<<

The Roles of the High and Low Molecular Weight Isoforms of 2

in Ischemia-Induced Revascularization

A dissertation submitted to the

Graduate School

of the University of Cincinnati

in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

in the Department of Pharmacology and Cell Biophysics

of the College of Medicine

by

Adeola Adeyemo B.S. University of Illinois at Chicago

March 2016

Committee Chair: Jo El J. Schultz, Ph.D.

ABSTRACT

Cardiovascular diseases are the underlying cause for majority (>30%) of deaths worldwide.

They coronary disease and peripheral disease, conditions characterized by limited

flow and inadequate oxygenation. Treatment strategies include management of symptoms

and risk factors, reduction of oxygen demand and surgical revascularization to increase circulation.

Therapeutic revascularization is a potential alternative for patients who are poor candidates for

these interventions due to advance disease or co-morbidities. Revascularization involves the

genetic or pharmacologic stimulation of vascular growth processes to facilitate tissue perfusion

and promote functional recovery. Adaptive vascular growth occurs via vessel growth

() and growth or remodeling of collateral (arteriogenesis). Delivery of angiogenic factors such as 2 (FGF2) to ischemia tissues can stimulate growth.

FGF2 consists of two classes of protein isoforms generated from alternative translation of the Fgf2 , high molecular weight (HMW) FGF2, and low molecular weight (LMW) FGF2.

Proof-of-concept studies in animal models of chronic ischemia provided evidence for the therapeutic potential of exogenous LMW FGF2. This promise, however, did not translate into successful clinical use. Currently, the functions of the endogenous FGF2 isoforms in ischemia- induced revascularization are not well understood. Elucidating the function(s) of the FGF2 isoforms in vascular growth is of great clinical importance and may lead to the development of novel pharmacological therapies for ischemic diseases.

Mice with a targeted deletion of all FGF2 isoforms (Fgf2-/-), HMW FGF2 (FGF2 LMW- only) and LMW FGF2 (FGF2 HMW-only) were employed to identify the distinct role(s) of the

ii

FGF2 isoforms in chronic ischemia. Revascularization was evaluated in mice subjected to chronic

hindlimb ischemia using measures of limb function, tissue viability, and vessel growth. FGF2

HMW-only mice had a faster recovery of limb function compared to wildtypes. Fgf2-/- and FGF2

LMW-only mice, however, had a significantly slower recovery of function. Additionally, no limb

necrosis was detected in FGF2 HMW-only mice while Fgf2-/- and FGF2 LMW-only mice had significant necrosis of their ischemic limbs. The early recovery of limb function in FGF2 HMW- only limbs was preceded by improved revascularization (angiogenesis and arteriogenesis). Vessel growth was significantly decreased in FGF2 LMW-only muscles and not different from wildtypes in Fgf2-/- mice, indicating that the presence of LMW FGF2 inhibited vascular growth. The effect of the HMW FGF2 isoforms on revascularization and protection from ischemic injury was not associated with activation of FGF receptor (FGFR). Angiogenesis-related proteins including

IGFBP-3, IGFBP-10 and CX3CL1 were increased in HMW FGF2-only muscles and are likely involved in the protective effect of HMW FGF2. Another possible mechanism mediating HMW

FGF2-induced recovery from ischemic damage is myocyte regeneration. Preliminary results indicate the presence of amplified satellite cell activation in HMW FGF2 muscles, which was represented by upregulation of Pax7 expression. Pax7 expression was coupled with increased expression of the myoblast differentiation , Myogenin and MRF4.

Together, these data, for the first time, show that the HMW FGF2 isoforms had a beneficial role in salvaging skeletal muscle from ischemic injury. This dissertation uncovered a biological role for HMW FGF2 isoforms in vascular growth. Though the mechanisms that mediate this function of HMW FGF2 isoforms remain to be thoroughly characterized, this dissertation provides significant and novel evidence for the role of the FGF2 HMW isoforms in ischemia-induced revascularization and preservation of muscle function.

iii

iv

ACKNOWLEDGMENTS

As I start to pen the final section of this thesis and with the final steps toward completion in sight, my prior belief that this part would be the easiest has turned out not to be true. So, it is with mixed emotions that I continue writing. I am both happy to have arrived at the end of this long journey and sad because of the strong attachments I feel to the people and the place that has nurtured my transition from student to scientist.

First and foremost, I am enormously indebted to my thesis advisor, Dr. Jo El Schultz for her unwavering support throughout my time as a graduate student. She always provided whatever

I needed to succeed including training, support, guidance, friendship and patience. I could always count on her to be available whenever I needed her, to be an attentive ear when I needed to vent as well a shoulder to cry on at times of frustration. She has and will continue to be a mentor, a role model and friend.

I would like to extend my gratitude to my dissertation committee, Drs. James Hoying, W.

Keith Jones, Ronald Millard, and Daria Narmoneva for their valuable time, assistance and encouragement. As a group, they have been an incredible resource for experimental advice, technical support and cardiovascular expertise.

I am also grateful to past and current members of the Schultz lab including Dr. Craig Bolte and Dr. Janet Manning for being amazing graduate student mentors; Dr. Yu Zhang for collaborating on some of my projects, Chahua Huang for being my lab buddy and teaching me a few words in Chinese and Colleen York for helping with our animal colony. Special thanks to honorary Schultz lab member, Stephanie Kelly for her insight into the treadmill experiment setup and her support from behind the scenes.

v

I am sincerely indebted to many colleagues and collaborators at the University of

Cincinnati and Cincinnati Children’s Hospital, including Drs. Swathi Balaji and Daria Narmoneva

for their assistance in developing the vessel staining protocol; Dr. Zhihua Qi, Dr. Lisa Lemen, and

Kati LaSance of the Vontz Core Imaging Lab for their guidance in the collection and analysis of

the micro-CT images; Dr. Craig Bolte and Robin Gear for constantly helping me to troubleshoot

experiments. In addition, members of the laboratories of Drs. Belcher, Fan, and Kranias,

who were all very generous with their reagents, lab equipment and time.

I am extremely grateful to the undergraduate and graduate students who I have had the

privilege to mentor during their rotations through the Schultz lab including Mykea Ruffin, Allyson

Hamlin, Allison Dixon, Jordan Redfield, Chang Zeng, Chris Johnson, and Andrew Stein. In

different ways, each of you made significant contributions to my research projects.

I owe a tremendous amount of gratitude to members of the Department of Pharmacology and the Molecular, Cellular, and Biochemical Pharmacology (MCBP) program. Drs. Scott

Belcher, Guochang Fan, Evangelia Kranias, Andrew Norman, Robert Rapaport, and Jo El Schultz for their steadfast belief in my capabilities as a scientist and helping to guide me towards my degree completion. A special thanks to Dr. Rapoport and Nancy Thyberg who towards the end were instrumental in keeping me on task, making sure paperwork was filed and deadlines were met.

Finally, I would like to thank family for their love, patience and support. To my wonderful parents for their encouragement, sacrifice, and unconditional support. I would like to express my profound appreciation for my brother Wunmi, and my sisters Tola and Toun, for always listening when I need to vent, never letting me wallow and always giving the best advice.

vi

TABLE OF CONTENTS Page

Abstract ii

Acknowledgments v

List of Figures and Tables xi

List of Abbreviations xv

Introduction

1. Coronary artery disease (CAD) 1

2. Peripheral artery disease (PAD) 2

3. Vascular growth 4

A. Therapeutic revascularization 4

B. Mechanisms of vascular formation and growth 4

Vasculogenesis 4

Angiogenesis 5

Arteriogenesis 6

Collateral remodeling 6

Vessel network formation 7

Ischemia-induced vascular growth 9

Mechanisms regulating post-ischemic vascular growth 10

4. Fibroblast Growth Factors (FGFs) 15

A. FGF receptors (FGFRs) 18

Heparan Sulfate Proteoglycans (HSPGs) 18

vii

FGFRs 19

B. FGF/FGFR signaling 22

C. Angiogenic FGFs 24

5. FGF2 28

A. Fgf2 gene and tissue expression 28

B. Phenotypes of Fgf2-/- mice 29

C. FGF2 protein isoforms 31

D. Biological functions of the FGF2 isoforms 34

E. Phenotypes of FGF2 isoform-specific knockout mice (role of endogenous FGF2

isoforms) 39

F. FGF2 in revascularization (Preclinical studies) 44

G. FGF2 in revascularization (Clinical studies) 56

6. Dissertation focus and hypotheses 55

Materials and Methods 64

Animals and exclusion criteria 64

Mouse generation and breeding 64

Generation of Fgf2-/- mice 64

Generation of FGF2 HMW-only mice 65

Generation of FGF2 LMW-only mice 68

Hindlimb ischemia surgery 69

Assessment of necrosis 71

Assessment of limb function 71

viii

Exercise treadmill testing 71

Fgf2 and Fgfr 72

RNA isolation 72

Reverse transcription and real-time polymerase chain reaction (PCR) 73

Myogenic satellite cell marker gene expression 75

FGF2 protein isoform expression (Western immunoblot) 78

FGFR expression and phosphorylation (Western immunoblot) 79

Expression of angiogenesis-related proteins 81

Immunohistochemistry 83

Griffonia simplicifolia (Bandeiraea) isolectin B4 83

Alpha smooth muscle actin 84

Myeloperoxidase (MPO) 85

Mac-3 (CD107b) 85

Micro-computed tomography (micro-CT) 86

Perfusion of contrast agent 87

Image acquisition 87

Comparison of contrast agents 88

Image analysis 93

Laser Doppler Perfusion Imaging 95

Statistical Analysis 95

Chapter 1: HMW and LMW isoforms of FGF2 in ischemia-induced revascularization and tissue recovery

Results I 96

ix

Discussion I 141

Chapter 2: Potential mechanisms of the FGF2 isoforms in adaptive revascularization during chronic ischemia

Results II 166

Discussion II 208

Chapter 3: FGF2 isoform-induced myogenesis during chronic hindlimb ischemia

Results III 233

Discussion III 248

Conclusions and Clinical Relevance 253

Future Directions 258

References 264

Appendix 345

x

LIST OF FIGURES AND TABLES

FIGURES

1. The mechanisms governing post-ischemic revascularization 8

2. Mouse/Human FGF2 gene and protein expression 34

3. LMW and HMW FGF2 isoform expression in non-ischemic skeletal muscles 67

4. The mouse hindlimb ischemia model 70

5. Representative 2-D slice views of a barium sulfate-perfused limb 90

6. Representative 2-D slice views of a lead oxide-perfused limb 91

7. Representative 2-D slice views of a Microfil-perfused limb 92

8. Representative 2-D slice views of a Microfil-perfused limb segmentation 94

9. Limb impairment scores 98

10. Treadmill exercise performance 100

11. Incidence of limb necrosis in ischemic limbs 103

12. Representative micro-CT images of sham and ischemic hindlimbs 108

13. Vessel volume, density and spacing calculated from 3-D micro-CT images 109

14. Histograms of vessel thickness distribution in sham and ischemic limbs analyzed from

micro-CT images 112

15. Volume of small-sized (34 -200μm) and medium-sized (200-400μm) vessels analyzed

from micro-CT images 115

16. Capillary vessel density ratio after 14 days of ischemia 118

17. Representative photomicrographs of GSI-Lectin stained vessels and capillary density in

sham and ischemic muscles after 42 days of ischemia 121

18. Capillary density of sham and ischemic muscles after 42 days of ischemia 122

xi

19. Gracilis, semi-membranosus, semi-tendinosus and gastrocnemius muscle capillary density

after 42 days of ischemia 123

20. vessel density ratio after 14 days of ischemia 126

21. Representative photomicrographs of α-SMA stained vessels, venule and arteriole density

in sham and ischemic muscles after 42 days of ischemia 129

22. Venue and arteriole vessel density of sham and ischemic muscles after 42 days of

ischemia 130

23. Gracilis, semi-membranosus, semi-tendinosus and gastrocnemius muscle venule density

in sham and ischemic muscles after 42 days of ischemia 132

24. Gracilis, semi-membranosus, semi-tendinosus and gastrocnemius muscle arteriole density

in sham and ischemic muscles after 42 days of ischemia 134

25. Wet weights of sham and ischemic gracilis, semi-membranosus, semi-tendinosus and

gastrocnemius muscles 136

26. Representative laser Doppler perfusion images and quantitative analysis of foot perfusion

over time 139

27. Schematic (Chapter 1) 165

28. Fgf2 mRNA expression in non-ischemic muscle 138

29. LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT, Fgf2-/-, and

FGF2 LMW-only non-ischemic muscles 167

30. LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT and FGF2

HMW-only non-ischemic muscles 169

31. Fgf2 mRNA expression in Fgf2-/- and FGF2 LMW-only sham and ischemic muscles170

32. Fgf2 mRNA expression in FGF2 HMW-only sham and ischemic skeletal muscles 173

xii

33. LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT, Fgf2-/-, and

FGF2 LMW-only sham and ischemic muscles 174

34. LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT and FGF2

HMW-only sham and ischemic skeletal muscles 176

35. Fgfr1 mRNA expression in non-ischemic, sham and ischemic muscles 180

36. Fgfr3 mRNA expression in non-ischemic, sham and ischemic muscles 183

37. Fgfr4 mRNA expression in non-ischemic, sham and ischemic muscles 186

38. FGFR activation (phosphorylation) and expression in non-ischemic muscle 189

39. FGFR1 activation (phosphorylation) and expression in ischemic muscles 191

40. FGFR3 activation (phosphorylation) and expression in ischemic muscles 193

41. FGFR4 activation (phosphorylation) and expression in ischemic muscles 195

42. Neutrophil density in non-ischemic, sham and ischemic muscles 203

43. density in non-ischemic, sham and ischemic muscles 207

44. Schematic (Chapter 2) 232

45. Schematic of satellite cell activation, differentiation and fusion 235

46. Pax7 mRNA expression in non-ischemic, sham and ischemic muscles 237

47. Myf5 mRNA expression in non-ischemic, sham and ischemic muscles 240

48. MyoD mRNA expression in non-ischemic, sham and ischemic muscles 241

49. Myogenin mRNA expression in non-ischemic, sham and ischemic muscles 243

50. MRF4 mRNA expression in non-ischemic, sham and ischemic muscles 246

51. Schematic (Chapter 3) 252

52. Working model of HMW FGF2 in vascular and skeletal muscle repair 257

xiii

TABLES

1. Phenotypes of Fgf knockout mice 17

2. FGF/FGFR binding specificities 21

3. Phenotypes of Fgfr knockout mice 22

4. FGF2 and other angiogenic growth factors in angiogenesis assays 47

5. Fgf2 gene or LMW FGF2 preclinical studies in models of myocardial ischemia 53

6. FGF2 preclinical studies in models of peripheral ischemia 54

7. FGF2 and other angiogenic growth factors in models of ischemia 55

8. Clinical trials results of FGFs in CAD and PAD patients 58

9. Forward and reverse primers for analysis of Fgf2 and Fgf receptor expression 76

10. Forward and reverse primers for the analysis of MRFs expression 77

11. List of angiogenesis-related proteins analyzed using antibody array 82

12. Descriptive scoring for the assessment of hindlimb use during chronic ischemia 97

13. Expression of angiogenic-related proteins in ischemic muscles 199

xiv

LIST OF ABREVIATIONS

AGENT angiogenic gene therapy

α-SMA alpha smooth muscle actin

Ang angiopoietin

ANOVA analysis of variance

ATE amino-terminal end

BAEC bovine aortic endothelial cell

BCEC bovine capillary endothelial cell

BM-MNC bone marrow mononuclear cell

CAD coronary artery disease

CAM chorioallantoic membrane

CHD coronary heart disease

CHIP chromatin immunoprecipitation

CLI critical limb ischemia

CNS central nervous system

CVD cardiovascular disease

CXCL chemokine (C-X-C) motif

DLL4 delta-like ligand 4

EC endothelial cell

ECM

EGF

EPC endothelial progenitor cell

ETT exercise tolerance test

xv

FGF fibroblast growth factor

FGFR fibroblast growth factor receptor

FIF FGF2-interacting factor

FIRST FGF Initiating RevaScularization Trial

GM-CSF granulocyte -colony stimulating factor

GSL IB4 griffonia simplicifolia Isolectin B4

HB-EGF heparin-binding EGF-like growth factor

HGF growth factor

HIF1α hypoxia inducible factor 1 alpha

HLI hindlimb ischemia

HMW high molecular weight

HSPG proteoglycan

IC intermittent claudication

IGF1 insulin like growth factor 1

IGFBP insulin-like growth factor binding protein

IL interleukin

INFS integrative nuclear FGFR1 signaling

I/R ischemic/reperfusion kDa kilodaltons

LAD left anterior descending

LDPI laser Doppler perfusion imaging

LMW low molecular weight

MAPK mitogen activated protein kinase

xvi

Micro-CT micro computed tomography

MIP-1 macrophage inflammatory protein-1

MCP-1 monocyte chemoattractant protein-1

MI

MMP matrix metalloproteinase

MPO myeloperoxidase

MRF myogenic regulatory factor

MyoD myogenic determination factor 1

Myf5 myogenic factor 5

NV1FGF nonviral 1 FGF

PAD peripheral artery disease

PAI-1 plasminogen activator inhibitor-1

Pax7 paired box 7

PBS phospho-buffered saline

PCR polymerase chain reaction

PEDF pigment epithelium-derived factor

PDGF platelet-derived growth factor

PD-ECGF platelet-derived endothelial cell growth factor

PIGF growth factor

PI3K phosphatidylinositol 3-kinase

PDGF platelet-derived growth factor

PKC protein kinase C qRT-PCR quantitative real time PCR

xvii

ROI region of interest

SDF stromal-derived factor

SMN survivor of motorneuron protein

TALISMAN Therapeutic Angiogenesis Leg Ischemia Study for the Management of

Arteriopathy and Nonhealing ulcers

TIMP tissue inhibitors of metalloproteinases

TNF-α -alpha

TRAFFIC TheRapeutic Angiogenesis with recombinant FGF2 for Intermittent

Claudication u-PA urokinase plasminogen activator

VEGF vascular endothelial growth factor vSMC vascular smooth muscle cell

VOI volume of interest

WT wildtype

xviii

INTRODUCTION AND BACKGROUND

Approximately 85.6 million individuals in the United States have one or more form of cardiovascular disease (CVD) (Mozaffarian et al., 2015). As recently as 2010, CVDs accounted for over 31% of all U.S deaths. CVD as defined by the International Classification of Diseases includes coronary heart disease (CHD), acute coronary syndrome, high blood pressure, myocardial infarction, angina pectoris, heart failure and stroke (cerebrovascular disease) and peripheral vascular (arterial) disease (Brämer, 1988). This dissertation will focused on coronary heart disease and peripheral vascular disease.

1. Coronary heart disease (CHD)

Coronary heart disease affects over 15.5 million people in the United States. This disease category includes acute myocardial infarction, acute ischemic (coronary) heart disease, angina pectoris, atherosclerotic cardiovascular disease, and chronic ischemic (coronary) heart disease

(Mozaffarian et al., 2015). CHD develops from hardening, narrowing and/or occlusion of one or more of the large coronary arteries of the heart that is caused by atherosclerotic plaque (obstructive lesions) formation. Atherosclerotic lesions are usually an accumulation of lipids, inflammatory cells and fibrous material that are deposited in the inner arterial walls (Falk, Nakano, Bentzon,

Finn, & Virmani, 2013; Libby & Theroux, 2005). In CHD, fixed stenosis of coronary arteries manifests as stable angina pectoris and patients experience reversible ischemic pain during exercise. This type of angina can progress to an unstable state where ischemic pain occurs at rest

(Dragneva, Korpisalo, & Ylä-Herttuala, 2013). Additional risk factors for CHD include cigarette smoking and alcohol use, physical inactivity, hypertension, obesity, mellitus and aging

(Silvestre, Smadja, & Lévy, 2013). Current therapies for chronic ischemic and coronary heart

1

diseases include controlling risk factors (diet, exercise and cessation of smoking), preventive

measures (statins, aspirin, antiplatelets and ACE inhibitors), revascularization procedures

(percutaneous coronary interventions, stents or bypass grafts) and anti-thrombotic agents (Libby

& Theroux, 2005; Rubanyi, 2013).

2. Peripheral arterial disease (PAD)

Peripheral arterial disease (PAD) results in reduction of blood supply to vessels beds other than the heart or the brain. PAD affects an estimated 8.5 million people in the U.S and it has been diagnosed in over 10-15% of adults over age 65 (Mozaffarian et al., 2015). PAD develops from occlusive and is characterized by loss of blood flow reserve from chronic reduction in the diameter of the lumen of the distal abdominal , pelvic or lower extremity arteries

(Annex, 2013). There are two symptomatic clinical manifestations of PAD; intermittent claudication (IC) and critical limb ischemia (CLI). In IC, patients encounter muscle pain or cramping during walking and exertion which is relieved upon resting, while CLI presents as limited blood flow to the limbs even at rest that can lead to gangrene, ulcers and in some cases amputation (Haas, Lloyd, Yang, & Terjung, 2012). The risk factors for PAD are similar to those for CHD; however, cigarette smoking and diabetes mellitus play a stronger causative role in PAD.

Other risk factors for PAD include hypertension, hyperhomocysteinemia, dyslipidemia, elevated

C-reactive protein, and age (Mascarenhas, Albayati, Shearman, & Jude, 2014; Muller, Reed,

Leuenberger, & Sinoway, 2013). PAD alone can be a strong predictor of CHD and cerebrovascular disease (CBVD) based on vascular effects. The co-diagnosis of CHD with PAD ranges from 35% to 60% of patients and for PAD with CBVD can be as high as 80% of patients. The severity of

PAD correlates with increased CVD-related mortality (Haas et al., 2012). Current treatments for

2

PAD are similar to those for CHD and include modification of risk factors like smoking, diet and

exercise (Muller et al., 2013). Smoking cessation slows the progression to critical limb ischemia

and reduces the risk of cardiovascular (CVD) events like myocardial infarction (Jha et al., 2013).

Diet modification and management of diabetes also decreases the risk of CVD events as well as

microvascular complications like retinopathies and neuropathies (Libby & Theroux, 2005; Muir,

2009; Muller et al., 2013). Statins also reduce risk and worsening of symptoms in PAD patients

by lowering serum cholesterol, improving endothelial function, and other atherosclerotic markers.

Antihypertensives like ACE inhibitors or β-AR blockers lower blood pressure and reduce the risk of new CVD events. Secondary prevention of CHD and PAD is conferred by antiplatelet agents which reduce the risks of nonfatal myocardial infarction, ischemic stroke, and vascular-related death (Muir, 2009; Muller et al., 2013; Norgren et al., 2007).

Overall, standard treatments for CHD and PAD are aimed at symptom relief, reducing atherosclerotic risk factors, decreasing oxygen demand and vascular interventions to physically reestablish blood flow to the ischemic regions. Despite these pharmacotherapies, a large and growing number of patients remain symptomatic. The presence of co-morbidities like diabetes and hypertension renders these patients ineligible for the surgical interventions; these patients may eventually develop heart failure or require limb amputations (Ahn, Frishman, Gutwein, Passeri, &

Nelson, 2008; Grochot-Przeczek, Dulak, & Jozkowicz, 2013; Rubanyi, 2013; Van Weel, van

Tongeren, van Hinsbergh, van Bockel, & Quax, 2008). The increasing numbers of these “no- option” patients illustrates the need for alternative treatment approaches for CAD and PAD

(Rubanyi, 2013; van Weel et al., 2008; Annex 2013 ).

3

3. Vascular growth

A. Therapeutic revascularization

“Therapeutic revascularization” is an emerging strategy that aims to augment or stimulate

blood vessel growth and remodeling for the improvement of perfusion and function in the ischemic

heart or lower limb (Annex, 2013; Rubanyi, 2013). This approach compensates for the loss of

blood supply that develops from arterial occlusions or stenosis by pharmacological treatment with

(s) of vessel growth.

B. Mechanisms of vascular formation and growth

Blood vessel growth and formation involves four distinct but complementary processes: 1)

vasculogenesis, 2) angiogenesis, 3) arteriogenesis, and 4) collateral growth (Annex, 2013;

Dragneva et al., 2013; Rubanyi, 2013; Silvestre et al., 2013). Vessel growth plays a key role in various physiological and pathological conditions, including the menstrual cycle, embryonic development, wound repair, , and tumor growth. At present, it is known that all four mechanisms of vascular growth can be triggered in response to ischemia but the relative contributions of each to the restoration of perfusion has yet to be determined (Figure 1).

Vasculogenesis

Vasculogenesis is the de novo formation of the earliest vascular structures (primitive plexus) in the developing embryo. In situ proliferation, migration and differentiation of endothelial progenitor cells (EPCs) (angioblasts in development) or bone-marrow-derived endothelial stem cells (in adult) results in tubules of endothelial cells (ECs) with very little organization (Carmeliet

& Jain, 2011; Rubanyi, 2013). Once considered to be a strictly embryonic process, vasculogenesis in the adult has been shown to occur when nascent vessels formed in response to ischemia were

4

proven to contain progenitor cells that mobilize to the site of and are incorporated into the developing vessels (Asahara, 1997; Fischer, Schneider, & Carmeliet, 2006).

Angiogenesis

“Angiogenesis” is frequently used to describe de novo formation or increase in diameter, length, or branching of vessels in any developmental, postnatal or pathological condition (Faber,

Chilian, Deindl, Royen, & Simons, 2014). The use of “angiogenesis” in this dissertation will refer only to capillary growth or formation occurring during embryogenesis or in response to arterial obstructions as described below.

The organization of the nascent primitive plexus formed during vasculogenesis into a network of functional microvessels is termed embryonic angiogenesis. This process involves the sprouting or intussusceptive (splitting into daughter vessels) growth of new vessels from the ends

and walls of the preexisting vascular beds (endothelial tubules) (Carmeliet & Jain, 2011; Fischer

et al., 2006). This expanded nascent network is now made up of narrow microvessels (<10 µm)

which are referred to as . A variation of this process that occurs postnatally or in the

adult also involves the proliferation of pre-existing vascular endothelial cells into new capillary

sprouts (sprouting) and/or bridging of preexisting vessels into smaller daughter vessels (splitting)

(Fischer et al., 2006; Semenza, 2007). Angiogenesis can be induced by physiological or pathological triggers including organ growth, physical exercise training, , tumor growth, and vascular occlusions (Carmeliet & Jain, 2011; Fischer et al., 2006; Persson, 2011).

Arteriogenesis

5

Arteriogenesis refers to the stabilization and maturation of nascent vascular networks

(Fischer et al., 2006). Mural cells ( and smooth muscle cells) of mesenchymal lineage are recruited to the naked EC sprouts. Deposition of the extracellular matrix and tightening of cell junctions occurs at this time, leading to a more organized and functional network. The origin, type and composition of the mural cells will depend on the site, function and arterial-venous identity of the vessels (Carmeliet & Jain, 2011; Silvestre et al., 2013). A collateral circulation is also formed at this time. Collaterals are artery-to-artery (collateral arteries) or arteriole-to-arteriole

(microvascular collaterals) anastomoses (connections) that form between feed arteries or arterial trees (Faber et al., 2014). They exist in a unique hemodynamic state of very low blood flow and high resistance during normal physiology and are present in the brain, spinal cord, heart, skin and skeletal muscle (Faber et al., 2014; Fischer et al., 2006; J. Liu et al., 2014; Silvestre et al., 2013).

Collateral remodeling

This process often described as an adaptive arteriogenic process, refers to the transformation of pre-existing collateral vessels with previously low blood flow into functional conductance vessels (Faber et al., 2014). Here, remodeling is defined as an anatomic enlargement of lumen diameter and an increase in vascular wall thickness (Heil, Eitenmüller, Schmitz-Rixen,

& Schaper, 2006; Silvestre et al., 2013). This outward remodeling takes place as a result of flow obstructions that induce increased flow in adjacent arterial trees. The pre-existing collateral network across from the occluded vascular tree is recruited to compensate for the change in pressure gradient (J. Liu et al., 2014; Schaper, 2009). The microvascular collaterals are typically

≤100 µm in diameter (in healthy tissues) and are capable of significant outward remodeling by

6

increasing their lumen diameter up to 5-10 fold. In the mouse circulation, they can be as small as

30 µm and can increase by 2-5 fold (Faber et al., 2014; Ziegler et al., 2010).

Vessel network formation

During physiological angiogenesis, formation of hierarchally structured three dimensional vascular networks requires more than sprouting of neovessels. Guided maturation, pruning,

regression of the blood vessel network is necessary for the formation of a stable, well-perfused

network (LeBlanc, Krishnan, Sullivan, Williams, & Hoying, 2012). Vessel maturation involves the recruitment and attachment of network-stabilizing perivascular cells like pericytes and vascular smooth muscle cells. The recruitment and differentiation of perivascular cells is mediated by factors released by endothelial cells including platelet endothelial growth factor (PDGF) and transforming growth factor-β (TGF-β).

Capillary regression is regulated by the balance between the expression of positive angiogenic factors like vascular endothelial factors (VEGF) and negative angiogenic factors such as -1 (TSP-1) (Olfert, 2015). This balance is evident in the mammalian corpus luteum. Growth and regression of the cyclic corpus luteum involves distinct phases of angiogenesis, vessel maturation, and vessel regression. Luteolysis leads to the physiologically triggered dissociation of the entire corpus luteum vasculature (Wietecha, Cerny, & DiPietro,

2013). Branch selection for pruning or regression is also regulated by blood flow and perfusion of vessels. The selected branch/vessel constricts until it occludes and the blood flow ceases. The endothelial cells then retract to undergo or may migrate away to be re-integrated elsewhere.

7

Figure 1: The four distinct but complementary mechanisms governing post-ischemic revascularization. 1) Vasculogenesis, the formation of blood vessels by differentiation of endothelial progenitors cells (EPCs); 2) angiogenesis, the sprouting or budding of new capillaries (Cap) from preexisting vessels; 3) arteriogenesis, the maturation and stabilization of nascent sprouts via recruited smooth muscle cells (SMC) and pericytes (Per) to form ; and 4) collateral (Coll) growth, the remodeling of preexisting arteriolar anastomoses into functional conductance vessels. Adapted from (Silvestre et al., 2013). Reprinted with permission of the American Physiological Society.

8

Ischemia-induced vascular growth

The vascular processes described above largely take place during embryonic development but they can be recapitulated in adult tissues under certain pathophysiological conditions like ischemia or physical damage (Ahn et al., 2008; Carmeliet & Jain, 2011; Haas et al., 2012; Silvestre et al., 2013). The physiological responses of cardiac or skeletal muscle to the initial vascular obstruction share some similarities with the wound healing processes observed in tissues like skin.

Tissue healing can be been divided into three chronologically overlapping phases: 1) destruction, 2) inflammatory phase, and 3) remodeling and regeneration phase (Philippou,

Maridaki, Theos, & Koutsilieris, 2012; Smith, Kruger, Smith, & Myburgh, 2008; Weber, Sun,

Bhattacharya, Ahokas, & Gerling, 2013). The destruction phase which begins just after the initial occlusion and persists for up to 2 days is characterized by hypoperfusion, hypoxia and metabolic dysregulation leading to myocyte death by apoptosis and/or necrosis. An acute inflammatory phase follows cellular death and can last for up to 7 days (Silvestre et al., 2013). This phase features degradation of the extracellular matrix to allow homing and migration of inflammatory cells

(neutrophils, , ) into the necrotic tissue for phagocytosis of cellular debris

(Cleutjens, 1999; Matsui, Morimoto, & Uede, 2010). The remodeling and regeneration phase which lasts up to 21 days is comprised of migration and differentiation of , , and ; synthesis and deposition of extracellular matrix proteins; and scar tissue formation to replace necrotic or apoptotic cells. The final phase is also associated with an immune-inflammatory balance, secretion of chemokines, growth factors and other which promote myocyte hypertrophy/regeneration and vessel formation or remodeling (Silvestre et al., 2013; Smith et al., 2008).

9

Mechanisms regulating post-ischemic vascular growth

There are many driving forces that initiate adaptive angiogenesis and collateral growth;

they include inflammatory balance, ischemia-induced hypoxia, and fluid shear stress (Heil et al.,

2006; Silvestre et al., 2013).

i. Inflammatory response in post-ischemic vascular growth

Acute inflammation is a crucial component of the repair/remodeling phase of the tissue

healing that occurs after ischemia (Silvestre et al., 2013; Smith et al., 2008). Vessel growth and

inflammation are linked in pathophysiological conditions of wound healing, diabetic retinopathy,

, and atherosclerosis (Frantz, Vincent, Feron, & Kelly, 2005; Mueller, 2008;

Serhan & Savill, 2005). Inflammation is capable of modulating vessel growth and vice-versa. The when it is activated by hypoxia secretes chemoattractant cytokines and adhesion molecules that serve as signals for inflammatory cell recruitment and activation (Stockmann,

Schadendorf, Klose, & Helfrich, 2014). In turn, the recruited inflammatory cells (leukocytes) stimulate the formation of new vessels in a paracrine fashion by secreting angiogenic growth factors, pro-and antiangiogenic cytokines and proteases (Smith et al., 2008). The recruitment and

transmigration of immune cells across the sites of ischemia and vascular occlusion is regulated by the expression and surface presentation of specific and chemotactic pathways. These include the diverse group of cell adhesion molecule, chemokine and interleukin protein families.

The types of inflammatory cells that found infiltrate the vasculature and muscle during

ischemic injury include monocytes, neutrophils, and lymphocytes (Cochain, Channon, & Silvestre,

2012; Mueller, 2008; Smith et al., 2008). The level of monocyte infiltration has been shown to

correlate with the extent of neovascularization. Monocytes are rapidly mobilized from the bone

10

marrow, spleen and bloodstream to the site of ischemia (Silvestre et al., 2013). Monocytes

differentiate into macrophages once they have extravasated into the tissue. Tissue macrophages

are made up of two subpopulations of cells: the classically-activated M1 macrophages which

express pro-inflammatory molecules and alternatively-activated M2 macrophages which express anti-inflammatory molecules and promotes vascular and tissue regeneration (Fung & Helisch,

2012; Jetten et al., 2014; Mueller, 2008; Silvestre et al., 2013). The phagocytic M1 cells are localized to inflamed tissues immediately after ischemia to promote phagocytosis and the M2 cells are largely present during tissue healing (Silvestre et al., 2013; Smith et al., 2008).

Neutrophils, unlike monocytes, are predominantly antiangiogenic (Naldini & Carraro,

2005; Nguyen & Tidball, 2003). Since they trigger the release of free radicals and lysosomal proteinases, neutrophils have been linked to the secondary tissue damage that occurs after ischemia especially to the endothelium (Frangogiannis, Smith, & Entman, 2002; Silvestre, Mallat, Tedgui,

& Lévy, 2008; Silvestre et al., 2013). They produce reactive oxygen species which can stimulate revascularization at low levels but can also damage endothelial cells and inhibit angiogenesis at high concentrations (Cochain et al., 2012; Silvestre et al., 2008; Sun, 2009). Lysosomal proteinases such as proteinase 3 and elastase are also generated by neutrophils. These proteinases can induce endothelial detachment and apoptosis (Cochain et al., 2012; Mueller, 2008; Silvestre et al., 2008).

However, neutrophils may contribute to angiogenesis by their secretion of vascular endothelial

growth factor (VEGF) (Cochain et al., 2012; Mueller, 2008).

T-lymphocytes, divided into two subtypes, CD4 or CD8, based on their expression of

membrane antigens, are also involved in post-ischemic inflammatory response (Silvestre et al.,

2013). Mice that are deficient in CD4 or CD8 (CD4-/- or CD8-/-) or both (CD4-/-/CD8-/-) have

impaired vascular responses to ischemia (Stabile et al., 2003, 2006). CD8 T-lymphocytes promote

11

vascularization by the release of growth factors and cytokines including VEGF, interleukins 10

and 16 (IL-10 and IL-16) that will also attract other inflammatory cells like CD4 lymphocytes and

monocytes (la Sala, Pontecorvo, Agresta, Rosano, & Stabile, 2012; Silvestre et al., 2013).

The overall role of inflammation on ischemia-induced revascularization is dependent on

the temporal and spatial balance between the activation of pro- and anti-inflammatory cells.

Additionally, the intensity of pro-/anti-angiogenic molecules secreted by the recruited immune

cells is also important for vascular remodeling (Carmeliet, 2003; Silvestre et al., 2008).

ii. Hypoxia in ischemia-induced angiogenesis

The primary functions of the vascular system include the delivery of O2 and supply of

nutrients to tissues and the removal of toxic metabolic wastes (Semenza, 2007). The reduction or

complete loss of blood supply that occurs in ischemic tissues disrupts this oxygen homeostasis,

and leads to hypoxia. Hypoxia is the local decrease in O2 partial pressure (pO2) that results from

either decreased supply of O2 or increased metabolic demand from tissues (Silvestre et al., 2013).

In hypoxic environments, angiogenesis is stimulated to re-establish the equilibrium between local oxygen need and supply and is mediated by the hypoxia-inducible factor (HIF) (Krock, Skuli, &

Simon, 2011). HIF is a heterodimeric protein composed of a regulatory (oxygen-sensitive) α- subunit and a constitutively expressed β-subunit. There are three isoforms of the HIF α-subunit

(1α, 2α and 3α) that are subjected to rapid polyubiquitination and proteosomal degradation under normal oxygen conditions. HIF1α is universally expressed while HIF2α is predominantly expressed in endothelial cells during development and in the adult (Hickey & Simon, 2006; Krock et al., 2011; Silvestre et al., 2013). HIF3α has been detected in the cerebral cortex, hippocampus and epithelial cells (Heidbreder et al., 2003; Q. F. Li, Wang, Yang, & Lin, 2006). During

hypoxic stress when tissue pO2 decreases below 5% (40 mmHg), the HIF1α subunit becomes

12

stabilized and translocates to the nucleus (Krock et al., 2011; Persson, 2011). After dimerization

with the β-subunit, HIF1 binds to hypoxia response elements within promoter regions of target

genes and activates gene transcription (Semenza, 2007). Pathways targeted by the active HIF1

include other transcription factors (NFκB and ETS1), vasodilatory nitric oxide (NO), glycolytic

enzymes, cell cycle inhibitors and angiogenic cytokines (Krock et al., 2011).

Angiogenic cytokines induced by hypoxia and HIF1 exhibit either pro- or anti-angiogenic

behavior. Endogenous and ischemia-induced vascular growth are both tightly regulated and are

governed by the balance between local expression of angiostatic and proangiogenic molecules

(Hickey & Simon, 2006; Krock et al., 2011). Some of these proangiogenic genes include

angiopoietin-1/2 (Ang-1/2), matrix metalloproteinases (MMPs), tissue inhibitor of

metalloproteinases (TIMP), tyrosine kinase with immunoglobulin-like and EGF-like domains 1

(Tie2), tumor necrosis factor-α (TNF-α), interleukins, chemokines, stromal-derived growth factor

(SDF), (HGF), placenta growth factor (PIGF), platelet-derived growth

factors (PDGFs, vascular endothelial growth factors (VEGFs), and fibroblast growth factors

(FGFs) (Krock et al., 2011; Murakami & Simons, 2008; Persson, 2011; Rubanyi, 2013). This dissertation will focus on the biological functions of fibroblast growth factor 2 (FGF2) in ischemia- induced angiogenesis.

iii. Fluid shear stress-induced collateral remodeling

Unlike ischemia-induced angiogenesis, collateral remodeling has been shown to be largely stimulated by inflammation and is not likely to involve hypoxia (Heil et al., 2006; Heil & Schaper,

2007). Evidence for this includes the lack of HIF1α upregulation in the collateral zone, localization of collaterals in tissue regions that are distal to the ischemic zone, and the occurrence of collateral

13

development much later than the appearance of capillary growth in the ischemic regions (Heil et

al., 2006; van Oostrom, van Oostrom, Quax, Verhaar, & Hoefer, 2008; Van Weel et al., 2008).

During arterial occlusion, the luminal surface of vessels are exposed to hemodynamic

forces such as tensional stretch and shear stress which have been shown to induce vessel growth

by activation of endothelium (Schirmer, van Nooijen, Piek, & van Royen, 2009; Van Weel et al.,

2008). In particular, collateral remodeling of pre-existing arterioles is triggered by an increase in

fluid shear stress (FSS) and circumferential wall stress. FSS results from the steep pressure

gradient between the preexisting arterioles (high pressure) connected upstream of the occlusive

region and those downstream (low pressure) from this point (van Oostrom et al., 2008). The

increased flow through the vessels leads to functional activation of the previously quiescent

endothelial cell layer. The shear stress is also transmitted to the smooth muscle cell layers and the

cytoskeleton via communication between the cells and with extracellular matrix proteins

(Semenza, 2007; Silvestre et al., 2013). The shift in phenotype from quiescent to “activated”

endothelial cells is followed by gene transcription of several genes which have shear stress

response elements (SSREs) in their promoters; these include , chemokines, endothelial

adhesion molecules, and growth factors (Haas et al., 2012; Heil & Schaper, 2007). The

upregulation of chemokines and cytokines induces the attraction and adhesion of bone marrow-

derived and circulating leukocytes particularly monocytes (Van Weel et al., 2008). Monocytes

accumulate in the perivascular space and then transmigrate across the vessels walls, by secretion

of MMPs and other protease that create gaps within the elastic lamina and basal membrane.

Monocytes also secrete additional growth factors like FGF2 that lead endothelial and smooth

muscle cells of the vessel wall to undergo a phenotypic shift from a contractile to a

proliferative/secretory state. The enzymatic breakdown of the extracellular matrix and basal

14

lamina also releases matrix-bound growth factors. Vascular cells proliferate, migrate and rearrange

causing thickening of the intima layer of the vascular wall (Fung & Helisch, 2012; Haas et al.,

2012; Heil & Schaper, 2004).

4. Fibroblast Growth Factors (FGFs)

Fibroblast growth factor 2 (FGF2) is a classical member of the FGF family. The FGFs are signaling polypeptides that are conserved among vertebrates and play an important role in a multitude of many processes including development, organogenesis, metabolism, neurobiology, tumorigenesis, and tissue homeostasis (Beenken & Mohammadi, 2009; Itoh & Ornitz, 2011). The

22 structurally-related proteins that make up the mammalian FGF family share 30-60% identity and contain ~150-300 amino acids. Crystal structures of the FGFs have revealed a core domain of

~140 amino acids with a conserved β-trefoil fold that consists of 12 anti-parallel β-strands (Belov

& Mohammadi, 2013; Zhu et al., 1991). The human FGF gene family has been grouped into seven subfamilies based on their and phylogeny: Fgf1 subfamily (Fgf1/2), Fgf4 subfamily (Fgf4/5/6), Fgf7 subfamily (3/7/10/22), Fgf8 subfamily (Fgf8/17/18), Fgf9 subfamily

(Fgf9/16/20), Fgf19 subfamily (19/21/23), and the Fgf11 subfamily (Fgf11/12/13/14) (Belov &

Mohammadi, 2013; Itoh & Ohta, 2013).

The FGFs can be classified as intracrine, endocrine or paracrine/autocrine factors based on their respective mechanisms of action. The intracrine FGFs (FGF11-FGF14) are generally referred to as FGF homologous factors (FHF) because they share high sequence homology with other FGFs but do not activate the FGF receptors (FGFR) (Belov & Mohammadi, 2013; Goldfarb, 2005). FHFs have been shown to interact with the intracellular domains of calcium channels and voltage-gated sodium channels (Hennessey, Wei, & Pitt, 2013; Laezza et al., 2007; Xu Zhang, Bao, Yang, Wu,

15

& Li, 2012). The endocrine FGF 15, 19, 21 and 23 are hormone-like in their actions, FGFR- dependent and regulate several metabolic pathways (Beenken & Mohammadi, 2009; Krejci,

Prochazkova, Bryja, Kozubik, & Wilcox, 2009; Laestander & Engström, 2014). Endocrine FGFs

(FGF 15, 19, 21, 23) also require additional co-receptors, the Klotho receptors (α or β) to form

active signaling complexes due to their extremely low affinities for heparan sulfates (Beenken &

Mohammadi, 2009; Belov & Mohammadi, 2013; Laestander & Engström, 2014). Fgf1(FGF1/2),

4(FGF4/5/6), 7(3/7/10/22), 8(8/17/18), and 9(9/16/20) subfamilies are the paracrine and/or

autocrine signaling FGFs. They are mostly secreted proteins that bind to and activate the cell

surface tyrosine kinase FGFRs with high affinity and with low affinity to heparan sulfate

proteoglycans (HSPGs) (Beenken & Mohammadi, 2009; Itoh & Ohta, 2013; Itoh & Ornitz, 2011).

FGF 1 and 2 are prototypes of the entire FGF family as well as the paracrine subgroup despite

lacking the classical N-terminal secretion present on the other paracrine FGFs (Itoh

& Ohta, 2013; Powers, McLeskey, & Wellstein, 2000). Elucidation of the biological functions of

the individual Fgf genes using knockout mice have begun and are listed in Table 1.

16

Gene Knockout mouse phenotype Physiological role

Fgf1 Normal Not established Viable, Loss of vascular tone, Slight loss of cortex Described in detail in section titled Fgf2 neuron, Reduced bone mass (Montero et al., 2000) “Phenotypes of Fgf2-/- mice” Inner ear development (agenesis) in humans Fgf3 Viable, Tail defects (Alvarez et al., 2003) (Tekin et al., 2008)

Embryonic lethal at E4-5 (Feldman, Poueymirou, Cardiac valve leaflet formation, Fgf4 Papaioannou, DeChiara, & Goldfarb, 1995) Limb development Abnormally long hair (Hébert, Rosenquist, Götz, & Fgf5 Martin, 1994) Hair growth cycle regulation Fgf6 Defective muscle regeneration Myogenesis

Fgf7 Matted hair, Reduced nephron branching in Branching morphogenesis Fgf8 Embryonic lethal at E8 Brain, eye, ear and limb development Postnatal death, Gender reversal, Gonadal development and Fgf9 Lung hypoplasia Organogenesis

Fgf10 Failed limb and lung development Branching morphogenesis Fgf11 N/A Not established Neurotransmission of motor function (Xu Fgf12 Viable, Muscle weakness Zhang et al., 2012) Microtubule stabilization (Q.-F. Wu et al., 2012) X-linked mental retardation Fgf13 Impaired learning and memory (Xu Zhang et al., 2012) Neurotransmission of motor function (Xu Fgf14 Ataxia, paroxysmal hyperkinetic movement disorder Zhang et al., 2012) Fgf16 Embryonic lethal Heart development Fgf17 Abnormal brain development Cerebral and cerebellar development Fgf18 Delayed long-bone ossification, Lethal at PD0 Bone development Bile acid homeostasis, Lipolysis, Gall Fgf15/19 Increased bile acid pool, Lethal at E13.5-PD7 bladder filling Fgf20 Viable Neurotrophic factor Fasting response, Glucose homeostasis, Fgf21 Viable Lipolysis and lipogenesis Fgf22 Viable Presynaptic neural organizer Impaired phosphate and Vitamin D metabolism, Phosphate homeostasis, Vitamin D Fgf23 Immature sexual organs, Lethal PW4-13 homeostasis

Table 1: Biological functions of the FGF family. Adapted from (Beenken & Mohammadi,

2009). E: embryonic day; PD: postnatal day; PW: post-wean day.

17

A. FGF Receptors

The paracrine and endocrine FGFs mediate diverse cellular response by interactions with

the low affinity heparan sulfate proteoglycan (HSPG) co-receptors and high affinity FGF receptors

(FGFRs). i. Heparan Sulfate Proteoglycans (HSPGs)

Heparan sulfate proteoglycans (HSPGs) are heavily glycosylated proteins with a core

protein and one or more covalently attached heparin/heparan sulfate glycosaminoglycan (GAGs)

chains (Bishop, Schuksz, & Esko, 2007). GAGs are linear polysaccharides consisting of repeating

N-acetylated or N-sulfated glucosamine units (N-acetylglucosamine or N-sulfoglucosamine) and

uronic acids (glucuronic acid or iduronic acid). HSPGs are classified into subfamilies based on

their cellular locations: membrane-spanning HSPGs (syndecans 1-4, betaglycan, neuropilin-1,

CD44v3), glycophosphatidylinositol-anchored HSPGSs (glypicans 1-6), ECM proteoglycans

(perlecan, agrin, type XVIII collagen) and secretory vesicle HSPGs (serglycin) (Sarrazin,

Lamanna, & Esko, 2011). The functions of HSPGs in cell physiology include acting as co-

receptors for growth factors and their tyrosine kinase receptors, serving as endocytic receptors for

the removal of bound ligands, protecting growth factors, cytokines, and chemokines from

proteolysis, and cooperating with integrins to support cell adhesion to ECM and cell motility

(Bishop et al., 2007; Sarrazin et al., 2011). Some HSPGs like members of the syndecan family are

able to function outside the FGF-FGFR signaling complex to activate signaling pathways in

response to growth factor ligand binding (Murakami, Elfenbein, & Simons, 2008). Syndecan-4

independently regulates activation and localization of protein kinase Cα (PKCα) at sites of focal

adhesion (Elfenbein & Simons, 2013; Keum et al., 2004).

18

HSPGs are abundantly expressed on cell surfaces and in basement membranes of almost

all cell types including endothelial cells and fibroblasts (Presta et al., 2005; Sarrazin et al., 2011).

HSPGs bind to a variety of ligands via the heparan sulfate chains: these include growth factors,

cytokines, enzymes, chemokines and ECM proteins. They function in regulating FGF binding to

and activation of FGF receptors (Moosa Mohammadi, Olsen, & Ibrahimi, 2005). HSPGs are co- receptors that facilitates proper presentation of FGFs to FGFR and increases the formation of stable

FGF/FGFR complexes as well as the half-life of the signaling complex (Belov & Mohammadi,

2013; Bishop et al., 2007; Presta et al., 2005). HSPGs of the ECM can act as reservoirs for FGFs, prevent proteolytic degradation and increase local gradients of FGF during stimulation of endothelial cells. Alternatively, free heparin/HS can trap or sequester FGFs in the blood and other extracellular spaces and inhibit FGF activity (Presta et al., 2005, 2007). The affinity of FGFs and other growth factors for the highly negatively-charged heparin/heparan sulfates has been crucial in the discovery and purification of these proteins and is still in use today in affinity matrices and other “pull-down” assays.

ii. Fibroblast growth factor receptors (FGFRs)

The FGF receptors are a group of tyrosine kinase receptors that are encoded by four genes;

FGFR1, FGFR2, FGFR3 and FGFR4. They share up to 70% sequence identity (Eswarakumar,

Lax, & Schlessinger, 2005). These highly conserved receptors have the classic tyrosine kinase structure with an extracellular ligand-binding domain, a single helical transmembrane domain and a split intracellular tyrosine kinase domain (Belov & Mohammadi, 2013; Eswarakumar et al.,

2005). The extracellular domain is composed of three extracellular immunoglobulin (Ig) domains

(I, II and III) and an unique eight acidic residue sequence in the linker connecting domains I and

II called the “acid box”. The domain II and III fragments are necessary for ligand and heparin

19

binding and IgI and the acid box may both be important for autoinhibition (Beenken &

Mohammadi, 2009). Alternative RNA splicing of Ig domain III in the FGFR1-3 genes yields

several isoforms; FGFR1-3IIIa, FGFR1-3IIIb and FGFR1-3IIIc and this determines ligand specificity and tissue expression (Turner & Grose, 2010)(Belov & Mohammadi, 2013). The IIIa

forms have no known signaling capability while the IIIb and IIIc forms are predominantly

expressed in epithelial and mesenchymal cells, respectively (Xiuqin Zhang et al., 2006). Fgfr4 is

not known to be alternatively spliced (Fon Tacer et al., 2010; Ornitz & Itoh, 2015). The specificity

of FGF-FGFR interaction is such that, for example, FGFs expressed in epithelial tissues bind

specifically to the FGFR2b and do not bind to FGFR2c (Eswarakumar et al., 2005) (Table 2).

Several gain-of-function (GOF) and loss-of-function (LOF) mutations in FGFR 1, 2 and 3 have been linked to human diseases of skeletal dysplasia, dwarfism and cancer (Du, Xie, Xian, &

Chen, 2012; Eswarakumar et al., 2005). FGFR2 GOF mutations have been identified in some cases of craniosyntosis (premature fusion of cranial sutures) and FGFR1 GOFs are present in glioblastoma brain tumors and 8P11myeloproliferative syndrome (Beenken & Mohammadi,

2009). Transmembrane mutations in FGFR3 are highly prevalent in all cases of achondroplasia

(dwarfism) (Eswarakumar et al., 2005). The pathophysiological functions of the individual Fgfr genes are beginning to be elucidated using knockout mouse models. The viability and phenotypes of these mouse models are listed in Table 3.

20

FGFR isoform Ligand specificity

FGFR1 IIIb FGF1, -2, -3, -7, -10, -22 FGFR1 IIIc FGF1, -2, -4, -5, -6, -19, -21, and -23

FGFR2 IIIb FGF1, -3, -7, -10, -22

FGFR2 IIIc FGF1, -2, -4, -5, -6, -9, -19, -21, and -23

FGFR3 IIIb FGF1, -8, -9, -16, -17, -18, -20

FGFR3 IIIc FGF1, -2, -4, -5, -6, -8, -9, -16, -17, -18, -19, -20, -21, and -23

FGFR4 FGF1, -2, -4, -5, -6, -19, -21, and -23

Table 2: Binding specificities of FGF ligands (except for intracrine FGFs) for the FGF receptors. Adapted from (Eswarakumar et al., 2005; Xiuqin Zhang et al., 2006).

21

FGF Knockout mouse phenotype Biological role receptor

Mesoderm/endoderm differentiation

Fgfr1 Embryonic lethal at E9.5-12.5 (Deng et al., 1994; Yamaguchi, Harpal, Henkemeyer, & Rossant, 1994)

Cell migration through primitive streak (Partanen, Fgfr1b Viable Schwartz, & Rossant, 1998)

Cell migration through primitive streak (Partanen et al., Fgfr1c Embryonic lethal at E9.5 1998)

Placenta and limb bud formation (Laestander & Fgfr2 Embryonic lethal at E4-5 Engström, 2014)

Development of the lung, limbs and teeth (Itoh & Fgfr2b Embryonic lethal at E0 Ornitz, 2011; Ornitz & Itoh, 2015)

Bone and skull development (Beenken & Mohammadi, Fgfr2c Viable 2009; Ornitz & Itoh, 2015)

Long bone overgrowth, Fgfr3 Early and postnatal bone development (Du et al., 2012) osteopenia

Cholesterol metabolism and bile acid synthesis Not Fgfr4 Viable clear (Eswarakumar et al., 2005; Ornitz & Itoh, 2015)

Table 3: Biological roles of the FGF receptors based on knockout mice phenotypes. Adapted

from (Beenken & Mohammadi, 2009; Eswarakumar et al., 2005; Itoh & Ornitz, 2011). E:

embryonic day

iii. FGF/FGFR signaling It is widely believed that the formation of a stable FGF-FGFR complex for intracellular

signaling requires the additional interaction with heparin or HSPGs (Belov & Mohammadi, 2013;

Schlessinger et al., 2000). There is growing evidence, however, to suggest that in the absence of heparin, functional FGF-FGFR complexes are still formed. Recently, Zakrzewska and colleagues

22 generated FGF1 mutants that had reduced affinity for heparin but retained their mitogenic activity due to the presence of stabilizing mutations (Zakrzewska et al., 2009; Zakrzewska, Marcinkowska,

& Wiedłocha, 2008). Once bound to a FGF ligand, the cell surface HSPG:FGF:FGFR and/or

FGF:FGFR units dimerize which facilitates transphosphorylation of the kinase domains on the intracellular side of the cell and activation of the receptor (Eswarakumar et al., 2005; Murakami et al., 2008; Schlessinger, 2000; Xiuqin Zhang et al., 2006). On FGFR1, Tyr653 and Tyr654, which reside in the activation loop, are the first residues to be phosphorylated and initiate an autophosphorylation cascade of other tyrosine residues (M Mohammadi et al., 1996).

Phosphorylated Tyr766 is a binding site for the SH2 domain of phospholipase Cγ (PLCγ) and the adaptor protein, Shb (Murakami et al., 2008). Tyr463 serves as a binding site for phosphotyrosine binding proteins, FGFR substrate 2α, β (FRS2α and FRS2β). FRS2α and β are docking proteins, that are also phosphorylated during receptor activation, and they assemble other downstream signaling molecules such as Src homolog (SH2) domain-containing adaptors Grb2 or Shp2. Grb2 recruits the docking protein Gab1 (Grb2-associated binder 1) and the resulting complex activates

Ras and the MAP kinase (ERK1/2, p38, JNK) pathways (Eswarakumar et al., 2005; Presta et al.,

2005; Turner & Grose, 2010). ERK1/2 activation is involved in endothelial cell proliferation, migration and differentiation while the p38 kinase is linked to FGF-mediated endothelial cell differentiation (Matsumoto, Turesson, Book, Gerwins, & Claesson-Welsh, 2002; Pintucci et al.,

2002; Presta et al., 2007). Gab also recruits and activates the cell survival phosphoinositide 3- kinase (PI3K)/Akt pathway via SH2 domains (Eswarakumar et al., 2005).

23

B. Angiogenic Fibroblast Growth Factors

Of the 22 proteins that make up the FGF family, FGF1, FGF2, FGF4 and FGF5 have been

shown to play a role in vascular growth and formation: (Presta et al., 2005). FGF1 and FGF2 are

the most extensively studied and are considered the prototypes of the vascular FGFs. FGF2 will

be discussed in further detail in the sections below.

FGF1 was first isolated from cultured tumor-derived endothelial cells as a mitogen for 343 fibroblasts (Folkman, Merler, Abernathy, & Williams, 1971). It was purified from bovine brain and pituitary using heparin affinity chromatography and found to have an isoelectric point of 5.6 which helped to distinguish it from basic FGF (FGF2, isoelectric point 9.6) which also had a high affinity for heparin and was mitogenic (Böhlen, Esch, Baird, & Gospodarowicz, 1985; Lobb &

Fett, 1984; Thomas, Rios-Candelore, & Fitzpatrick, 1984). Sequencing of bovine brain FGF1 and bovine pituitary FGF2 helped to further characterize FGF1 and to differentiate it from FGF2 though the two growth factors were found to have significant structural homology (Böhlen et al.,

1985; Esch et al., 1985). FGF1 is the “universal” FGF due to its ability to bind and activate all subtypes of the FGFRs (Eswarakumar et al., 2005; Itoh & Ornitz, 2011; Xiuqin Zhang et al., 2006).

The 17.5 kDa FGF1 protein is widely distributed including in primary ECs and vSMCs (Antoine et al., 2005; Fon Tacer et al., 2010). The biological activity of Fgf1 was examined by the generation of mice lacking FGF1. These mice were found to be viable and possess normal neuronal physiology and wound healing capabilities (Miller, Ortega, Bashayan, Basch, & Basilico, 2000).

The mitogenic and angiogenic properties of exogenous FGF1 have been observed in cultured human umbilical ECs, the chick chorioallantoic membrane and subcutaneous matrigel plugs (Lobb, Alderman, & Fett, 1985; Lobb & Fett, 1984; Mühlhauser et al., 1995;

Thomas et al., 1985). Topical application of FGF1 was shown to increase the healing rate of

24

excisional dermal wounds in healing-impaired diabetic mice (Mellin et al., 1995). Administration

of FGF1 DNA plasmids or recombinant protein to experimental models of PAD produced positive

results and highlighted the potential for the use of FGFs in treatment of ischemic diseases. Rabbits

(Pu et al., 1993, 1994; Tabata, 1997; Witzenbichler et al., 2006), rats (Rosengart et al., 1997) and

hamsters (Caron et al., 2004) subjected to unilateral femoral artery ligation and treated with recombinant or plasmid FGF1 had increased numbers of angiographically visible collateral vessels and enhanced blood flow. The positive results from these animal studies and a phase I safety study of recombinant FGF1 in 20 patients with three-vessel coronary artery disease confirmed the

clinical utility of growth factors in arterial stenoses (Schumacher, Pecher, von Specht, &

Stegmann, 1998). An open-label phase I safety study of NV1FGF (nonviral recombinant human

FGF1 DNA plasmid) added to the growing list of proof-of-concept (safety) studies and further encouraged the development of larger double-blind placebo controlled II/III trials (Comerota et al., 2002). The Therapeutic Angiogenesis Leg Ischemia Study for the Management of Arteriopathy

and Non-healing ulcers (TALISMAN), phase II trial of 125 CLI patients was the largest of the

phase II studies. NV1FGF intramuscular treatment did not produce differences from the primary

endpoint of ulcer healing. However, secondary endpoints (occurrence of amputation or death) were significantly reduction (Nikol et al., 2008). This study was followed by the TAMARIS trial, a phase III double-blind placebo-controlled multinational clinical trial (Belch et al., 2011). The trial enrolled 525 CLI patients with ischemic lesions and is the largest gene therapy that has been carried out for PAD. The NV1FGF treated-group did not differ from the placebo patients in the primary endpoint of time to major amputation or death. Despite the disappointing results of NVIFGF treatment, subsequent follow-up studies have revealed no adverse ischemic events, retinopathies or malignancies (Belch et al., 2011; Silvestre et al., 2013).

25

The Fgf4 and Fgf5 genes were both discovered during screenings of human tumor for genes with a transforming effect on 3T3 fibroblasts (Bovi et al., 1987; Yoshida et al., 1987; Zhan,

Bates, Hu, & Goldfarb, 1988; Zhan, Culpepper, Reddy, Loveless, & Goldfarb, 1987). They were classified as oncogenic proteins and fibroblast growth factors based on 43% (FGF4) and 50%

(FGF5) amino acid sequence homology to FGF2 (Yoshida et al., 1987; Zhan et al., 1988). The functions of the Fgf4 and Fgf5 were investigated with the development of mice containing targeted deletion of the genes. Fgf4 was found to be necessary for development as embryos died after uterine implantation at embryonic day 4-5 (Feldman et al., 1995). Fgf5-/- mice were viable but had abnormal long hair and suggests an inhibitory role for FGF5 in regulation of the hair growth cycle (Hébert et al., 1994). Fgf4 and Fgf5 are both widely expressed during development while adult expression of Fgf4 is low and limited to the intestines and testes (Feldman et al., 1995; Fon

Tacer et al., 2010; Haub & Goldfarb, 1991). High Fgf5 expression is observed in the adult central nervous system with moderate expression in muscle and bone (Fon Tacer et al., 2010; Haub,

Drucker, & Goldfarb, 1990).

The mitogenic activity of recombinant FGF4 protein has been established in cultured human umbilical vascular endothelial cells (Miyagawa et al., 1988), rat avascular cornea assay and in the chick chorioallantoic membrane (Dell’Era et al., 2001; Yoshida et al., 1994). In cDNA- transfected murine aortic endothelial cells, endogenous Fgf4 increased proliferation but had no effect on morphogenesis in anchorage-dependent models such like 3D fibrin and EC matrix gels.

The proliferative capacity of FGF4 and affinity for FGFRs was also reduced compared to FGF2

(Dell’Era et al., 2001; Rissanen et al., 2003). The therapeutic angiogenic potential of FGF4 for treatment of ischemic disease was also investigated similarly to FGF1 in animal models of hindlimb and myocardial ischemia. Replication-incompetent serotype 5 adenoviruses (Ad5) were

26

implemented to deliver human Fgf4 (Ad5FGF4) to a rabbit lower limb ischemia model. The

Ad5FGF4 therapy promoted angiogenesis and arteriogenesis and increased popliteal blood flow

and muscle perfusion (Rissanen et al., 2003). The vascular effects of Ad5FGF4 were reported to

be direct via FGF4-induced EC proliferation and indirect via increased expression of endogenous

VEGF (Deroanne et al., 1997). Similar improvements in regional tissue perfusion and function

were observed in stress-induced porcine myocardial ischemia models with intracoronary delivery

of Ad5FGF4 (Gao et al., 2004; Roth et al., 2006).

The preclinical results served as the basis for the Angiogenic Gene Therapy (AGENT) trial,

a first-in-man study of intracoronary administration of Ad5FGF4. The randomized placebo- controlled study enrolled only 79 patients with chronic stable angina and after 12 weeks,

improvements from baseline in the primary endpoint (exercise tolerance time, ETT) were observed

(Grines et al., 2002). Another small study, the Phase II AGENT 2 trial, was performed to determine if Ad5FGF4 could improve myocardial perfusion in 51 patients with stable angina and reversible ischemia. The study proved to be too small and underpowered for conclusive results; however, additional safety and feasible data were obtained (Grines et al., 2003). The safety profile of

Ad5FGF4 and the results of the small AGENT/AGENT2 trials propelled the initiation of two nearly-identical double-blind place-controlled phase III trials, AGENT3 and AGENT4. The former was carried out exclusively in the United States while the later was conducted in Europe,

Latin America and Canada (Henry et al., 2007). Together the AGENT3 and AGENT4 trials enrolled 532 recurrent stable angina patients who received intracoronary infusions of placebo or

Ad5FGF4. Single pooled analysis of results from both trials revealed no significant improvements in the primary endpoint of change from baseline of total ETT at 4 weeks, 12 weeks or 6 months.

27

5. Fibroblast Growth Factor 2 (FGF2)

FGF2 was originally isolated from bovine pituitary in the mid-1970s as a 146 amino acid protein that was mitogenic for 3T3 fibroblasts and endothelial cells (Y. Cao et al., 2011; Ribatti,

2014). FGF2 was also referred to as basic FGF (bFGF) due to its isoelectric point of 9.6 and the loss of biological activity at acidic pH values (Gospodarowicz, Bialecki, & Greenburg, 1978;

Gospodarowicz, 1975). In 1984, a growth factor derived from a rat chondrosarcoma was discovered and purified by Shing and co-workers based on the high binding affinity for heparin

(Shing et al., 1984). This factor was described as a cationic polypeptide with a molecular weight

of about 18 kilodaltons (kDa) with the ability to stimulate proliferation of BALB/C 3T3 fibroblasts

and capillary endothelial cells in vitro as well in vivo angiogenesis in the chorioallantoic membrane

(CAM) and yolk sac membrane of the developing chick (Shing et al., 1985). A heparin-binding

angiogenic factor was also purified from human placenta and a human hepatoma cell line and

shown to stimulate protease production, DNA synthesis, and migration capillary endothelial cells

(Moscatelli, Presta, & Rifkin, 1986; Presta, Moscatelli, Joseph-Silverstein, & Rifkin, 1986). The

factor was then sequenced and identified as fibroblast growth factor 2 (FGF2) (Esch et al., 1985).

Soon after, it was discovered that additional FGF2 polypeptides were translated from the human

Fgf2 gene and these proteins were described as high molecular weight forms of FGF2 (Robert Z

Florkiewicz & Sommer, 1989; Prats et al., 1989). FGF2 has since been studied extensively and

shown to have pleiotropic and mitogenic effects on myriad cell types from endoderm, mesoderm,

and neuroectoderm lineages (Okada-Ban, Thiery, & Jouanneau, 2000). The functions of the low

(18kDa) and high molecular weight isoforms (21, 22, 24 and 34kDa) of FGF2 will be discussed in

further detail below.

28

A. Fgf2 gene and its tissue expression

The single copy human Fgf2 gene is mapped to q26 on 4, and is composed of three separated by two 16 kilobase and extends to 40 kilobases (Okada-Ban et al., 2000; Ornitz & Itoh, 2001). The gene is highly conserved across many species including chimpanzee, Rhesus monkey, dog, cow, mouse, rat, chicken, zebrafish, and frog (Okada-Ban et al., 2000).

FGF2 is widely distributed in cells of mesenchymal, mesodermal and neuroectodermal lineages and is expressed during development and in the adult. In situ hybridization and RT-PCR analysis detected expression of Fgf2 in all stages of Xenopus and rodent embryonic development

(Gonzalez, 1990; Lea, Papalopulu, Amaya, & Dorey, 2009; Yaylaoglu et al., 2005). In adult murine tissues, Fgf2 mRNA is expressed in the brain, heart, liver, aorta, gut, skin, bone, skeletal muscle, reproductive organs and blood vessels (Fon Tacer et al., 2010). A similar expression pattern was observed in adult mouse tissue sections that were analyzed by immunostaining. In the vasculature, FGF2 was present in the basement membrane and smooth muscle of the of mid-size vessels while capillary vessels showed expression in the basal lamina and endothelium

(Cordon-Cardo, Vlodavsky, Haimovitz-Friedman, Hicklin, & Fuks, 1990). Primary co-cultures of human umbilical vascular endothelial and arterial smooth muscle cells also had strong expression of Fgf2 (Antoine et al., 2005).

B. Phenotypes of Fgf2-/- mice (Role in development, homeostasis and injury)

Several, independently developed lines of mice deficient in the Fgf2 gene expression have been produced to describe the roles of FGF2 in development and physiology. House and colleagues have shown that normal growth of capillary and smooth muscle-containing vessels occurs in FGF2

29

knockout of mice bred on a 50:50 129/Sv and Black Swiss background (House et al., 2003).

A similar result was observed by the Murray group in Fgf2-/- mice of a different strain, C57BL/6J

(Virag et al., 2007). Mice bred on a C57BL/6J x 129/Sv mixed background, however, had impaired myocardial capillary density when Fgf2 gene expression was partially (+/-) or completely (-/-)

disrupted (Amann et al., 2006). Fgf2-deficient mice have normal retinal vascularization and

neovascularization responses to retinal or choroid layer injury (Ozaki et al., 1998; Tobe et al.,

1998). Ablation of Fgf2 delayed healing of dorsal skin excisional wounds (Ortega, Ittmann, Tsang,

Ehrlich, & Basilico, 1998). Sullivan and investigators showed that in Fgf2-/- hindlimbs, vascular

growth occurs to a similar degree as wildtypes in response to chronic ischemia (Sullivan,

Doetschman, & Hoying, 2002). FGF2 deficiency decreased the effect of the sex hormone 17β- estradiol (E2) on the proliferation and migration of cultured endothelial cells (Garmy-Susini et al.,

2004). After carotid arterial vessel injury, E2-mediated acceleration of re-endothelialization and

endothelial progenitor cell (EPC) mobilization was attenuated in the absence of FGF2 (Fontaine

et al., 2006), while flow-dependent remodeling was not changed (Sullivan & Hoying, 2002) and

normal neointimal hyperplasia was present (Zhou et al., 1998).

Other cardiac phenotypes of the Fgf2-/- mice include resting hypotension, decreased portal

vein vascular smooth muscle contractility, normal basal cardiac function and impaired

baroreceptor reflex (Dono, Texido, Dussel, Ehmke, & Zeller, 1998; Ortega et al., 1998; Pellieux et al., 2001; Schultz et al., 1999; Zhou et al., 1998). Normotensive Fgf2-/- mice exhibited baseline dilated cardiomyopathy with increased left ventricular chamber diameter at systole or diastole

(Pellieux et al., 2001). When subjected to different experimental hypertension models, including -induced hypertension (Pellieux et al., 2001), pressure-overload (Schultz et al., 1999), or chronic β-adrenergic stimulation (House et al., 2010); Fgf2-/- hearts were protected from

30 compensatory cardiomyocyte hypertrophy. This blunted hypertrophic response was also present in Fgf2-ablated hearts after permanent occlusion of the left-descending coronary artery, a model of chronic ischemia (Virag et al., 2007) and in response to a closed-chest model of cardiac ischemia-reperfusion injury (House et al., 2015). In both models, cardiac cell death (measured as infarct size) was significantly greater relative to wildtype hearts. The later study also reported reduced cardiac vessel density after 7 days of reperfusion (House et al., 2015). Infarct size was also increased in response to low-flow I/R injury and this resulted in reduced recovery of post- ischemic cardiac function in Fgf2-/- hearts (House et al., 2003).

Non-cardiovascular phenotypes in the FGF2-deficient mice were observed, suggesting a role for FGF2 in bone and nerve patho(physiology). For example, bone mass and formation are decreased in Fgf2-/- femora and tibiae and this osteopenia progressed with age (Montero et al.,

2000). Reduced neuron density and disorganization has been shown in the cerebral cortices of

FGF2-deficient mouse brains (Dono et al., 1998; Korada, Zheng, Basilico, Schwartz, & Vaccarino,

2002; Ortega et al., 1998; Raballo et al., 2000). The neuronal deficit in Fgf2-/- mice was not a factor of defective neural stem cell (NSC) proliferation during embryonic or adult neurogenesis, but rather an impairment in the differentiation of the NSCs into mature neurons (Werner, Unsicker,

& von Bohlen und Halbach, 2011). In response to peripheral nerve crush injury, increased macrophage invasion, regeneration of myelinated axons and greater axon and myelin diameters resulted from an absence of Fgf2 (Jungnickel, Claus, Gransalke, Timmer, & Grothe, 2004).

Additionally, sciatic nerve-injured knockout mice exhibited faster mechanosensory recovery

(Jungnickel et al., 2010).

C. FGF2 protein isoforms

31

The Fgf2 gene produces several protein isoforms from a single mRNA. The protein

isoforms are grouped into two classes: low molecular weight (LMW) and high molecular weight

(HMW). A single LMW FGF2 (18kDa) is translated from a classical AUG codon while the HMW

isoforms are translated from in-frame alternative translational sites at CUG codons that are located

upstream of the LMW start site (Touriol et al., 2003). The 18kDa LMW FGF2 isoform is present

in mice and humans. Humans express four HMW isoforms (22, 22.5, 24, and 34kDa) while mice

express two HMW isoforms (21.5 and 22kDa) (Okada-Ban et al., 2000) (C. H. Chen, Poucher, Lu,

& Henry, 2004; Liao et al., 2009; Okada-Ban et al., 2000). Structurally, the HMW isoforms are co-linear extensions of the LMW isoform with amino terminal nuclear localization signals that localize them to the cell nucleus (Chlebova et al., 2009; Yu, Ferrari, Galloway, Mignatti, &

Pintucci, 2007) (Figure 2). In neuronal cells, HMW isoforms appear in a punctuate pattern within the nucleoplasm, around nucleoli and associated with chromatin (Claus et al., 2003; Förthmann,

Grothe, & Claus, 2015). LMW FGF2 is mostly a cytosolic protein but has been found to localize to nucleoli in the nucleus and is stored in the extracellular matrix (Claus et al., 2003; Förthmann et al., 2013; Liao et al., 2010; Quarto, Fong, & Longaker, 2005).

The precise manner of secretion of LMW FGF2 is still unknown because like FGF1, it lacks a conventional export signal sequence for transport via the canonical endoplasmic reticulum

(ER)/Golgi secretory pathway (Mignatti, Morimoto, & Rifkin, 1992). Proposed mechanisms for release of FGF1 and FGF2 into the extracellular space include an ATP-dependent manner that implicates Na+/K+-ATPase, vesicular shedding from plasma membranes, or from non-lethal

damage or disruption of membranes (Florkiewicz, Anchin, & Baird, 1998; Sørensen, Nilsen, &

Wiedłocha, 2006; Taverna et al., 2003). Specific HMW or LMW FGF2 isoform detection has been

performed using Western immunoblotting. Several groups have reported expression of the LMW

32

(18kDa) and HMW (21, 22kDa) FGF2 isoforms in mouse brain, liver, heart, lung and spleen

(Azhar et al., 2009; Claus, Werner, Timmer, & Grothe, 2004; Coffin et al., 1995; Garmy-Susini et

al., 2004; House et al., 2003; Miller et al., 2000). Selective expression of LMW FGF2 was

observed in the kidney and skeletal muscle (Coffin et al., 1995). In rat embryonic whole brain,

only 18kDa and 21kDa FGF2 were expressed until birth when the 22kDa could then be detected.

A similar pattern was observed in embryonic spinal cord and cortex extracts while adult extracts

expressed higher levels of 21 and 22kDa FGF2 (S. Giordano, Sherman, Lyman, & Morrison,

1992). In the adult cerebellum, however, 18kDa FGF2 was the predominant form and the

expression of the 21kDa protein disappears at birth (S. Giordano et al., 1992). Adult rat heart

ventricles had selectively higher expression of the 18kDa protein in contrast with neonatal

ventricles where the 22kDa isoform was preferentially expressed (L. Liu, Doble, & Kardami,

1993).

33

Figure 2: Canonical (AUG, LMW) and alternative (CUGs, HMW) FGF2 protein isoform translation. Adapted from (Liao et al., 2009).

D. Biological functions of the FGF2 isoforms

FGF2 has been shown to play a pathophysiological role in diverse biological processes, including development, tumorigenesis, wound healing, hematopoiesis, neuronal development, limb formation, cardiac hypertrophy, and vascular growth (C. H. Chen et al., 2004; Fei,

Gronowicz, & Hurley, 2013; Kardami et al., 2007; Liao et al., 2009; Ortega et al., 1998; Presta et al., 2005; Schultz et al., 1999; Shing et al., 1985; Zhou et al., 1998). The pleiotropic effects of

FGF2 in a multitude of biological systems can be linked to the alternative translation of its protein isoforms and their different subcellular localization.

i. Vascular Biology

The initial discovery and characterization of FGF2 was based on its isolation from and mitogenic effects on tumor-derived endothelial cells (ECs) and vascular smooth muscle cells

34

(vSMCs) (Gospodarowicz et al., 1978; Shing et al., 1985). The role of FGF2 has since expanded to include complex cellular processes that contribute to physiological and/or tumor-associated angiogenesis (Auguste, Javerzat, & Bikfalvi, 2003; Javerzat, Auguste, & Bikfalvi, 2002; Presta et al., 2007). These include EC and vSMC proliferation, chemotactic migration, protease production, extracellular matrix (ECM) degradation, morphogenesis and assembly of nascent or remodeled vascular structures (Carmeliet & Jain, 2011; Fischer et al., 2006; Presta et al., 2005). In blood vessels, FGF2 is produced and secreted by vascular ECs and its mechanism of action can be intracrine, paracrine and/or autocrine. Expression of similar levels of Fgfr1IIIc and 3IIIc mRNA

have been shown in ECs and vSMCs, while lower amounts of Fgfr3IIIc and no Fgfr4 were detected

(Antoine et al., 2005). In MCF-7, aortic endothelial and 3T3 cells, only FGFR1 protein expression was detected (Piotrowicz, Ding, Maher, & Levin, 2001). FGF2 is also found in the ECM where it is bound to matrix components including fibronectin, αvβ3 integrins, vimentin, collagen type I,

thrombospondin-1 and HSPGs (Presta et al., 2005, 2007). The ECM-immobilized FGF2 is a

stimulus for angiogenesis when it is freed by proteases and glycosidases and it promotes cell

adhesion, activation and motility via recruitment of αvβ3 receptors and focal adhesion kinases

(Folkman et al., 1988; Presta, Maier, Rusnati, & Ragnotti, 1989; Tanghetti et al., 2002).

To date, only the LMW FGF2 (18kDa) isoform is commercially available in recombinant

form and this protein has been applied exogenously to determine the functions of FGF2. Due to

the distinct subcellular locations and properties of the FGF2 isoforms, it has become critical to

employ genetic models where only one class of the FGF2 isoforms are expressed to examine the

specific functions of the endogenous FGF2 proteins. These mouse models will be described in the

following sections “Phenotypes of FGF2 isoform knockout mice”.

35

ii. Non-vascular effects of exogenous FGF2 isoforms

Studies on the mitogenic functions of exogenous FGF2 isoforms extend beyond their effects on vascular-derived cells to include cells/tissues of mesodermal origins including cardiomyocytes and myofibroblasts in the heart (Fujiu & Nagai, 2014; Kardami et al., 2007;

Santiago et al., 2011). Exogenous administration of recombinant LMW FGF2 to intact rodent hearts undergoing surgical corornary artery occlusion or ex vivo hearts being subjected to acute ischemia significanlty reduced the dysfunction associated with these models of myocardial infarction (MI) (Jiang et al., 2002; Jiang, Srisakuldee, Soulet, Bouche, & Kardami, 2004; Sheikh,

Sontag, Fandrich, Kardami, & Cattini, 2001). Futhermore, myocyte injury (infarct size) was decreased in the FGF2-treated hearts. The protective effect of the LMW isoform was maintained for up to 6 weeks (in vivo studies) and was not dependent of the timing of the ischemic injury; i.e., exogenous LMW FGF2 was beneficial if given to the hearts before, during or after the development of ischemia (during reperfusion) (Jiang et al., 2002, 2004; Kardami et al., 2007). A mutant form of the LMW isoform (S117A) lacking mitogenic or angiogenic ability still retained the potential for cardioprtection (Jiang et al., 2004).

When injected into infarcted rat hearts, recombinant HMW FGF2 (23kDa) alone was also effective in preventing myocyte death and improving functional recovery (Jiang et al., 2007). This level of protection observed was similar to hearts that had been treated with the LMW isoform.

The LMW FGF2 group had sustained cytoprotective effects and protection from contractile dysfunction for up to 6 weeks after the initial treament (Jiang et al., 2007). Increased small vessel

(arteriole) density was also detected after 6 weeks of MI in the LMW FGF2-infarcted hearts. HMW

FGF2 had no effect on vessel growth but increased heart dimensions (LV mass) and cardiomyocyte size were observed after 8 week of MI. This hypertrophic response to HMW FGF2 was confirmed

36 in cultured neonatal cardiomyocytes and non-cardiac myocytes treated with the exogenous isoform (Jiang et al., 2007; Kardami et al., 2004; Santiago et al., 2011). The acute protection by exogenous LMW or HMW FGF2 was found to be dependent on activation of FGFR1 and related downstream signaling proteins including PKCα, ε, δ, and ζ (Jiang et al., 2002, 2004). Several studies using Fgf2-/- mice (described in “Phenotypes of Fgf2-/- mice”) have documented a role for the Fgf2 gene in chronic adrenergic stimulation or pressure overload-induced cardiac hypertrophy. The actions of the recombinant HMW isoform would imply that nuclear signaling by

HMW FGF2 and not LMW FGF2 is responsible for the role of Fgf2 in development of cardiac hypertrophy.

The in vitro and in vivo studies described here to characterize the properties of FGF2 have utilized exogenous (recombinant) forms of the proteins. While LMW FGF2 is exported from the cell and can signal via the cell surface FGFRs, HMW FGF2 isoforms are nuclear proteins which share considerable sequence homology with LMW FGF2 including receptor binding sites. This would suggest that activities of exogenous HMW FGF2 may mimic that of LMW FGF2 signaling, however, physiological functions of HMW FGF2 would have to be confirmed in studies with endogenous HMW FGF2 (Chlebova et al., 2009; Gualandris, Urbinati, Rusnati, Ziche, & Presta,

1994).

iii. Effect of overexpression of FGF2 isoforms

The functional effects of the FGF2 isoforms have also been studied in models of Fgf2 gene overexpression in vitro and in vivo. Mouse aortic endothelial (MAE) cells transfected with human

Fgf2 cDNA displayed an angiogenic phenotype when plated on ECM gels and in the avascular rabbit cornea. This activity was inhibited by a neutralizing FGF2 antibody and suggests an autocrine mechanism for this phenotype (Gualandris et al., 1996). Aortic vascular smooth muscle

37

cells from transgenic mice expressing all the human FGF2 isoforms, all the HMW FGF2 isoforms

or only the 18kDA isoform had enhanced DNA synthesis with increased proliferation observed

when only the nuclear HMW isoforms were overexpressed (Davis et al., 1997). Preservation of

the increased cell growth in presence of antibody neutralization, established an intracrine role for

FGF2 in vascular cell growth (Davis et al., 1997). When 3T3 fibroblasts overexpress only LMW

or only HMW FGF2, different cellular phenotypes emerge. The HMW-expressing cells had

impaired growth relative to wildtype cells, while the LMW FGF2 cells resembled transformed

cells with enhanced saturation density and growth in low serum (Quarto, Talarico, Florkiewicz, &

Rifkin, 1991; Wieder et al., 1997).

In response to transient overexpression of LMW FGF2 or HMW FGF2 cDNA in rat

neonatal ventricular myocytes, similar increases in myocyte DNA synthesis and proliferation were

observed (Pasumarthi, Kardami, & Cattini, 1996). When HMW FGF2 (22/21.5kDa)

overexpression and not LMW FGF2 was present, there was increased binucleation of the myocytes

that was not inhibited by FGF2 neutralizing antibodies. Bi- or multi-nuclear phenotype has also

been reported in HMW FGF2-overexpressing neuronal cells (Nindl et al., 2004). These studies

provide further evidence for an intracrine (nuclear) pathway of HMW FGF2 signaling. Mice with

cardiac-specific overexpression of LMW FGF2 had unchanged heart weight-to-body weight ratios and arteriolar vessel density as well as normal expression of cardiac differentiation markers

(Sheikh et al., 2001). Capillary density and expression of MAP kinases (JNK, p38) and PKC isoforms (α, ε) were increased in unstressed LMW FGF2 transgenic hearts. After ex vivo global ischemia-reperfusion (I/R) injury, overexpression of LMW FGF2 was shown to elevate intra- and extracellular FGF2 expression in addition to preventing myocyte death (Sheikh et al., 2001). These results are partially consistent with a study in which cardiac specific overexpression of all human

38

FGF2 (18, 22-34 kDa) isoforms likewise produced no basal changes in vessel densities in the heart

but resulted in protection from I/R injury-induced myocyte loss (House et al., 2003). Conversely,

House and co-workers reported improvements in recovery of myocardial function with normal

expression of MAP kinases and PKC isoforms with the exception for PKCε whose expression was

reduced (House et al., 2003; House, Branch, Newman, Doetschman, & Schultz, 2005; House,

Melhorn, Newman, Doetschman, & Schultz, 2007). Alterations in the activation of p38 and JNK

kinases only occurred in response to I/R injury as overexpression of all human FGF2 isoforms

already had modified p38 and JNK activation independent of ischemia (House et al., 2005). The

human FGF2 transgenic hearts exhibited an exaggerated hypertrophic response to isoproterenol

stimulation in the form of decreased fractional shortening, increased heart weight-to-body weight

ratios and myocyte cross-sectional area.

E. Phenotypes of FGF2 isoform knockout mice (role of endogenous FGF2 isoforms)

FGF2 is a widely expressed growth factor and promotes mitogenic activity in a variety of

cell types including vascular cells (Davis et al., 1997), fibroblasts (Gospodarowicz, 1975),

cardiomyocytes (Pasumarthi et al., 1996), skeletal myocytes, neurons (Grothe, Haastert, &

Jungnickel, 2006), and osteoblasts (Xiao et al., 2003). LMW FGF2-deficient (FGF2 LMWKO) or

HMW deficient (HMWKO) mouse models are viable and fertile. Using microvascular endothelial

cells isolated from ovariectomized LMWKO mice, Garmy-Susini and co-workers observed that

HMW FGF2 was necessary for 17 β-estradiol (E2)-mediated migration and angiogenesis (Garmy-

Susini et al., 2004). This effect of HMW FGF2 was FGF receptor-independent but involved the

anti-apoptotic protein FGF2-interacting Factor (FIF), a putative intracellular signaling partner of

HMW FGF2 (Chlebova et al., 2009; Krejci et al., 2007). To date, this in vitro work by the Arnal

39

group represents the only study undertaken to examine the in vivo effects of the endogenous HMW

FGF2 isoforms in vascular biology. FGF2 isoform knockouts have provided an opportunity to further examine the unique functions of the endogenous FGF2 isoforms in the vascular pathophysiology. Three different groups have performed experiments using these unique mouse models to distinguish the biological roles of the FGF2 isoforms in the heart, bone and brain.

Findings from these studies are summarized below. Taken together, they confirm that when expressed endogenously, the HMW and LMW isoforms of FGF2 play different roles in tissue homeostasis and pathophysiology.

Heart

In parallel with the characteristics of Fgf2-/- mice (described in “Phenotypes of Fgf2-/-

mice”), FGF2 LMWKO and FGF2 HMWKO mice exhibit normal vascular growth and altered

FGF2 expression. Capillary and arteriole density is not significantly different between FGF2

LMWKO (HMW expressing) or FGF2 HMWKO (LMW expressing) and WT hearts. While the

expression of the 21 and 22kDa FGF2 (HMW) isoforms are unchanged in the LMWKO hearts

(Liao et al., 2007), 18kDa (LMW) FGF2 is increased by 1.5 fold in HMWKO hearts (Liao et al.,

2010). The subcellular localization of the HMW isoforms was confirmed to be exclusively nuclear

and LMW FGF2 was expressed in the cytoplasm and nucleus (Azhar et al., 2009; Liao et al., 2009).

The increased level of LMW FGF2 in the absence of HMW FGF2 played a protective role in

ischemia–reperfusion injury

When subjected to an ex vivo work performing model, the knockout hearts displayed no

changes in baseline contractile and relaxation parameters (Liao et al., 2007, 2010). After induction

of low-flow ischemia and reperfusion (I/R), impaired recovery of post-ischemic contractile

40 function was observed in LMWKO hearts (Liao et al., 2007). Conversely, improved recovery of function occurred in HMWKO hearts (Liao et al., 2010). The cardioprotective effect of LMW

FGF2 was shown to be dependent on the activation of FGFR1 and inhibition of JNK signaling

(Liao et al., 2007). Creatine kinase release and myocardial infarct size, measures of cell injury were not different from WT indicating a requirement of LMW and HMW FGF2 isoforms in the protection from cell death (Liao et al., 2007, 2010).

Recently, the Doestchman group has performed echocardiographic measurements to characterize the interactions between sex and the FGF2 isoforms in regulating cardiac physiology and modulating remodeling in (Nusayr & Doetschman, 2013; Nusayr, Sadideen, & Doetschman,

2013). In unstressed hearts, male and female mice LMWKO (HMW expression only) hearts had impaired basal diastolic function (compliance and relaxation) while only female LMWKO hearts had abnormal cardiac development with smaller LV masses (Nusayr & Doetschman, 2013).

HMWKO (LMW expression only) males had significantly higher systolic functional parameters

(cardiac output, stroke volume and fractional shortening) than their female cohorts as well as in comparison to male WTs or LMWKOs. After chronic β-adrenergic stimulation, only female

LMWKO hearts had an elevated hypertrophic response while their male cohorts had increased and cardiac remodeling (Nusayr et al., 2013). Male HMWKOs had enhanced expression of the cardiac stress marker, atrial natriuretic factor (ANF) which protected them from the higher cardiac fibrosis that was present in WT and LMWKO male hearts (Nusayr et al., 2013). These observations suggest that in addition to the isoform-specific effects on injury response, there is a sex-specific regulation of cardiac patho(physiology) by FGF2.

Bone growth and remodeling

41

FGF ligands and FGF receptors play important roles in bone formation, homeostasis, repair

and remodeling. In humans, several gain-of-function and loss-of-function point mutations in

FGFRs (FGFR1, 2 and 3) result in deficits of digital, cranial, and skeletal development including craniosynostosis and skeletal dysplasias (Belov & Mohammadi, 2013; Eswarakumar et al., 2005;

Marie, Miraoui, & Sévère, 2012). Phenotypes of FGFR knockout mouse models (summarized in

Table 3) are consistent with the human disorders. Mutations in several FGFs including FGF3,

FGF8, FGF9, FGF10 and FGF23 have also been shown to cause human skeletal diseases (Du et al., 2012).

The role of the FGF system in bone regulation is also evident in mouse models with altered expression of FGF2. Abnormal bone development occurs when LMW and HMW FGF2 are overexpressed globally. These mice have moderate macrocephaly and dwarfism (shortening of the femora and humeri) (Coffin et al., 1995). Additionally, bone mineral density (BMD) and trabecular bone volume were reduced in the femora of the transgenic mice (Sobue et al., 2005). Targeted overexpression of LMW FGF2 in cells of osteoblastic lineages established a role for the LMW isoform as a positive regulator of bone mass. BMD, tibial and femoral bone volume and trabecular and cortical bone thickness were increased in the LMW FGF2 transgenic mice (Xiao et al., 2014).

Conversely, global LMW FGF2 knockout mice have decreased BMD and bone mineral content

(Xiao et al., 2009). When the HMW isoforms are selectively overexpressed in mouse osteoblasts, dwarfism, decreased BMD, osteomalacia and hypophosphatemia were observed (Xiao et al., 2010;

Xiao, Esliger, & Hurley, 2013). This hypophosphatemia phenotype is also present in the ostoeblast-specific overexpression of another FGF, FGF23 (Xiao et al., 2010). FGF23 in the serum and bone mRNA and protein were increased in the HMW transgenic mice. HMW FGF2 regulation of bone phosphate metabolism and inhibition of bone formation was shown to be dependent on

42

FGF23 activation of FGFRs and the MAPK pathway (Xiao et al., 2013). In cultured calvarial

osteoblasts (COBs) isolated from LMW FGF2 or HMW FGF2 newborn transgenic mice,

differential regulation of bone morphogenic protein-2 (BMP-2) was found to mediate the disparate

effects of the FGF2 isoforms on bone growth (Sabbieti et al., 2013). BMP-2 is potent inducer of

osteoblast differentiation and bone formation. When added to COBs, BMP-2 promoted or inhibited

bone differentiation in response to LMW FGF2 or HMW FGF2 overexpression, respectively

(Naganawa et al., 2008; Sabbieti et al., 2013).

Altered expression of FGFs and FGFRs have been found in bone during fracture healing

and bone regeneration (Du et al., 2012; Fei et al., 2013). Bone mRNA expression of Fgf1, Fgf2,

Fgf5, Fgfrs and other markers of bone development are increased in rodent models of tibial and femoral fracture (Du et al., 2012; Schmid, Kobayashi, Sandell, & Ornitz, 2009). When subjected bilateral calvarial defects, male LMW transgenic mice had enhanced healing compared with non- transgenic mice (Xiao et al., 2009). Treatment with low doses of BMP-2 further improved the bone healing at later stage and led complete healing of the defect (Fei et al., 2013).

Neurogenesis and neurodegeneration

FGF2 has been shown to be important for the development of the central nervous system

(CNS) and in adult neurogenesis. FGF2 and FGFR expression is temporally and spatially regulated during development (Woodbury & Ikezu, 2013; Yu et al., 2007). Furthermore, Fgf2 mRNA and

protein expression is present in spinal ganglia and is up-regulated in response to peripheral nerve

injury (Grothe, Meisinger, Hertenstein, Kurz, & Wewetzer, 1996). The FGF2 isoforms were

differentially elevated in the nerve lesions with increased 18kDa isoform expression 5 hours after

injury, while the HMW FGF2 isoforms had a greater increase at 7 days (Grothe, Heese, et al.,

43

2000; Meisinger & Grothe, 1997a). Conversely, no changes in the expression of the FGF2 protein

isoforms or FGF receptor transcripts were observed in neurotoxin-induced lesions of dopaminergic

neurons (Claus, Werner, et al., 2004). Transfection of immortalized Schwann cells (SC) and cultured PC 12 cells with modified rat cDNAs expressing endogenous 18kDa or 21/23 kDa resulted in differential growth and morphological phenotypes (Grothe, Meisinger, Holzschuh,

Wewetzer, & Cattini, 1998). Proliferation of the 18kDa overexpressing Schwann and PC12 cells was decreased compared to the HMW FGF2 cells (Müller-Ostermeyer, Claus, & Grothe, 2001).

The HMW FGF2 isoforms emerged as neurotrophic factors when HMW FGF2 released from implanted SCs lead to improvements in the lengths and numbers of regenerating myelinated axons in a rat model of sciatic nerve axotomy (Timmer, Robben, Müller-Ostermeyer, Nikkhah, &

Grothe, 2003). In Parkinson’s-diseased rat brains, co-transplantation of dopaminergic grafts with

HMW FGF2 SCs also enhanced re-innervation, survival and functional recovery compared to the

18kDa SCs (Timmer et al., 2004). In vivo, grafted FGF2 isoform-modified SCs had differential effects on nerve regeneration across long gaps. The 18kDa grafts were inhibitory for myelination of regenerating axons while the 21/23kDa expressing SCs stimulated long-distance axon myelination and early recovery of sensory nerve functions (Haastert, Lipokatic, Fischer, Timmer,

& Grothe, 2006).

F. Exogenous LMW FGF2 in revascularization in vivo (Preclinical studies)

Models of angiogenesis

Exogenous FGF2 is a potent stimulator of angiogenesis in experimental angiogenesis assays (Presta et al., 2005, 2007; Shing et al., 1985). These include in vitro (EC proliferation,

Boyden chamber migration, EC tubule formation assays, 3-D collagen matrix gels); ex vivo organ

44

cultures (rodent aortic ring) or in vivo (chick embryo chorioallantoic membrane (CAM), subcutaneous ECM gel sponges, dorsal skin fold chamber or avascular rodent cornea) assays

(Baker et al., 2012; Staton, Reed, & Brown, 2009). In early studies characterizing the activities of tumor-derived factors, an 18kDa protein stimulated proliferation in capillary bovine capillary endothelial cells (BCECs) and angiogenesis in chorioallantoic membranes and rat cornea (Shing et al., 1984, 1985). This protein, which was soon named FGF2, also induced DNA synthesis, motility, protease production in BCECs (Moscatelli et al., 1986; Presta et al., 1986) and capillary- like tubule formation in BCECs cultured on 3D collagen matrix gel (Montesano, Vassalli, Baird,

Guillemin, & Orci, 1986). After expression and purification from transformed Escherichia coli cells, exogenous treatment with recombinant 24kDa FGF2 was shown to have a similar effect as

18kDa FGF2 in promoting proliferation, protease production and chemotaxis of cultured bovine aortic endothelial cells (BAECs). Induction of angiogenesis in the rabbit cornea and binding of

FGF receptors was also similar in response to 18kDa or 24kDa FGF2 (Gualandris et al., 1994).

LMW FGF2-stimulated angiogenesis in cultured BCECs and in CAMs requires activation of the phosphoinositide 3-kinase (PI3K) pathway and is negatively regulated by the p38 MAP kinase pathway (Matsumoto et al., 2002; Qi et al., 1999).

Levin and colleagues observed that human HMW FGF2 isoforms have an inhibitory effect on BAECs and mammary carcinoma cells (MCF-7) in a Boyden chamber type assay (Piotrowicz,

Maher, & Levin, 1999). When both cell types were cultured in a monolayer and subjected to a wounding assay, recombinant HMW FGF2 inhibited migration of only BAECs (Piotrowicz et al.,

1999). HMW FGF2 had a similar proliferation-promoting effect on cells as the 18kDa. The inhibitory function of HMW FGF2 on cell migration was shown to be dependent on receptor α and regulated by the amino-terimal extension (ATE) of HMW FGF2 (Ding et al., 2002;

45

Piotrowicz et al., 2001). The inhibitory role of the ATE of HMW FGF2 on cell migration was also observed in vivo and found to be sufficent to suppress tumor angiogenesis in murine dorsal skin- fold model (Levin, Sikora, Ding, Rao, & Sriramarao, 2004). In another study utilizing exogenous

FGF2 in vivo, rat hearts had increased arteriole density in the peri-infarct after permanent coronary occlusion when treated with LMW FGF2. Conversely, HMW FGF2 did not have any effect on microvesels numbers (Jiang et al., 2007).

These in vivo models of angiogenesis have also been used to compare the efficacy of FGF2 with other proangiogenic growth factors. In particular, the corneal pocket assay has been employed to examine the vascular response to exogenous FGF2 relative to vascular endothelial growth factor

(VEGF) and platelet-derived growth factors (PDGF). The cornea is an avascular transparent tissue in the body, so any vessels penetrating into the corneal stroma are newly formed and easily quantified. The results from a number of studies have been summarized in Table 4. These have been limited to studies where FGF2 was compared directly with other growth factors to eliminate the possibility that difference in efficacies are due to variations in experimental parameters.

Overall, FGF2 induced corneal angiogenesis to a similar degree as PDGF-BB. Both FGF2 and

PDGF-BB were more efficacious in stimulating capillary network growth relative to other PDGF isoforms, PDGF-AA and PDGF-AB.

46

Delivery Neovascularization Assay FGF2 PDGF-AA PDGF-AB PDGF-BB VEGF-A Reference (protein) endpoint Corneal Slow release Vascular area (H. Lu et +++ + ++ +++ NT micropocket polymer (mm2) al., 2007) Corneal Vascular area (R. Cao et Micropellets +++ + +++ +++ ++ micropocket (mm2) al., 2003) Corneal Slow release Vascular area (Nissen et +++ NT NT +++ NT micropocket polymer (mm2) al., 2007) Corneal Vascular area (J. Zhang et Micropellets +++ + ++ ++ NT micropocket (mm2) al., 2009)

Table 4: Angiogenic responses induced by recombinant LMW FGF2 protein relative to other proangiogenic growth factors in corneal micropocket assay. For the endpoint (vascular area), VEGF and PDGF are responses were estimated relative to the

FGF2 response .

47

The early in vitro, ex vivo, and in vivo assays highlighting the proliferative effects of FGF2

and other growth factors served as the rationale for animal studies to evaluate the potential

pharmacological application of these “angiogenic” factors to stimulate angiogenesis and/or

collateral vessel formation in models of myocardial or peripheral ischemia. Experimental

myocardial and peripheral ischemia models have been beneficial in testing the efficacy of

angiogenic gene or protein delivery. For myocardial ischemia, a majority of initial experiments

focused on the vascular responses to progressive occlusion of coronary artery or arterial stenosis

in canine and porcine models (Hearse, 2000; Rubanyi, 2013; Teunissen, Horrevoets, & van Royen,

2012). Since then, other species including mouse, rats, and rabbits have also been studied.

Animal models of myocardial ischemia

The extent of the pre-existing collateral circulation has proved to be an important determinant of the appropriate animal model for analyzing post-ischemic revascularization

(Schaper, 2000). An extensive study by Maxwell, Hearse and Yellon has shown that a wide range

exists between mammalian species in the extent of the pre-formed collateral network (Maxwell,

Hearse, & Yellon, 1987). This network provides collateral flow upon coronary artery occlusion

and is responsible for outward vascular remodeling. Pigs, baboons, and rabbits have little to no

collateral flow (0-6% of pre-ischemia) after ischemia; dogs and cats have intermediate baseline

collateral flow (12-17%) whereas guineas pigs have total compensation (100%) of flow after

arterial occlusion of a single coronary artery. The canine models were the most commonly utilized

models, and the work by the Yellon group and others revealed an extensive collateral network in

dogs that was capable of providing 40% of baseline perfusion after acute coronary artery occlusion

(G. C. Hughes, Post, Simons, & Annex, 2003; Maxwell et al., 1987; Teunissen et al., 2012). These

48

studies prompted a re-evaluation of the most appropriate experimental model for use in researching

arterial disease therapeutics for human patients. The minimal pre-existing collateral network present in pigs (1% of normal flow), which is similar to that found in humans, renders it a more suitable model of proof-of-concept studies (Hearse, 2000; Maxwell et al., 1987; Rubanyi, 2013).

Additionally, the heart weight-to-body weight ratio, functional and vascular response to myocardial ischemia of the pig is similar to that of humans (Elmadhun et al., 2013;

Rubanyi, 2013; Teunissen et al., 2012).

The ameroid model of myocardial ischemia developed in both dogs and pigs is widely used and is considered the gold standard for pre-clinical studies of pro-angiogenic therapeutics (G. C.

Hughes et al., 2003; Radke et al., 2006; Rubanyi, 2013). In this model, an ameroid constrictor is placed around the left circumflex coronary artery (LCx) which supplies 20% of the left ventricle in pigs and 25% in dogs (G. C. Hughes et al., 2003; Reimer, Ideker, & Jennings, 1981). The ameroid constrictor is a metal device with an inner ring of casein that is encased in a stainless steel sheath. Casein, a hygroscopic compound, gradually swells as it absorbs moisture in the body (G.

C. Hughes et al., 2003; Slatter, 2003). The steel ring is designed to force inward swelling of the casein and which gradually occludes the enclosed vessel over time (3-7 weeks) (Becker, Schuster,

Jugdutt, Hutchins, & Bulkley, 1983; F. J. Giordano et al., 1996; G. C. Hughes et al., 2003; Laham,

Rezaee, et al., 2000; Radke et al., 2006; Unger et al., 1994). The ameroid placement recapitulates the gradual occlusion of coronary arteries that occurs in human patients of CAD.

Animal models of peripheral ischemia

The capacity of the pre-formed collateral network was also determined in the periphery

(limbs) of animal models. As was observed in the heart, the extent of the pre-existing collateral

49

circulation of the hindlimb is heterogeneous between animal species (Teunissen et al., 2012;

Waters, Terjung, Peters, & Annex, 2004). After acute femoral artery occlusion, mice, rabbits, and

dogs have low collateral flow (1-7%) when compared to pigs and rats (10-30%) (Hoefer, van

Royen, Buschmann, Piek, & Schaper, 2001; MacGabhann & Peirce, 2010; Rakue et al., 1998;

Teunissen et al., 2012). The slight baseline collateral formation in mice, rabbits and dogs is similar

to that observed in human PAD patients. A large proportion of preclinical and proof-of-concept

studies targeting therapeutic revascularization in PAD has focused on rodents (Haas et al., 2012;

Waters et al., 2004). Surgical ligation and/or excision of the common femoral or iliac arteries have

been reported to induce limb ischemia in these models. Due to the considerable level of pre-formed

collaterals in the rat, ligation of the femoral artery does not affect resting blood flow but does limit

muscle perfusion upon exercise activity (G. L. Tang, Chang, Sarkar, Wang, & Messina, 2005;

Teunissen et al., 2012). This model mimics the intermittent claudication in patients with PAD

whose symptoms are only observed during activity but not at rest (Mascarenhas et al., 2014; Muir,

2009; Muller et al., 2013). The rabbit and mouse are well-characterized models of resting vascular insufficiency that have been employed to test therapeutics for PAD (Couffinhal et al., 1998;

Dragneva et al., 2013; Haas et al., 2012; Madeddu et al., 2006; Pu et al., 1994).

LMW FGF2 gene or protein delivery in chronic ischemia

The potential of FGF2 as an agent of therapeutic revascularization has been examined and reported in several preclinical models of myocardial and limb ischemia. The results of these studies are summarized in Tables 5 and 6.

For the myocardial ischemia studies, canine or porcine models were treated with human recombinant LMW FGF2 protein exogenously administered by injection to the pericardium or

50

myocardium or by infusion through the coronary arteries or (Laham, Rezaee, et al., 2000;

Lazarous et al., 1995; Rajanayagam et al., 2000; Unger et al., 1994; Watanabe et al., 1998). In all

but one of these cases, LMW FGF2 was delivered to the ischemic myocardium 7-14 days after the

ameroid placement surgery. The effects of the growth factor on infarct size, regional blood flow

and left ventricular (LV) function, were evaluated from 2 to 4 weeks after growth factor treatment.

The exception to this course of study was the work done by Yanagisawa-Miwa and co-workers where intracoronary infusion of LMW FGF2 was performed within hours of the occlusion of the coronary arteries in dogs which more closely resembles an acute ischemia model and not the human disease more accurately modeled by use of the ameroid occluders (Yanagisawa-Miwa et al., 1992). Furthermore, improvements in LV function and reduction in infarct size were observed after 7 days of ischemia. As previously established in “Mechanisms of vascular formation and growth” section, the vascular repair response in chronically ischemic tissues typically begins after

7-10 days lasting up to 21 days. The positive effects of LMW FGF2 observed only 1 week after treatment suggests an acute non-mitogenic effect of FGF2 on vasodilation and may not be a true revascularization response which requires 1-2 weeks to develop. A study by Tiefenbacher and colleagues showed that exogenous LMW FGF2 stimulates vasodilation of isolated coronary arterioles in a dose-dependent manner (Tiefenbacher & Chilian, 1997). Similar results of FGF2- induced vasoactivity was present in isolated cremaster muscle (H. M. Wu, Yuan, McCarthy, &

Granger, 1996) or intact pial arterioles (Rosenblatt, Irikura, Caday, Finklestein, & Moskowitz,

1994) and perfused whole hearts (Hampton et al., 2000). These responses were FGFR-mediated and were dependent on endothelium-derived NO production (Z. Huang, Chen, Huang, Finklestein,

& Moskowitz, 1997; Tiefenbacher & Chilian, 1997).

51

In canine or rodent models, LMW FGF2 protein was exogenously administered via intra-

arterial, intravenous or intramuscular injections into the ischemic hindlimbs for a period ranging

from 1 week to 4 weeks after the hindlimb ischemia surgery. The effect of LMW FGF2 on indices

of angiogenesis and/or arteriogenesis was evaluated at 7-42 days, postoperatively. These include

the number of angiographically-visible vessels (collaterals), collateral ischemic/normal limb

perfusion (blood flow) ratio, and microvascular density determined by histology.

The studies described above and in Table 5 and 6 demonstrated a positive role for FGF2

(the LMW isoform) in promoting vascular growth/remodeling and improving post-ischemic blood

flow. In the decade between the late 1980s and late 1990s, many studies of similar nature were

undertaken for a variety of angiogenic factors, including other members of the FGF family,

vascular endothelial growth factor (VEGF), hepatocyte growth factor (HGF), and platelet-derived growth factor (PDGF), [reviewed here (Haas et al., 2012; G. C. Hughes et al., 2003; Katoh, 2013;

Madeddu et al., 2006; Waters et al., 2004)]. Several studies have performed comparisons between

FGF2 and other angiogenic factors using myocardial and peripheral ischemia models. The results of these studies are summarized in Table 7. Taken together, these studies suggested that FGF2 and

HGF stimulated the development of angiographically-visible or α-SMA-positive vessels to a similar degree. Furthermore, FGF2 was more efficacious at inducing collateral vessels compared to VEGF or several PDGF isoforms.

52

Assessments of vascular FGF Model Species Delivery Findings Reference growth and tissue function

Thrombus injection in LVG, CAG, Histology, ↑LVEF, ↓Infarct size, (Yanagisawa- FGF2 Dog IC coronary artery Infarct size after 1 week ↑Capillaries and arterioles Miwa et al., 1992)

↑BF at 2 or 4 weeks, ↑Vessels Ameroid constriction Serial collateral BF, (Unger et al., FGF2 Dog IC ≥20µm, ↔Capillaries, ↑EC of the LCx Histology after 4 weeks 1994) proliferation, acute vasodilation

Ameroid constriction Serial collateral BF, ↑BF at 2 weeks, ↓MAP, (Lazarous et al., FGF2 Dog IC of the LCx Histology after 9 weeks ↔Infarct area, ↑Vessels >10µm 1995)

Ameroid constriction Single Bolus, IV, ↑BF in IC group, ↔BF in bolus, (Rajanayagam et FGF2 Dog BF, infarct size of the LCx Pericardial, or IC pericardial or IV al., 2000)

(Watanabe et al., FGF2 LAD embolization Pig Myocardial Injection Histology after 4-5 weeks ↔BP, ↔Capillaries, ↑Arterioles 1998)

↔MAP, ↑BF, ↑LV function, Ameroid constriction BF, CAG, MRI, Histology (Laham, Rezaee, FGF2 Pig Pericardial Injection ↑Collaterals, ↓Infarct zone, of the LCx after 4 weeks et al., 2000) ↑Capillaries

Table 5: Preclinical studies with the Fgf2 gene or LMW FGF2 in models of myocardial ischemia. BF, blood flow measured by microspheres;

CAG, Coronary angiography; IC, intracoronary; LAD, left anterior descending coronary artery; LCx, left circumflex coronary artery;

LVEF, left ventricular ejection fraction; LVG, left ventriculography; MAP, mean arterial pressure. ↓, Decrease; ↑, Increase; or ↔, No

change from control.

53

Methods for assessing FGF Model Species Delivery Findings Reference vascular growth

Bilateral caudal and TcPO2, Angiography, ↑Angiographic collaterals, ↑Calf or (Baffour et al., FGF2 Rabbit IM injection iliac artery ligation Histology thigh TcPO2, ↑Capillaries 1992) ↑BPR, ↑Angiographic score, Unilateral femoral Doppler probe, FGF2 Rabbit IA injection ↔Collateral diameter, ↑Capillaries (Asahara 1995) artery excision Angiography, Histology at day 30 BF, Muscle function, ↑Calf BF, ↑Muscle function, Bilateral femoral artery FGF2 Rat IA infusion Vascular casting, ↑Collateral vessels, ↑Capillaries at (Yang et al., 1996) ligation Histology day 28 IM injection Unilateral femoral and slow Angiography, ↑Angiographic collaterals, (J. Zhang et al., FGF2 Rat artery ligation release Histology, LDPI ↑Arterioles, ↔ BF at day 42 2009) polymer Angiography, ↑Angiographic collaterals, Unilateral femoral Histology, (Rakue et al., FGF2 Dog IV injection ↑Collateral diameter, ↑Collateral artery ligation Hemodynamic flow 1998) flow volume at day 28 probe

Table 6: Preclinical studies on the effects of FGF2 in models of peripheral ischemia. BF, muscle blood flow measured by microspheres;

BPR, ischemic limb/normal limb systolic pressure ratio; IA, intra-arterial; IM, intramuscular; Laser Doppler Perfusion Imaging, LDPI;

TcPO2, transcutaneous oximetry. ↓, Decrease; ↑, Increase; or ↔, No change from control.

54

Assessment of Model Species FGF2 PDGF-AA PDGF-AB PDGF-BB VEGF-A HGF Reference vascular growth (H. Lu et LAD ligation Pig α-SMA+ vessels +++ NT NT ++ NT NT al., 2007) (Hao et al., LAD ligation Rat α-SMA+ vessels ++ NT NT +++ NT NT 2004) Femoral artery (R. Cao et Rat α-SMA+ vessels ++++ NT NT +++ NT NT ligation al., 2003) Femoral artery Angiographic (J. Zhang et Rat +++ +++ +++ NT NT NT ligation collateral vessels al., 2009)

Femoral artery (J. Zhang et Rat α-SMA+ vessels +++ + ++ NT NT NT ligation al., 2009) Femoral artery Angiographic (R. Cao et Rabbit +++ NT NT ++ NT NT ligation collateral vessels al., 2003) Femoral artery (J. Li et al., Rabbit α-SMA+ vessels +++ NT NT +++ NT NT ligation 2010) Femoral artery (Masaki et Mouse α-SMA+ vessels ++++ NT NT NT ++ NT ligation al., 2002) Femoral artery (Marui et Mouse α-SMA+ vessels +++ NT NT NT NT +++ ligation al., 2005)

Table 7: Efficacy of LMW FGF2 gene or recombinant protein relative to proangiogenic growth factors in models of myocardial and peripheral ischemia. LAD, left anterior descending coronary artery; α-SMA, alpha smooth muscle actin; NT, not tested

55

F. FGF2 therapy in PAD and CAD (Clinical studies)

The plethora of preclinical data revealing the beneficial effects of treating animal models

of ischemia with FGF2 and other growth factors propelled the field to attempt to translate these

results to CAD and PAD patients. To date, six human trials for recombinant FGF2 (LMW isoform)

have been conducted (Aviles, Annex, & Lederman, 2003). Four of these trials were Phase I/II

designed to evaluate the safety profile of FGF2 after intra-arterial, intracoronary, or intramuscular

administration (Laham et al., 1999; Laham, Chronos, et al., 2000; Lazarous et al., 2000; Unger et

al., 2000). None of these trials had more than 60 patients, and no signs of non-target organ

angiogenesis, malignancy, or altered vascular permeability were observed. However, in two of the

Phase I/II trials, low rates of proteinuria and resolving incidents of hypotension were observed

(Laham, Chronos, et al., 2000; Unger et al., 2000).

The two larger randomized, double-blinded, placebo-controlled studies were each aimed

at patients of PAD or CAD (Table 8). The Therapeutic Angiogenesis with FGF-2 for Intermittent

Claudication Trial (TRAFFIC) tested the efficacy of bilateral intra-arterial infusion placebo, single

bolus or double bolus of 30 µg/kg of FGF2 in 190 patients with moderate-to-severe IC (PAD)

(Lederman et al., 2002). The primary endpoint was change in peak walking time (PWT) at 90 days.

In a non-parametric intention-to-treat analysis (not the pre-specified ANOVA) of the results, single

bolus treatment with FGF2 increased PWT and no change was observed with double bolus FGF2

(Aviles et al., 2003; Lederman et al., 2002). Secondary outcomes of the study, claudication onset

time or quality of life measures, did not show improvement with FGF2 treatment. The FGF

Initiating RevaScularization Trial (FIRST) included 337 patients with CAD that were not eligible

for surgical revascularization (Simons et al., 2002). Patients received placebo or one of three doses

(0.3, 3 or 3 µg/kg) of FGF2 by intracoronary injection. The primary outcome for the study was

56

change (improvement) in exercise tolerance test (ETT) time from baseline at 90 days after

treatment. While ETT was the standard clinical endpoint for CAD, it is not ideal as it does not

measure myocardial function but skeletal, pulmonary and vascular function and serves as a clinical

surrogate for vascular growth and reduction in myocardial infarction (Annex & Simons, 2005;

Aviles et al., 2003; Silvestre et al., 2013). ETT times for any FGF2 dose group were not

significantly different from placebo. Secondary endpoints, which include change in ETT time at

180 days and change in stress perfusion defect (measured by single-photon emission computed tomography, SPECT), were also not changed relative to placebo at any dose of FGF2.

57

Primary FGF Drug Clinical Disease n Delivery Endpoint, Outcomes Reference (trial name) Formulation Phase followup

FGF2 Hr LMW IC Intra-arterial PWT and ABI, 90 ↑in PWT w/ single bolus, ↔ (Lederman et II 190 (TRAFFIC) FGF2 (PAD) infusion days w/ double bolus al., 2002)

FGF1 NV1 DNA II CLI Intramuscular Ulcer healing, 25 (Nikol et al., 125 No diff from placebo (TALISMAN) plasmid DB, R (PAD) injection weeks 2008)

FGF1 CLI Intramuscular Major amputation (Belch et al., NV III 525 No diff from placebo (TAMARIS) (PAD) injection or death, 1 year 2011)

Hr LMW II (Simons et al., FGF2 (FIRST) CAD 337 Intracoronary ETT, 90 days No diff from placebo FGF2 DB, R 2002)

FGF4 Angina Treadmill exercise (Henry et al., Ad5FGF4 III/III 532 Intracoronary No diff from placebo (AGENT3/4) (CAD) time, 12 weeks 2007)

Table 8: Summary of clinical trial results of FGFs in CAD and PAD patients. ABI, ankle brachial index; Ad5, adenovirus type

5; CLI, critical limb ischemia; DB, R double blind, randomized; ETT, exercise treadmill test time; IC, intermittent

claudication; Hr, Human recombinant; NV, Non-viral. ↑, Increase; ↔, No change from placebo

58

6. Dissertation focus and hypotheses

FGF2 is a potent stimulator of wound healing, tissue repair and vascular growth processes

(Grothe, Haastert, & Jungnickel, 2006; Masahiro Murakami & Sakurai, 2012; Ortega, Ittmann,

Tsang, Ehrlich, & Basilico, 1998; Presta et al., 2005). These properties put forward FGF2 as

prototypic growth factor for use in the study of therapeutic revascularization (Katoh, 2013;

Murakami & Simons, 2008). Use of exogenous LMW FGF2 in in vitro and in vivo vascular growth studies has revealed a function for the growth factor in regulating extracellular matrix remodeling, blood vessel destabilization, endothelial and vascular smooth muscle cell migration, proliferation, and tube formation; all of which are critical aspects of the vascular growth processes including angiogenesis and arteriogenesis (Laham et al., 1999; Presta et al., 2005).

The central hypothesis of this dissertation is that the two classes of FGF2 isoforms have

distinct roles in chronic ischemia-induced vascular and tissue repair/remodeling. There is recent,

and still emerging, evidence to suggest that the two classes of FGF2 isoforms mediate cellular

pathways that both converge and diverge in their modulating effects on vascular biology, cardiac

injury, bone development/remodeling, and neuronal development/homeostasis (Coffin et al., 1995;

Garmy-Susini et al., 2004; Grothe et al., 2006; Jiang et al., 2007; Liao et al., 2009). Further

evidence to support the over-arching hypothesis that the isoforms have functionally distinct roles

include their different translational mechanisms, subcellular and subnuclear distribution, factors

controlling release/secretion/trafficking from cells, gene expression profiles and receptor

activation (Claus et al., 2003; Liao et al., 2009; Ma et al., 2007; Quarto et al., 2005). This

dissertation project contains three working hypothesis.

59

Hypothesis 1: The HMW and LMW FGF2 isoforms promote differential levels of

revascularization and tissue recovery in response to chronic ischemia of skeletal muscle

The FGF2 isoforms have been shown to have contrasting functions in protecting the heart

from cardiac dysfunction but have similar roles in preserving cardiomyocyte viability during acute

ischemia/reperfusion injury (Liao et al., 2007, 2010). The opposing activities of the FGF2 isoforms have also been observed in estrogen-stimulated endothelial cell proliferation and migration

(Garmy-Susini et al., 2004). To establish the specific roles of endogenous LMW FGF2 and HMW

FGF2 isoforms in chronic ischemia-induced revascularization, wildtype (WT) mice, mice with a targeted deletion of the Fgf2 gene (Fgf2-/-), mice with specific ablation of the 21kDa and 22kDa

FGF2 isoforms (FGF2 LMW-only) and mice deficient in 18kDa expression (FGF2 HMW-only) will be subjected to a model of unilateral hindlimb ischemia. This is an in vivo model where vascular perfusion of the lower hindlimb is interrupted leading to chronic ischemia of the underlying skeletal tissue. Surrogate outcome measures for functional recovery (active movement scores and the extent of toe necrosis) of the ischemic limb were assessed. Revascularization

(angiogenesis and arteriogenesis; collateral vessel formation) will be evaluated by histological quantification of capillary and arteriole vessel density, respectively and micro-computed topography. Functional determinants of post-ischemic recovery of skeletal muscle perfusion and arterial conductance were determined via serial Doppler imaging measurements of lower limb

(foot) blood flow.

Hypothesis 2: The LMW and HMW isoforms differentially modulate FGF2 expression and

FGF receptor activation in addition to stimulating a unique set of angiogenic growth factor and cytokines in response to chronic hindlimb ischemia

60

The expression of Fgf2 gene and protein isoforms has been shown to be tissue- and context- specific (Claus, Werner, et al., 2004; Coffin et al., 1995; S. Giordano et al., 1992; Liao et al., 2009).

In the peripheral nervous system (PNS), Fgf2 gene and/or protein expression as well as distribution are altered in response to nerve lesion injury. FGF2 protein expression is also upregulated under cardiac stress conditions including low-flow ischemia and adrenergic overstimulation (House et al., 2003; Padua & Kardami, 1993). To demonstrate the effects of deletion of a specific FGF2 isoform in the absence or presence of chronic ischemia on overall Fgf2 expression in skeletal muscle, the expression levels of Fgf2 transcript and protein, LMW FGF2 and HMW FGF2 were evaluated in non-ischemic, sham-operated or ischemic WT, Fgf2-/-, FGF2 LMW-only (HMWKO) and FGF2 HMW-only (LMWKO) hindlimbs.

In the heart and brain, the differential effects of the FGF2 isoforms have been studied under

acute and chronic injury, respectively. These studies have revealed that the activation of FGF2

signaling pathways are different in the presence of only the LMW or only the HMW isoforms

(Grothe et al., 2006; Kardami et al., 2007; Liao et al., 2009; Nindl et al., 2004). These processes

are paracrine, autocrine and receptor-dependent (LMW) or nuclear/intracrine and receptor-

independent (HMW). To identify the role of FGF receptor expression and activation in HMW and

LMW FGF2 responses to chronic ischemia, skeletal muscle isolated from non-ischemic, sham-

operated and ischemic hindlimbs of WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice

will be analyzed for Fgfr1b, Fgfr1c, Fgfr2b, Fgfr2c, Fgfr3b, Fgfr3c, or Fgfr4 mRNA and protein

expression levels.

In addition to its own effect on vascular growth, FGF2 can regulate the expression and

release of other angiogenic cytokines, which in turn, have their own specific roles in

61

revascularization (Murakami & Simons, 2008). This positions FGF2 as a “master regulator” of ischemia-induced revascularization. In cultured NIH 3T3 fibroblasts, HMW FGF2-transfected cells display a phenotype distinct from LMW FGF2-expressing cells by inducing the expression of a unique transcriptome (Quarto et al., 2005). To characterize the repertoire of angiogenic cytokines whose protein expression is altered in response to FGF2 isoform-stimulated vascular growth and remodeling, WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only sham-operated or ischemic muscle were evaluated using an angiogenesis proteome array.

During ischemia, angiogenic factors activate the microvascular endothelium to express cell-adhesion molecules and chemokines. These chemoattractants then recruit and activate inflammatory cells to invade the site of injury (Silvestre et al., 2008). Equally, inflammatory cells release angiogenic cytokines that can also regulate the revascularization process (Heil & Schaper,

2004; la Sala et al., 2012). A cross-talk exists between inflammatory and angiogenic responses during FGF2-mediated revascularization (Andrés et al., 2009; Presta et al., 2005; Presta, Andrés,

Leali, Dell’Era, & Ronca, 2009). Inflammatory cells can also play a role that is counter to vascular reparation by releasing oxidative metabolites that are cytotoxic and cytolytic to ischemic muscle

(Tidball, 2011). To establish the role of inflammatory cell infiltrates in FGF2 isoform-induced repair and regeneration, markers of inflammation were measured during several timepoints of ischemia in WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only non-ischemic, sham or ischemic skeletal muscle.

Hypothesis 3: The effect of the HMW and LMW isoforms FGF2 on ischemia-induced revascularization and tissue recovery is dependent upon differential modulation of tissue regeneration pathways

62

In the early periods following skeletal muscle injury (chronic ischemia), necrotic tissue

clearance is required for muscle healing and the re-establishment of muscle homeostasis. This stage of injury also coincides with activation and proliferation of quiescent resident progenitors called satellite cells that differentiate into mature myofibers (Le Grand & Rudnicki, 2007; Meisner,

Annex, & Price, 2014; Tedesco, Dellavalle, Diaz-Manera, Messina, & Cossu, 2010). The stages of the satellite cell activity (proliferation, migration and differentiation) are controlled by a family of muscle-specific transcription factors known as the myogenic regulatory factors (MRFs)

(Ceafalan, Popescu, & Hinescu, 2014). FGF2 and other growth factors have been shown to function as “wound hormones” when released from degenerating myocytes where they can stimulate the satellite cell activation which leads to muscle regeneration (Allen & Boxhorn, 1989;

Anderson, Liu, & Kardami, 1991; DiMario, Buffinger, Yamada, & Strohman, 1989). To examine

the contribution of the endogenous LMW and HMW isoforms of FGF2 to satellite cell activation

and differentiation in ischemic muscle injury and recovery, expression of several myogenic

regulatory factors (MRFs) , markers of satellite cell activity was determined. Transcript levels of

Pax3, Pax7, MyoD, myogenin, Myf5 and MRF4 was quantified in non-ischemic, sham-operated or

ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only hindlimbs.

This dissertation research project was designed to provide novel evidence of the specific

roles of the endogenous FGF2 isoforms in ischemia-induced revascularization, in particular, HMW

FGF2 in ischemic-induced revascularization. Additionally, the molecular signaling pathways that

mediate the effect of the FGF2 isoforms on vascular and tissue regeneration was explored. These

findings contribute to the understanding of the FGF2 isoforms in vascular biology and

pathophysiology. Furthermore, they may one day lead to the development of novel therapies for

the treatment of coronary and peripheral arterial diseases.

63

MATERIALS AND METHODS

Animals and Exclusion Criteria

Mice were housed in a pathogen-free facility and handled in accordance with standard use

protocols, animal welfare regulations, and the NIH Guide for the Care and Use of Laboratory

Animals. All protocols were approved by the University of Cincinnati Institutional Animal Care

and Use Committee. All mice were bred on a (50%) Black Swiss and (50%) 129/Sv mixed

background. Any animals with severe necrosis and/or auto-amputation of tissue in the surgical

limb were excluded from the study and prematurely euthanatized.

Generation of Fgf2-/- mice

Mice with a targeted ablation of the Fgf2 gene (Fgf2-/-) were generated using a tag and

exchange construct by Ming Zhou in the laboratory of Dr. Thomas Doetschman as previously

described (Zhou et al., 1998). Briefly, a 0.5-kb NarI/XbaI portion of the Fgf2 gene containing the

proximal promoter region and the entire first was replaced with an Hprt (hypoxanthine

phosphoribosyl transferase) minigene in E14TG2a embryonic stem (ES) cells derived from 129

strain mice. A herpes virus thymidine kinase gene inserted outside of the homologous region at

the 3’ end of the targeting construct was used for negative selection. ES cells with the target vector

inserted were selected with HAT (hypoxanthine, aminopterin, thymidine) medium. Loss of the

thymidine kinase gene was selected with the use of gancyclovir. Double-resistant cells were confirmed for the targeted allele with polymerase chain reaction (PCR) and Southern blotting.

The targeted ES cells were then microinjected into C57BL/6 blastocysts and chimeric mice were generated from foster mothers. Germline transmission was tested by breeding the chimeric mice with 50% 129 and 50% Black Swiss mice. Wildtype (+/+), heterozygous (+/-) and knockout (-/-)

64

offspring produced from this breeding and other subsequent pairings were identified with PCR

analysis and loss of FGF2 expression was confirmed with Southern, Northern and Western blotting

(Figure 3).

For genotyping with PCR, DNA was isolated from mouse pup tail clips that were placed

in a digestion buffer (100 mM NaCl, 10 mM Tris-Cl pH 8, 25 mM EDTA, 0.5% SDS) along with

100 µg proteinase K (Roche) at 60oC overnight. After centrifugation to remove cellular debris, the

DNA was precipitated in 100% isopropyl alcohol and washed with 70% ethanol. The pelleted

DNA was then air dried and dissolved in TE buffer overnight or at 55oC for 2 hours. PCR was

performed with two sets of primers using Taq DNA polymerase (Roche): wildtype allele, 5’- CGA

GAA GAG CGA CCC ACA C - 3’, 5’- CCA GTT CGG GGA CCC TAT T - 3’ and knockout

allele, 5’- AGG AGG CAA GTG GAA AAC GAA - 3’, 5’- CCC AGA AAG CGA AGG AAC

AAA - 3’. 1 µL of DNA was added to a 19 μL PCR master mixture (Roche 1.5 mM Mg PCR

buffer, 2.5 mM dNTP, 5X Cresol Red, H2O, and Taq DNA polymerase) and the following

amplification reaction was carried out. 1 cycle of (95°C for 90 seconds), 35 cycles of (95°C for 30

seconds for denaturing, 58 °C for 50 seconds for annealing and 72°C for 90 seconds for

elongation), followed by 1 cycle of 95°C for 30 seconds, 58°C for 50 seconds and 72°C for 10

minutes. The PCR products were run on a 1% agarose gel for 30 minutes at 140 mV. A band at

1299 base pairs (bp) signifies the presence of the targeted (knockout) allele while the wildtype

allele was indicated by a 185bp product.

Generation of FGF2 HMW-only (LMWKO) mice

65

FGF2 HMW-only (LMWKO) mice were generated in the laboratory of Dr. Thomas

Doetschman as previously described (Azhar et al., 2009; Garmy-Susini et al., 2004). Briefly, the

Hprt minigene in the targeting vector described for the ablation of the Fgf2 gene was exchanged

by gene targeting with a mutated exon 1 where the wildtype CCATGC sequence is replaced with

CTGCAG. This replaces the ATG translational start site for FGF2 LMW protein with a CGA as

well introducing a PstI site that serves as a diagnostic for the exchanged allele. After selecting for

loss of the Hprt gene with 6-thioguanine (6-TG), ES cells carrying the exchanged allele were

microinjected into C57BL/6 blastocysts and chimeric mice were generated from foster mothers.

Germline transmission was tested by breeding male chimeric mice with 50% 129 and 50% Black

Swiss mice. Wildtype (+/+), heterozygous (+/-) and knockout (-/-) offspring produced from this

breeding and other subsequent pairings were identified with PCR analysis and absence of LMW

isoform expression was confirmed with Western blotting (Figure 3). The colony was subsequently

maintained on a mixed (50%) 129/Sv x (50%) Black Swiss background (Garmy-Susini et al.,

2004).

For genotyping with PCR, DNA was isolated from mouse pup tail clips that were placed

in a digestion buffer (Qiagen Puregene Cell Lysis Solution) along with 30µg proteinase K (Roche) at 60oC overnight. After treatmet with Qiagen Puregene Protein Precipitation Solution to remove

protein and cellular debris, the DNA was precipitated in 100% isopropyl alcohol and washed with

70% ethanol. The pelleted DNA was then air dried and dissolved in Qiagen Puregene DNA

Hydration solution overnight or at 55oC for 2 hours. PCR was performed with Epicentre Tfl

polymerase and the following primer set: 5’- CCC GCA CCC TAT CCT TAC ACA - 3’and 5’-

GCC GCT TGG GGT CCT TG - 3’ to amplify both the wildtype and mutated alleles. 1 µL of

DNA was added to a 19 μL PCR master mixture (Epicentre 20X buffer, Epicentre 25mM Mg2+,

66

Epicentre 10X Enhancer, 2.5 mM dNTP, 5X Cresol Red, H2O, and Tfl DNA polymerase) and the following amplification reaction was carried out. 1 cycle of 97˚C for 5 minutes, 35 cycles of 97˚C for 60 seconds for denaturation, 58˚C for 60 seconds for annealing and 72˚C for 2 minutes for elongation and followed by 1 cycle of 72oC for 10 minutes. The PCR products were incubated for

1 hour at 37oC with PstI restriction enzyme, Buffer 3 and BSA (New England Bio Labs). The digested PCR products were run on a 2% agarose gel for 40 minutes at 140 mV. A band at 566 bp signifies the presence of the wildtype allele while two bands at 476 bp and 90 bp indicate the presence of the knockout allele.

Figure 3. FGF2 protein isoform expression in non-ischemic skeletal muscles. Representative immunoblots of LMW (18 kDa) or HMW (21 kDa and 22 kDa) FGF2 isoforms in non- ischemic wildtype (WT), Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only skeletal muscles.

67

Generation of FGF2 LMW-only (HMWKO) mice

FGF2 LMW-only (HMWKO) mice were generated in the laboratory of Dr. Tom

Doetschman as previously described (Azhar et al. 2009). Briefly, the Hprt minigene in the

targeting vector described for the ablation of the Fgf2 gene was exchanged by insertion of a 14 bp

oligonucleotide (CTAGTCTAGACTAG) that contains a XbaI into a SmaI site located upstream

of the LMW FGF2 translational start site. This insertion introduced stop codons in all reading

frames and a frameshift mutation of all translation at the alternative start sites that lie upstream of

the LMW FGF2 AUG start codon. Loss of the Hprt gene was selected for in ES cells resistant to

6-TG and was confirmed with Southern blotting, DNA sequencing, PCR, and XbaI digestion. The

cells were then microinjected into C57BL/6 blastocysts and chimeric mice were generated from

foster mothers. Germline transmission was tested by male breeding the chimeric mice with female

Black Swiss mice. Wildtype (+/+), heterozygous (+/-) and knockout (-/-) offspring produced from

this breeding and other subsequent pairings were identified with PCR analysis. Insertion of the

14bp oligonucleotide and the XbaI was confirmed by Southern blotting and DNA sequencing

respectively. Loss of HMW FGF2 expression was confirmed by Western blotting (Figure 3).

For genotyping with PCR, DNA was isolated from mouse pup tail clips that were placed

in a digestion buffer (Qiagen Puregene Cell Lysis Solution) along with 30µg proteinase K (Roche) at 60oC overnight. After treatmet with Qiagen Puregene Protein Precipitation Solution to remove

protein and cellular debris, the DNA was precipitated in 100% isopropyl alcohol and washed with

70% ethanol. The pelleted DNA was then air dried and dissolved in Qiagen Puregene DNA

Hydration solution overnight or at 55oC for 2 hours. PCR was performed with Epicentre Tfl

polymerase and the following primer set: (Forward 5’- CCC AAG AGC TGC CAC AG - 3’ and

Reverse 5’-CGC CGT TCT TGC AGT AGA G -3’) to amplify both the wildtype and knockout

68

alleles. 1 µL of DNA was added to a 19 μL PCR master mixture (Epicentre 20X buffer, Epicentre

2+ 25 mM Mg , Epicentre 10X Enhancer, 2.5mM dNTP, 5X Cresol Red, H2O, and Tfl DNA

polymerase) and the following amplification reaction was carried out. 1 cycle of 97˚C for 5

minutes, 35 cycles of 97˚C for 60 seconds for denaturation, 58˚C for 60 seconds for annealing and

72˚C for 2 minutes for elongation and followed by 1 cycle of 72oC for 10 minutes. The PCR products were run on a 4% agarose gel for 2.5 hours at 75mV. A band at 166 bp signifies the presence of the knockout allele while a 152 bp band indicates the presence of the wildtype allele.

Hindlimb ischemia surgery

Age- (10 -12 weeks) and sex-matched WT, Fgf2-/-, FGF2 LMW-only, and FGF2 HMW- only mice were anesthetized with 2.5% Avertin (2, 2, 2-tribromoethanol and tertiary amyl alcohol in 0.9% saline) at a dose of 0.02 mL/g, i.p and placed on a heating pad at 37oC. Unilateral hindlimb

ischemia was induced surgically as previously described (Couffinhal et al. 1998; Sullivan et al.

2002). A longitudinal incision was made in the skin overlying the middle portion of the left

hindlimb to expose tissue . The femoral artery and vein were then exposed and dissected free of

the associated nerve (Figure 4). Preservation of the femoral nerve was important to prevent atrophy of the muscle fibers. The vessel pair was ligated proximal to the external iliac artery and distally

at the bifurcation into the saphenous and popliteal branches. The intervening vessel segment was

then excised. The overlying skin was then closed with 5-0 silk sutures and tissue adhesive. A sham

procedure was performed on the contra-lateral (right) hindlimb, wherein a longitudinal skin incision and sutures were placed without ligation.

69

A

B.

Femoral artery, Nerve vein & nerve

R L

Figure 4: The mouse hindlimb ischemia model. A) Schematic for the surgical induction of

hind limb ischemia. Modified from (Sullivan et al., 2002) B) Photograph of mouse lower

limbs after surgical excision of femoral artery-vein pair with intact nerve in ischemic right

hind limb (R) and intact femoral artery-vein-nerve in the sham limb (L).

70

Assessment of necrosis

Necrosis was defined as the apparent darkening of one or more toes on the ischemic foot.

Darkening of ischemic feet began as a redness and in some cases, discoloration progressed to black

and ischemic toe(s) were lost. These mice were characterized as having necrotic limbs.

Assessment of limb function

The incidence of hindlimb tissue necrosis and limb function was assessed daily up to 14

days of ischemia and then weekly up to 42 days of ischemia. Functional use of the ischemic limb

was measured on the following scale modified from (Stabile et al., 2003) (4 = dragging of foot, 3

= no dragging of foot but no plantar flexion, 2 = plantar flexion but no flexing of toes and 1=

normal function, with flexing of toes to resist gentle traction on the tail).

Treadmill exercise testing

Mice were exercised on a modified OmniPacer LC4 motor-driven treadmill apparatus

(Omnitech, Columbus, OH) (Fewell et al., 1997; Lerman et al., 2002). The four-lane mouse treadmill was equipped with belts of adjustable speeds (0-100 m/min) and shock bars of adjustable amperage (0-2 mA) at the rear of each belt to stimulate the mice to run. Two pairs of infrared (IR) emitters/detectors were placed across from one another at the head of the shock bars. When the double IR beam was blocked by a mouse, a trip signal was generated and recorded. The trip signals

(beam breaks) were recorded every second for the length of each mouse’s exercise run. For each run, the number of breaks in the detection beam is a measure of the exercise performance of the animal. Therefore, the higher the number of beam breaks per minute, the less able the mice are to keep pace with the belt speed. Before each exercise test, the treadmill was placed at a 7o incline

71

and the mice were acclimated over a 2-day period. On the first day of acclimation, mice were left undisturbed in the treadmill for 15 minutes with the instrument powered on, shock bars turned off and the belt speed set at 0 m/min. The belt was then turned on and speed was set at 2 m/min for 2 minutes, 5 m/min for 5 minutes and 7 m/min for 10 minutes. On the second day of acclimation, the shock grid was turned on, set at 1 mA, and the protocol from the previous day was repeated.

After the acclimation, exercise treadmill testing was performed at baseline (1-3 days before hindlimb ischemia surgery) and 14 days after surgery. Each test began with mice placed in the treadmill for 15 minutes and left undisturbed with the shock grid turned on to reduce any effect of stress on the exercise performance. The belt speed was then set at 7, 10 and 15 m/min for 5 minutes and at a maximum speed of 20 m/min for either 15 minutes or the length of time before which the animal fails to keep up the running pace. Failure to sustain the speed was indicated by the mouse lagging at the rear of the belt on the metal bars despite the stimulus of the shock current. If the animal received >10 shocks in a row, the test was halted and the mouse allowed to rest. The beam breaks were calculated over each 5 minute period of treadmill speed and over the final 15 minute run for animals able to complete the test. For animals that failed to reach 15 minutes, the beam breaks were calculated over their respective run times.

Assessment of gene expression in skeletal muscles

RNA isolation

Skeletal muscles were harvested from non-ischemic, sham or ischemic hindlimbs of WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice. Sham or ischemic thigh and calf muscles

were dissected free of the femur and tibia/fibula, respectively and snap-frozen in liquid nitrogen

at 1, 3, 7 and 14 days of ischemia. For quantitative real-time polymerase chain reaction (qRT-

72

PCR), 50-80 mg of skeletal muscle was placed in 1 mL of Tri-Reagent (Molecular Research

Center) for RNA isolation. The tissue was immediately homogenized with a benchtop motorized homogenizer, after which 1 mL of the homogenate was transferred to 1.5 mL microcentrifuge tube and centrifuged at 12,000x g at 4oC for 10 minutes. A phase separation to extract the RNA was performed with 100 µL of 1-bromo-3-chloropropane (BCP, Molecular Research Center) followed with centrifugation at 12,000x g at 4oC for 15 minutes. The aqueous phase was transferred into a new 1.5 mL microcentrifuge tube where the RNA was precipitated with 500 µL of isopropanol and centrifuged again at 12,000x g at 4oC for 10 minutes. The RNA pellet was then washed twice with 1 mL of 75% ice-cold ethanol followed by centrifugation at 12,000x g at 4oC for 4 minutes.

After air drying on ice for 5 minutes, the pellets were resuspended the RNA pellet in 10µL of nuclease-free water. To assess the integrity of the RNA, 1 µL of each RNA sample was added to

1 µL of 6X bromophenol blue and 4 µL of H2O and the solution was run on a 1% agarose gel at

90V for 40 minutes. The presence of two ribosomal bands at ~5 kilobases (28S) and at 2 kilobases

(18S) indicates intact RNA while a smear or the absence of RNA suggests that the RNA has been degraded. To determine the purity and the concentration of RNA, a 1:10 dilution of the sample in nuclease-free water was measured using a Nanodrop ND1000 spectrophotometer. The total concentration of RNA in each sample was calculated with the absorbance at 260 nm, and the purity was determined from the 260 nm/280 nm absorbance ratio. Samples with a 260/280 ratio of at least 1.8 or above were saved and stored at -80oC and utilized for reverse transcription.

Real-Time Quantitative Reverse Transcription PCR

Prior to reverse transcription of RNA to cDNA, 500 ng each of the isolated RNA was purified using RQ1 RNase-Free DNase (Promega). For reverse transcription with the SuperScript

73

II Reverse Transcriptase (Life Technologies), 250 ng RNA was incubated with 10 mM dNTPs,

o random hexamers and H2O and at 65 C for 5 minutes and placed immediately on ice. The RNA

mix was then added to 10X RT buffer, 25 mM MgCl2, 0.1 M DTT and RNAseOUT and incubated

at room temperature for 2 minutes. 1 µL of SuperScript II reverse transcriptase was added to the

mixture which was then incubated at room temperature for 10 minutes 42oC for 50 minutes

followed by 15 minutes at 70oC, 1 minute at 4oC where the reaction was paused for addition of 1

µL of RNAse H to each sample. This step degrades any remaining template RNA. To confirm the

presence of cDNA, a control RNA sample that was converted to cDNA is amplified with

conventional PCR. A PCR reaction for each of four dilutions of the cDNA (1:1000, 1:10,000,

1:100,000 or 1:1,000,000) was evaluated using Taq Polymerase (Roche) and the following reaction

mixture: 10X + Mg buffer, 2.5 mM dNTP, Primer A (5’- GAC ATG GAA GCC ATC ACA GAC

- 3’), Primer B (5’- AGA CCG TTC AGC TGG ATA TTA C - 3’) 5X Cresol Red, and H2O.

For FGF2, qRT-PCR analysis of mRNA was performed in triplicates using the comparative

cycle time (ΔΔCT) method and Step One Plus ABI Real Time PCR system. 40 cycles of

amplification was performed in a 25 μL volume containing SYBR GREEN PCR master mix (Life

Technologies) and a 10μM primer mix, containing primers for the Fgf2 gene or for the

housekeeping gene, 18S. The following PCR parameters were used to amplify the cDNA (95o for

10 minutes, 95o for 15 seconds and 60o for 60 seconds) after determining that 18S (reference gene)

expression was not altered in the Fgf2-/-, FGF2 LMW-only or FGF2 HMW-only skeletal muscle relative to WT.

For FGF receptor (FGFR) expression, qRT-PCR analysis of mRNA was performed in triplicates using the ΔCT method and Step One Plus ABI Real Time PCR system. 40 cycles of

amplification was performed in a 25 μL volume containing SYBR GREEN PCR master mix (Life

74

Technologies) and a 10μM primer mix, containing primers for one of the following Fgfr1b, Fgfr1c,

Fgfr2b, Fgfr2c, Fgfr3b, Fgfr3c, or Fgfr4 or for the 18S housekeeping gene (Table 9). The

following PCR parameters were used to amplify the cDNA (95o for 10 minutes, 95o for 15 seconds

and 60o for 60 seconds) after determining that 18S (reference gene) expression was not altered in

the Fgf2-/-, FGF2 LMW-only or FGF2 HMW-only skeletal muscle relative to WT.

For expression of the myogenic regulatory factors (MRFs), qRT-PCR analysis of mRNA

was performed in triplicates using the ΔCT method and Step One Plus ABI Real Time PCR system.

40 cycles of amplification was performed in a 25 μL volume containing SYBR GREEN PCR master mix (Life Technologies) and a 10μM primer mix, containing primers for one of the following Pax3, Pax7, Myf5, MyoD, MyoG or Mrf4 or for the 18S housekeeping gene (Table 10).

The following PCR parameters were used to amplify the cDNA (95o for 10 minutes, 95o for 15 seconds and 60o for 60 seconds) after determining that 18S (reference gene) expression was not altered in the Fgf2-/-, FGF2 LMW-only or FGF2 HMW-only skeletal muscle relative to WT.

75

Primer Sets Forward, 5’- AGT CCC TGC CCT TTG TAC ACA -3’ 18S Reverse, 5’- CCG AGG GCC TCA CTA AAC - 3’ Forward, 5’- CAA CCG GTA CCT TGC TAT GA -3’ Fgf2 Reverse, 5’- TCC GTG ACC GGT AAG TAT TG -3’ Forward 5’- CAA CTT GCC GTA TGT CCA GAT C -3’ Fgfr1b Reverse 5’- CTC CGC ATC CGA GCT ATT AA -3’ Forward 5’- GCC AGA CAA CTT GCC GTA TG -3’ Fgfr1c Reverse 5’- ATT TCC TTG TCG GTG GTA TTA ACT C-3’ Forward 5’- GGG CTG CCC TAC CTC AAG -3’ Fgfr2b Reverse 5’- CTT CTG CAT TGG AGC TAT TTA TCC -3’ Forward 5’- CCC GGC CCT TCA -3’ Fgfr2c Reverse 5’- GTT GGG AGA TTT GGT ATT TGG TT -3’ Forward 5’- GCA CGC CCT ACG TCA CTG TA -3’ Fgfr3b Reverse 5’- GCG TCT GCC TCC ACA TTC T -3’ Forward 5’- ACG GCA CGC CCT ACG T -3’ Fgfr3c Reverse 5’- CTC CTT GTC GGT GTT AGC -3’ Forward 5’- CGC CAG CCT GTC ACT ATA CAA A -3’ Fgfr4 Reverse 5’- CCA GAG GAC CTC GAC TCC AA -3’

Table 9. List of forward and reverse primers for analysis of Fgf2 and Fgf receptor expression in WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only non-ischemic, sham and ischemic skeletal muscle tissues.

76

Primer Sets Forward 5’- CAA CTT GCC GTA TGT CCA GAT C -3’ Pax3 Reverse 5’- CTC CGC ATC CGA GCT ATT AA -3’ Forward 5’- GCC AGA CAA CTT GCC GTA TG -3’ Pax7 Reverse 5’- ATT TCC TTG TCG GTG GTA TTA ACT C-3’ Forward 5’- GGG CTG CCC TAC CTC AAG -3’ Myf5 Reverse 5’- CTT CTG CAT TGG AGC TAT TTA TCC -3’ Forward 5’- CCC GGC CCT CCT TCA -3’ MyoD Reverse 5’- GTT GGG AGA TTT GGT ATT TGG TT -3’ Forward 5’- GCA CGC CCT ACG TCA CTG TA -3’ MyoG (myogenin) Reverse 5’- GCG TCT GCC TCC ACA TTC T -3’ Forward 5’- ACG GCA CGC CCT ACG T -3’ Mrf4 Reverse 5’- CTC CTT GTC GGT GGT GTT AGC -3’

Table 10. List of forward and reverse primers for analysis of myogenic regulatory factors

(MRFs) expression in WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only non-ischemic, sham and ischemic skeletal muscle tissues.

Gene expression (data analysis)

To quantify gene expression, the comparative CT method was used to measure the fold

changes in expression of Fgf2, Fgfr1-4, Pax3, Pax7, Myf5, MyoD, MyoG or Mrf4 in Fgf2-/-, FGF2

LMW-only or FGF2 HMW-only muscles relative to WT tissues (Bookout et al., 2006). Relative

expression of each gene of interest was determined relative to expression of 18S rRNA, an internal

control gene. Briefly, relative expression (ΔCT) = CTgene of interest – CT18S) and fold changes due to

(-ΔΔCT) selective protein expression = 2 (where ΔΔCT = CTFgf2-/-, FGF2 HMW-only, FGF2 LMW-only –

CTWT)(Livak & Schmittgen, 2001; Schmittgen & Livak, 2008).

77

FGF2 protein isoform expression (Western immunoblot)

Skeletal muscles were harvested from non-ischemic, sham and ischemic hindlimbs of WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice. Sham and ischemic thigh and calf muscles were dissected free of the femur and tibia/fibula, respectively and snap-frozen in liquid nitrogen at 1, 3, 7 and 14 days of ischemia. 100-120 mg of skeletal tissues were pulverized and powdered and then placed in an ice-cold glass dounce tissue grinder in an extraction buffer containing 20 mM Tris, 2 mM EDTA, 2 M NaCl, 1% NP40, EDTA-free protease inhibitor cocktail (Roche, 1

tablet/10 mL), and 1 mM PMSF. The homogenate was centrifuged at 15,000g for 15 minutes and

the supernatant was collected. Protein concentration was determined using a Bio-Rad DC protein

assay, and 2 or 5 mg (FGF2 LMW-only) of total protein was added to a 75% heparin sepharose

bead slurry (GE Healthcare). The samples and beads were incubated at 4°C for one hour, pelleted

and washed three times with a wash buffer (0.6 M NaCl and 10 mM Tris-HCl pH 7.4). 30 mL of

10X protein sample buffer was added to the beads and boiled for 10 minutes. The entire protein

sample was then loaded on a 15% SDS-PAGE gel and transferred to nitrocellulose membrane.

Transfer efficiency and loading equality were examined by staining the membrane with 0.1%

Ponceau S in 5% acetic acid. The membranes were blocked in 5% dry milk in 0.1% PBS/Tween

solution for 1 hour to prevent non-specific binding and then incubated with rabbit polyclonal primary antibody to FGF2 (1:1000 dilution, Santa Cruz Biotechnology) and incubated overnight at 4°C. After primary antibody incubation, membranes were washed and then incubated at room temperature in anti-rabbit IgG secondary antibody (1:7000 dilution, Santa Cruz Biotechnology)

two hours. The blots were then visualized using a chemiluminescence kit (Amersham ECL, GE)

and densitometry of protein bands were quantified using a Fluorchem 8800 gel imager.

78

FGF2 protein isoform expression (quantitative data analysis)

To evaluate FGF2 isoform protein expression, the intensity of each protein band was

determined as described above. For comparison of expression across all genotypes, FGF2 protein

levels in Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only were normalized to WT protein

measures. This allows for statistical assessment of the FGF2 isoforms in tissues that required

different amounts of starting protein (2 mg for Fgf2-/- and FGF2 LMW-only; 5 mg for FGF2

HMW-only).

FGFR expression and phosphorylation (Western immunoblotting)

Skeletal muscles were harvested from non-ischemic, sham or ischemic hindlimbs of WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice. Sham or ischemic thigh and calf muscles

were dissected free of the femur and tibia/fibula, respectively and snap-frozen in liquid nitrogen

at 3 and 7 days of ischemia. 2 mg of skeletal tissues were pulverized and powdered and then placed

in an ice-cold glass dounce tissue grinder in an extraction buffer containing 25 mM Hepes, 150

mM NaCl, 1% Triton X-100, 5 mM EDTA, 1% glycerol, 1 mM sodium orthovanadate, 25 mM β–

glycerolphosphate, 50 mM sodium fluoride, 0.5 mM okadaic acid, 100 mM calpain inhibitor,

Pefabloc Stock 1 and 2 (Roche), Roche phosphatase inhibitor (1 tablet/10 mL), Roche complete

mini EDTA-free protease inhibitor cocktail (1 tablet/10 mL), and 1 mM PMSF. Protein

concentration was determined using a Bio-Rad DC protein assay. 200 µg of protein was then loaded on a 8% SDS-PAGE gel and transferred to nitrocellulose membrane. Transfer efficiency and loading equality were examined by staining the membrane with 0.1% Ponceau S in 5% acetic acid. The membranes were blocked in 5% BSA in 0.1% TBS/Tween solution for 1 hour to prevent non-specific binding and then incubated overnight at 4ºC with primary antibodies against phospho-

79

Y654 FGFR1 (1:1000 dilution, Abcam), phospho-Y742 FGFR3 (1:50000 dilution, Abcam) or phospho-Y642 FGFR4 (1:500 dilution, Abcam) followed by 1 hour incubation at room temperature in anti-rabbit IgG secondary antibody (1:5000, Santa Cruz). The phosphorylation of

FGFR1, FGFR3, and FGFR4 expression was visualized by ECL or ECL Plus (phospho-FGFR3) and densitometry of the protein bands were quantitated using a Fluorchem 8800 gel imager. The membranes were then stripped of the phospho-antibodies with stripping buffer (62.5 mM TRIS, pH 6.8, 2% SDS, and 100 mM β-mercaptoethanol) for 45 minutes at 56ºC. After washing, the membranes were blocked again in 5% BSA in 0.1% TBS/Tween solution for 1 hour before incubation with primary antibodies against total FGFR1 (1:400, Abcam), FGFR3 (1:1000, Abcam) or FGFR4 (1:500, Santa Cruz) overnight at 4ºC. After primary antibody incubation, membranes then incubated for 1 hour at room temperature with anti-rabbit IgG secondary antibody (1:5000,

Santa Cruz). Total receptor expression of FGFR1, FGFR3, and FGFR4 was visualized by ECL or

ECL Plus (FGFR3) and densitometry of the protein bands were quantitated using a Fluorchem

8800 gel imager.

FGFR phosphorylation (quantitative data analysis)

To evaluate FGFR activation (i.e. phosphorylation), the intensity of each protein band was determined as described above.

80

Expression profile of angiogenesis-related proteins

Skeletal muscles were harvested from non-ischemic, sham and ischemic hindlimbs of WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice. Sham and ischemic thigh and calf muscles were dissected free of the femur and tibia/fibula, respectively and snap-frozen in liquid nitrogen at 7 days of ischemia. 100-120 mg of skeletal tissues was pulverized and homogenized on ice in a glass dounce in an extraction buffer containing 20 mM Tris, 2 mM EDTA, 2 M NaCl, 1% NP40,

EDTA-free protease inhibitor cocktail (Roche, 1 tablet/10 mL), and 1 mM PMSF. The homogenate was centrifuged at 15,000g for 15 minutes and the supernatant was collected. After protein concentration was determined using a Bio-Rad DC protein assay, 600 µg of total protein in 1.5 mL of array buffer from the proteome profiler mouse angiogenesis antibody array kit (R&D Systems) was mixed with 15 µL of a detection antibody cocktail and incubated at 25oC for one hour. Each

sample/antibody mixture was then added to a nitrocellulose membrane containing 53 different

capture antibodies (Table 11) dotted in duplicate and incubated at 4oC overnight. After washing,

2 mL of buffered Streptavidin-HRP was added to each membrane and allowed to incubate at 25oC

for 30 minutes. Array membranes were developed using chemiluminescence with the provided

Chem Reagent mix. The densitometry of the antibody/protein dots were quantified using a

Fluorchem 8800 gel imager.

81

ADAMTS1 EGF IGFBP-1 MIP-1α/CCL3 Osteopontin Amphiregulin CD105/Endoglin IGFBP-2 MMP-3 PD-ECGF Angiogenin Endostatin IGFBP-3 MMP-8 (pro) PDGF-AA Angiopoietin-1 -a IL-1α MMP-9 PDGF-BB Angiopoietin-3 FGF1 TIMP-1 IGFBP-9/NOV Pentraxin-3 Coagulation factor III/Tissue Factor FGF2 IL-10 KGF/FGF7 SDF/CXCL12 CXCL16 Fractalkine/CX3CL1 CXCL10/IP-10 CXCL4/PF4 Serpin E1/PAI-1 IGFBP-10/Cyr61 GM-CSF CXCL1, KC PIGF-2 Serpin F1/PEDF DLL4 HB-EGF Leptin/OB Prolactin Thrombospondin-2 CD26/DPPIV HGF MCP-1/CCL2 Proliferin TIMP-4 VEGF-A VEGF-B

Table 11. List of angiogenesis-related proteins analyzed for relative expression in WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only ischemic skeletal muscle tissues.

Expression profile of angiogenesis-related proteins (data analysis)

To determine the relative expression of each protein (Table 11), the intensity of each array dot was quantified as described above. Expression was determined as the average intensity of the pair of duplicate spots representing each angiogenesis related protein subtracted from an averaged background signal. For each protein, expression levels in ischemic WT, Fgf2-/-, FGF2 LMW-only, or FGF2 HMW-only arrays were normalized to the corresponding sham muscle expression.

82

Immunohistochemistry

Animals were sacrificed before (non-ischemic) or 1, 3, 7, 14, or 42 day(s) of ischemia.

Skeletal muscles harvested from the thigh (gracilis, semi-membranosus, and semi-tendinosus) and

calf (gastrocnemius) regions of the ischemic limb were immersion-fixed overnight in 4% formaldehyde followed by immersion in 70% ethanol for processing and embedding.

Griffonia simplicifolia isolectin B4 staining

Five μm thick transverse paraffin-embedded sections were deparaffinized and rehydrated with xylenes and alcohol/water (100%, 95% and 70%) solutions, respectively. A 10 minute antigen retrieval step was performed at 100oC using a pH 6.0 citric acid solution. Endogenous peroxidase

activity was blocked with H2O2 for 12 minutes followed by treatment with 5% normal goat serum

(Vector Labs) and 1% bovine serum albumin (Vector Labs) at room temperature for 1 hour to

reduce nonspecific binding. Serial sections were then incubated overnight at 4oC with biotinylated

Griffonia simplicifolia (Bandeiraea) isolectin B4 (IB4 1:100, Vector Labs) to detect endothelial

cells. IB4 from the African-based black bean Griffonia simplicifolia binds with high affinity to

terminal α-galactosyl residues of glycoproteins on the cell surface of brain, cardiac muscle and

skeletal muscle endothelial cells (Laitinen, 1987; J. E. Lee et al., 2008). The antigen-Lectin

complexes were labeled utilizing the Vectastain ABC reagent kit (Vector Labs) and

immunoperoxidase staining was detected with 3, 3’-diaminobenzidine (DAB, Vector Labs).

Sections were counterstained with Mayer’s Hematoxylin (Life Technologies) to detect nuclei,

dehydrated, and coverslipped with Permount mounting medium (Fisher Scientific). Vessels with

IB4-positive staining (small brown circular structures) and an outer diameter of 7-10 μm were

classified as capillaries. Images were captured under 400X magnification from each section (2

83

sections/limb, 40 fields/section). Vessel densities were determined as the total number of vessels

per myofiber number in each field.

Alpha-smooth muscle actin staining

Five μm thick transverse paraffin-embedded sections were deparaffinized and rehydrated with xylenes and alcohol/water (100%, 95% and 70%) solutions, respectively. Endogenous peroxidase activity was blocked with H2O2 followed by treatment with 5% normal goat serum at

room temperature for 1 hour to reduce nonspecific binding. Serial sections were then incubated

overnight at 4oC with rabbit polyclonal antibody against α-smooth muscle actin (α-SMA 1:15000,

Abcam) to detect vascular smooth muscle cells. A biotinylated goat anti-rabbit secondary antibody

(1:500, Vector Labs) was then applied to the sections for 2 hours at room temperature. The antigen- antibody complexes were labeled utilizing the Vectastain ABC reagent kit (Vector Labs) and immunoperoxidase staining was detected with 3, 3’-diaminobenzidine (DAB, Vector Labs).

Sections were counterstained with Mayer’s Hematoxylin (Life Technologies) to detect nuclei, dehydrated, and coverslipped with Permount mounting medium (Fisher Scientific). α-SMA positive vessels with a 20 μm or more outer diameter were identified as arterioles and venules.

Venules were distinguished from arterioles by their collapsed luminal structure and thinner layers of smooth muscle. Images were captured under 200X magnification from each section (2 sections/limb, 20 fields/section). Vessel densities were determined as the total number of vessels per myofiber number in each field.

84

Myeloperoxidase (MPO) staining

Five μm thick transverse paraffin-embedded sections were deparaffinized and rehydrated with xylenes and alcohol/water (100%, 95% and 70%) solutions, respectively. Endogenous peroxidase activity was blocked with H2O2 followed by treatment with 5% normal rabbit serum

to reduce nonspecific binding. Serial sections were then incubated overnight at 4oC with goat

polyclonal antibody against MPO (1:5000, R&D Systems) to detect neutrophils. A biotinylated

rabbit anti-goat secondary antibody (1:500, Vector Labs) was then applied to the sections for 2

hours at room temperature. The antigen-antibody complexes were labeled utilizing the Vectastain

ABC reagent kit (Vector Labs) and immunoperoxidase staining was detected with 3, 3’- diaminobenzidine (DAB, Vector Labs). Sections were counterstained with Mayer’s Hematoxylin

(Life Technologies) to detect nuclei, dehydrated, and coverslipped with Permount mounting medium (Fisher Scientific). Images were captured under 400X magnification from each section (2 sections/limb, 40 fields/section). The number of MPO positive cells were counted in each images and presented as the total number of cells per field.

Mac-3 staining

Five μm thick transverse paraffin-embedded sections were deparaffinized and rehydrated with xylenes and alcohol/water (100%, 95% and 70%) solutions, respectively. Endogenous peroxidase activity was blocked with H2O2 followed by treatment with 5% normal goat serum to reduce nonspecific binding. Serial sections were then incubated overnight at 4oC with rat-anti

mouse monoclonal antibody against Mac-3 (CD107b 1:1000, BD Biosciences) to detect macrophages. A biotinylated mouse absorbed rabbit anti-rat secondary antibody (1:500, Vector

Labs) was then applied to the sections for 2 hours at room temperature. The antigen-antibody

85

complexes were labeled utilizing the Vectastain ABC reagent kit (Vector Labs) and

immunoperoxidase staining was detected with 3, 3’-diaminobenzidine (DAB, Vector Labs).

Sections were counterstained with Mayer’s Hematoxylin (Life Technologies) to detect nuclei,

dehydrated, and coverslipped with Permount mounting medium (Fisher Scientific). Images were

captured under 400X magnification from each section (2 sections/limb, 40 fields/section). The

number of Mac-3 positive cells were counted in each images and presented as the total number of

cells per field.

Micro-Computed Tomography Imaging of Vasculature

Micro-CT (Perfusion of contrast agent)

During preliminary perfusion experiments, three different contrast agents with high radio

density were tested for their ability to provide the X-ray attenuation necessary for micro-CT

imaging of vascular structures: barium sulfate, lead oxide and lead chromate. To prepare the

barium sulfate or lead oxide for perfusion, 105% w/v barium sulfate (E-Z-EM Canada Inc) or

100% w/v lead oxide (Sigma Aldrich) and 5% porcine gelatin (Sigma Aldrich) were added to

saline and heated with stirring to 60oC until the solution was completely dissolved. The solution was cooled to and maintained at 37oC in a water bath until a 1 mL syringe was filled and injected

at 1 mL/h. Samples were immediately stored at 4oC overnight to induce polymerization of the

gelatin suspensions within the vasculature until they were ready for CT scanning. For Microfil

perfusion, a curing agent (5% v/v) was added to catalyze the components (compound and diluent

mixed in equal amounts, according to the manufacturer recommendations; Flow-Tech Inc) to give

a working time of 20 minutes before the start of polymerization. 0.5 mL of the prepared compound

was drawn in a 1 mL syringe and injected at 1 mL/h with care taken to remove any perfusate that

86

contaminated the surrounding tissue, skin or fur. Perfused specimens were allowed to sit at room

temperature for 30 minutes and then stored at 4oC overnight to allow for contrast agent

polymerization.

For all perfusion experiments, a hand-pulled catheter (made from polyethylene [PE] tubing,

1/4" ID, 3/8" OD) was attached to a three-way tubing connector via 23-gauge needle adapters.

Standard PE-50 tubing were attached to the Y branches of the tubing connector; one side to deliver

the perfusates via a syringe pump and the other to connect to a pressure transducer and blood

pressure analyzer. The pressure across the tubing was monitored and maintained between 95 and

100 mmHg with adjustments made to the pump speed if pressure exceeded this range. Non-

ischemic (in preliminary experiments) and ischemic mice were anesthetized and overdosed with

sodium pentobarbital at a dose of 80 mg/kg, i.p. A horizontal midline incision was made right

below the diaphragm to avoid pneumothorax and to ensure blood flow during catheter insertion.

The abdominal aorta was cannulated, the catheter secured with 5-0 suture and a 6 mL heparinized

(100 U/mL) saline solution containing a mixture of vasodilators (10 µM acetylcholine, 10 µM

adenosine, 100 µM papaverine and 200 µM sodium nitroprusside; Sigma-Aldrich) was infused at

1 mL/min with a 3 mL plastic syringe (Becton Dickson #309659) for blood washout. The inferior

vena cava was severed to allow for drainage of blood and the infusing solution.

Micro-CT image acquisition

Non-ischemic, sham and ischemic limbs were dissected at the head of the femur for individual scanning at high resolution using the micro-CT scanner in the Tri-modal Siemens

Inveon (Siemens Healthcare) and the Inveon Acquisition Workplace (IAW) (version 1.5.28) software. Each limb was placed in a field of view of 18 x 25 mm, positioned on the dorsal side,

87

and scanned from the thigh to the ankle. CT scanning parameters were 80 kVp and 300 μA (voltage

and current) with 0.5 mm Al filtration. A 192o half scan was performed for 384 steps with an exposure time of 2100 ms for each step. Center offset and dark/light calibrations were performed immediately prior to scanning. The acquisition protocol included using high system magnification with 2x2 binning to generate an isotropic voxel size of 17 μm3. For the raw data reconstruction,

beam hardening was applied and was accurately Hounsfield Unit (HU) calibrated. Hounsfield units

are values assigned to voxels (3-D pixels) within a CT image that reflect the range of electron

densities and is proportional to the degree of X-ray attenuation (Kline & Ritman, 2012). Siemens

Inveon Research Workplace, (IRW version 4.0) and the 3-D Visualization tool were used to

generate 2-D cross-sections in yz-(sagittal), xy-(coronal), and xz-(axial) planes and 3-D volume

rendered images (Figures 5-7).

Comparison of contrast agents

Limbs perfused with either barium or lead contrast agents were not suitable for vascular

visualization or analysis. In the case of barium sulfate perfused-limbs, bone was very clearly

visualized and distinguishable from tissue but no vessels could be seen (Figure 5). Deposits of the

contrast agent were visible near the top of the limbs (i.e. adjacent to the site of perfusion)

suggesting that perfusion of the vessels was not successful. The lead oxide-perfused limbs produced images showing clear distinction of the bone and vessels (perfused with contrast agent) from muscle. The range of the HU values of the lead oxide-containing perfused vessels exceeded the upper limits of the viewing software and resulted in oversaturated voxels that could be not be analyzed (Figure 6). The Microfil (lead chromate with silicone rubber) contrast agent also produced images with clear distinctions of perfused vessels from bone and tissue (Figure 7). These

88 images had HU values within the range of the viewing software and were suitable for vascular visualization and quantification.

89

A AXIAL

C D

SAGITTAL 3D MIP

B CORONAL

Figure 5: Representative 2-D cross-sectional axial (A), coronal (B) and sagittal (C) slice views of a non-ischemic WT limb perfused with a barium sulfate and gelatin suspension which did not provide vascular contrast due possibly to improper perfusion or vasodilation. (D) A high resolution 3-D maximum intensity projections (MIP) of the same limb where the contrast agent was visible at the top of the image but did not perfuse the vessels (red arrows).

90

A AXIAL

D C 3D MIP SAGITTAL

B

CORONAL

Figure 6: Representative 2-D cross-sectional axial (A), coronal (B) and sagittal (C) slice views of a non-ischemic WT limb perfused with a lead oxide and gelatin suspension which perfused the vessels and provided contrast. (D) A high resolution 3-D maximum intensity projections (MIP) of the same limb. The high electron density of the contrast agent resulted in images that could not be analyzed due to oversaturation (red arrows) of the vascular structures.

91

A AXIAL

C D SAGITTAL 3D MIP

B CORONAL

Figure 7: Representative 2-D cross-sectional axial (A), coronal (B) and sagittal (C) slice views of a non-ischemic WT limb perfused with a lead chromate silicone rubber suspension (Microfil) which perfused vessels and provided contrast. (D) A high resolution 3-D maximum intensity projections (MIP) of the same limb. This contrast agent produced images with adequate contrast of vessels from tissue and 3-D volumes suitable for visualization and quantitative analysis of vascular structures. 92

Micro-CT image analysis

For 3-D vessel morphology analysis, bones were digitally removed and the vasculature was

segmented from the surrounding tissue (background) based on the density of the X-ray attenuated-

voxels of the Microfil. Briefly, manual segmentation or automatic segmentation was performed to designate regions of the scanned volumes as bone, vessel or non-vascular tissue (Figure 8). The manual segmentation entailed manually drawing volumes of interest (VOI) that surrounded the bone regions in each of the serial 2-D sections (~500 slices in the coronal plane) that make up the

3-D volumes. This was followed by automated segmentation based on the density of the voxels in the Microfil-filled vessels (thresholding). A threshold range of 1000 -7000 HU was chosen and

used to evaluate all the scanned volumes. This range was selected based on comparing the ability

of different ranges of HU values to effectively separate tissue and vessels into distinct volumes

(binarization) in post-segmentation images.

Histomorphometric parameters including vessel density (number of vascular

structures/mm), volume (volume of segmented voxels normalized to limb volume), diameter (local

thickness of voxels), and spacing (average distance/separation between vessels quantified from

background voxels) were calculated for each defined volume of interest (sham or ischemic calf or

thigh) using the Siemens IRW Bone Morphometry Tool. The values were reported as a percent

ratio of ischemic to non-ischemic limbs for each animal to minimize any variations in the vascular

morphology parameters that could be attributed to the contrast agent perfusion.

To determine and compare the vessel thickness (diameter) distribution in the mouse limbs,

bone-subtracted 2-D slices were analyzed for local thickness in Fiji (NIH). For each 3-D data set,

an in-house Matlab (Mathworks; Natick, MA) subtraction program was applied to generate images where all bone had been removed. These stacked data were converted to DICOM files and

93

evaluated for “local thickness” in Fiji. The local thickness tool defines a volume-based local

thickness at every voxel in the limb volume as the diameter of the largest sphere that both contains

the voxel and is entirely within the volume of structure (Zagorchev et al., 2010). Histograms were

generated to display the volume and range of vessel sizes diameter (34-400 μm) across the sham and ischemic limbs.

Figure 8: Representative 2-D cross-sectional coronal (A), sagittal (B), and axial (C) slice views of a non-ischemic limb WT after manual (to subtract bones) and automatic (HU-based

94

thresholding of vessels) segmentation. The calf region is highlighted in red with bones

segments digitally deleted (white open arrows) and vessels segmented based on HU voxel

intensity (white arrowheads).

Laser Doppler Perfusion Imaging

Perfusion was measured using images acquired with an infra-red laser Doppler perfusion imager (Moor LDI2, Moor Instruments). Anesthetized animals were placed ventrally with legs outstretched on a heating pad at 37oC and positioned to allow for scanning of the both lower

hindlimbs (sham and ischemic). Scanning of the lower hindlimbs (paws) was performed before

surgery (baseline) and on days 1, 14 and 42 after induction of chronic ischemia. Three consecutive

color-coded images representing the flow distribution were obtained per animal

for the calculation of perfusion values. In each image, high perfusion is signified by the red colors,

yellow and green depict the mid-range and the blues indicate low perfusion. Relative perfusion

was expressed as the ratio of the mean perfusion values of the ischemic to sham (non-ischemic)

feet.

Statistical Analysis

Data are expressed as mean ± S.E.M. Limb use/function data were compared with two-

way repeated measures analysis of variance (ANOVA). Necrosis data are presented as percentages

and were analyzed using Chi-Square test for categorical data. All other analyses were performed

using one-way or two-way ANOVAs. When ANOVA evaluations produced significance (p≤

0.05), Bonferroni post-hoc tests were carried out for multiple or pair wise comparisons.

95

Chapter 1. HMW and LMW isoforms of FGF2 in ischemia-induced revascularization and

tissue recovery

RESULTS I

The roles of the HMW and LMW FGF2 protein isoforms in embryonic vascular

development of the heart have been previously studied (Amann et al., 2006; Liao et al., 2009).

Liao and colleagues analyzed non-ischemic hearts from adult FGF2 HMWKO (LMW expression

only) and FGF2 LMWKO (HMW expression only) mice for capillary and smooth muscle- containing vessels (arterioles). Those experiments produced no differences from WT in the number of cardiac capillaries and arterioles in either knockout group. The purpose of this dissertation was to evaluate the roles of the FGF2 isoforms in ischemia-induced vascular growth and remodeling.

To that end, WT (all FGF2 isoforms expressed), Fgf2-/- (no LMW or HMW FGF2 expression)

FGF2 HMWKO (no HMW FGF2 expression), and FGF2 LMWKO (no LMW FGF2 expression), mice were subjected to a surgical model of hindlimb ischemia (Couffinhal et al., 1998; Murohara et al., 1998; Sullivan et al., 2002). This is a well-characterized model where a section of the superficial femoral artery and vein pair is ligated and then excised (Figure 4). A majority of the mice survived the surgery and all deaths that occurred were related to anesthesia and not the procedure. The mice were allowed to recover and monitored daily for any health or procedure-

related issues. For enhanced clarity of the following figures and results, the FGF2 HMWKO and

FGF2 LMWKO mouse groups will, henceforth, be referred to as FGF2 LMW-only and FGF2

HMW-only, respectively.

Accelerated recovery of ischemic limb use in the presence of HMW FGF2

Evaluation of ischemic hindlimb use was performed at specific time points after induction

of ischemia to address the roles of the FGF2 isoforms in recovery of the ischemic limb. A semi-

96

quantitative scoring scale modified from Stabile and colleagues (Helisch et al., 2006; Stabile et al.,

2003) was utilized to perform daily and weekly assessments of ischemic limb impairment or use

(Table 12). Mice from all groups experienced impaired limb use immediately after the ischemia surgery which presented as dragging of the ischemic foot and this observation persisted for the next three days of ischemia (Figure 9), most likely due to the ischemic pain and/or surgical procedure. Soon after, all groups of mice began to exhibit improved use of their ischemic limbs to varying degrees. By 28 days, WT had resumed normal limb function and their ischemic limbs could no longer be differentiated from sham limbs. Mice expressing only LMW FGF2 had a more prolonged recovery time and ischemic limb use did not return to normal until day 42. Conversely, the FGF2 HMW-only limbs had a faster recovery (2-3x) of the ischemic limb with normal use regained at 14 days compared to the other groups (p<0.05, Figure 9). Mice with no expression of

FGF2 (Fgf2-/-) did not completely recover ischemic limb use by the end of the study at 42 days.

Score Hindlimb Use

4 Dragging of ischemic foot

No dragging of ischemic foot but no plantar 3 flexion

2 Plantar flexion but no flexing of toes

Normal use of ischemic limb with flexing of 1 toes to resist gentle traction on the tail

Table 12: Descriptive scoring for the assessment of hindlimb use during chronic ischemia.

97

WT 5 Fgf2 -/-

4 FGF2 LMW-only FGF2 HMW-only

3

2 *†‡ Limb Impairment Score Impairment Limb 1 *†‡ *†‡ †‡ †‡

0 0 3 7 10 14 21 28 35 42

Days post-surgery

Figure 9: Average limb impairment scores of WT (n=35), Fgf2-/- (n=26), FGF2 LMW-only

(n=31) and FGF2 HMW-only (n=25) ischemic limbs after induction of hindlimb ischemia.

A descriptive scoring system listed in Table 12 was used to assign scores daily for up to 14 days of ischemia and then weekly for up to 42 days of ischemia. Data are presented as mean

± SEM. *p<0.05 vs. WT, †p<0.05 vs. Fgf2-/-, ‡p<0.05 vs. FGF2 LMW-only.

98

Loss of FGF2 expression negatively influences limb functional capacity

As a physiological alternative to the hindlimb use assessment, the effect of FGF2 isoform

expression on functional recovery of the ischemic was also determined by measuring the forced

exercise capacity of the mice in this study before and 14 days after the ischemia surgery was

performed. At 1-3 days prior to the surgery (baseline test), a cohort of WT, Fgf2-/-, FGF2 LMW-

only and FGF2 HMW-only mice were subjected to exercise treadmill testing at multiple belt

speeds and the same group was again tested 14 days later. The speed was gradually increased over

time and mice were allowed to run for 5 minutes at the 7, 10 and 15 m/min speeds. When a speed

of 20 m/min was reached, the mice were allowed to run for either 15 minutes or length of time

before which the animal was unable to keep the pace with the motorized treadmill belt; as indicated

by a continuous break in both of the infrared beams located at the end of each treadmill belt (Fewell

et al., 1997; Q. Yang et al., 2001). For the entire length of the test, the number of breaks in the dual

infrared beam by a part of the animal (usually tail and/or hindlimbs) was detected and recorded

(Fewell et al., 1997; Lerman et al., 2002).

In absence of ischemia (baseline), no differences in the levels of exercise performance was detected between WT and Fgf2-/- mice at lower treadmill speeds (7 or 10 m/min) (Figure 10A).

When the speed was increased to 15 m/min, the Fgf2-/- mice performed poorer than WT as

evidenced by the increased number of beam breaks/minute. When the treadmill speed was further

increased to 20 m/min, the performance of Fgf2-/- mice was unchanged from WT levels. After two weeks of ischemia, the performance of Fgf2-/- mice was unchanged from WTs at 7, 10 or 15 m/min. However, exercise activity appears to be decreased when the maximum speed of 20 m/min was reached. For either baseline or post-ischemic exercise performance of FGF2 LMW-only or

FGF2 HMW-only mice, levels of exercise capacity similar to WT mice were observed at all speeds

99

of the treadmill testing protocol. These results provide preliminary evidence suggesting that in the

absence of Fgf2 expression, there is an impaired functionality of the musculoskeletal system occurring in the absence or presence of ischemic injury.

A. Performance at baseline 50 WT Fgf2-/- 40 FGF2 LMW-only FGF2 HMW-only 30

20 Beam breaks/min Beam 10

0 7 10 15 20 Speed (meters/min)

100

B. Performance after 14 days of ischemia 50 WT Fgf2-/- 40 FGF2 LMW-only FGF2 HMW-only 30

20 Beam breaks/min Beam 10

0 7 10 15 20

Speed (meters/min)

Figure 10. Treadmill exercise performance of WT (n=1), Fgf2-/- (n=2), FGF2 LMW-only

(n=2) and FGF2 HMW-only (n=2) ischemic limbs before (A) and 14 days after (B) induction of hindlimb ischemia. Performance data are presented as mean beam breaks ± SEM.

101

HMW FGF2 promotes while LMW FGF2 abrogates gross post-ischemic tissue viability

In conjunction with recovery of ischemic limb function evaluation, daily examinations

were performed to assess the FGF2 isoform-induced gross tissue viability of the ischemic limb.

None of the mice had any necrosis of the sham-operated limb. Of the 124 male and female mice included in the study, ~16% (20/124) exhibited signs of tissue necrosis in the ischemic limb.

Necrosis was defined as the darkening of one or more toes or part of the footpad within 1-3 days of surgery but did not include limb loss (auto-amputation). Only 4 of the 124 mice with necrosis experienced auto-amputation and these mice were immediately euthanized and excluded from the rest of the study. In the mice with necrotic limbs, the tissue necrosis was not progressive and they were healed within 2 weeks of the surgical induction. WT mice had a 22.5% (7/31) incidence of necrosis (Figure 11). Fgf2-/- and FGF2 LMW-only mouse groups had increased incidences that

were not statistically different, 21.9% (7/32) and 23.16% (6/26), respectively. Conversely, mice

with expression of only the HMW FGF2 had no signs of tissue necrosis (0/26, Figure 11, p<0.05)

compared to the other groups. Interestingly, the mice that had to be excluded from the study due

to limb auto-amputation were from the WT or FGF2 LMW-only groups (2 of each). This fact

coupled with the observations that limbs expressing only HMW FGF2 had no necrosis and those

expressing only LMW FGF2 were more susceptible to necrosis suggests a protective role for

HMW FGF2 in tissue viability and a deleterious function for LMW FGF2.

102

100

80

60

40 % ANIMALS

20

0 WT Fgf2-/- FGF2 LMW- FGF2 HMW- only only NECROSIS NO NECROSIS

B. MALE MICE * † ‡ 100

80

60

40 % ANIMALS

20

0 WT Fgf2-/- FGF2 LMW- FGF2 HMW- only only NECROSIS NO NECROSIS

103

FEMALEAll mice MICE C. 140 * † ‡ 100 WT 120 Fgf2 -/- 80 100 FGF2 LMW-only FGF2 HMW-only 60 80

60

40% Animals

% ANIMALS 40

20 20

0 0 NECROSIS NO NECROSIS WT Fgf2-/- FGF2 LMW- FGF2 HMW- only only NECROSIS NO NECROSIS

Figure 11. (A) Incidence of limb necrosis (defined as discoloration of toes and/or foot observed daily for up to 14 days of ischemia compared to the sham-operated) in WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only ischemic limbs. (B) Incidence of limb necrosis in male ischemic limbs. (C) Incidence of limb necrosis in female WT (n=14), Fgf2-/-

(n=13), FGF2 LMW-only (n=18) and FGF2 HMW-only (n=10) ischemic limbs. For each sex/genotype, the presence or absence of necrosis is expressed as a percentage of the number of animals in which hindlimb ischemia surgery was performed. Data is presented as percentages of all animals in the study. *p<0.05 vs. WT, †p<0.05 vs. Fgf2-/-, ‡p<0.05 vs. FGF2

LMW-only.

104

Expression of only HMW FGF2 enhanced post-ischemic vascular network

To identify the role of the individual FGF2 isoforms in ischemia-related vascular and collateral growth and remodeling, both contrast-enhanced micro-CT imaging and immunohistochemistry were performed. The next set of figures (Figure 12-15) depict the outcomes from the micro-CT experiments. This imaging technique is relies on the radiopacity of material and resultant X-ray attenuation to generate sequential tomographic images (slices) which are then reconstructed to produce three dimensional (3-D) volumes of the scanned object (Bouxsein et al.,

2010a; Downey et al., 2012). The low inherent contrast of vessels and muscle (due to low atomic numbers of carbon and oxygen) requires the use of a radiodense contrast medium to increase image contrast (A. S. Lin et al., 2007; Zagorchev et al., 2010). In this case, a lead-based material, Microfil was perfused into the hindlimbs of WT, Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only mice to visualize and quantitatively analyze the vascular network at 42 days of ischemia (Figure 12).

The 3-D images of the sham and ischemic limbs reveal the presence of perfused superficial femoral arteries in the former limbs and the absence in the latter. After digitally removing the bone structures and automatic segmentation of tissue from vessels in the 3-D volumes, morphometric analysis was performed to determine the volume, number and spacing of the vessels in the calf or thigh regions. The delineation of the calf from the thigh allows for the distinction of angiogenesis from arteriogenesis in ischemic vascular repair (Limbourg et al., 2009). Femoral artery ligation and excision has been shown to predominantly induce angiogenesis in the calf muscles of the lower hindlimb and collateralization in the upper thigh regions (Oses et al., 2009; G. Tang, Charo, Wang,

Charo, & Messina, 2004; Tirziu et al., 2005). To minimize the possibility of vascular over- or

under-estimation resulting from variations in contrast-agent perfusion, the vascular morphology

parameters of each ischemic limb was reported as the percent change from contralateral, non-

105

ischemic (sham) limb. The increase in vessel volume (volume of segmented vessels normalized to

calf volume) in the ischemic calf of WT mice was similar to Fgf2-/- limbs (47% vs. 47%) (Figure

13A). Expression of only LMW FGF2 produced a significantly smaller change in vascular volume

(18%, p<0.05 vs WT). The FGF2 HMW-only ischemic calves had the greatest increase in vessel

volume (69%, p<0.05 vs WT). The change in vessel density (number of vessels/mm of tissue) and

intervessel spacing (distance between vessels) were also determined (Figure 13B-C). Consistent

with the vascular volume results, the ischemic limbs, expressing only LMW FGF2, had the least

change in vessel density and was significantly different from WT or HMW-only limbs (p<0.05).

A decrease in intervessel spacing in the ischemic limbs relative to the sham limb is indicative of

an increase in the portion of the tissue occupied by vessels. This factor was only decreased by 10%

in FGF2 LMW-only calves while WT and FGF2 HMW-only had significantly “higher” (i.e. more negative) changes in vessel separation with 40.68% and 40.73%, respectively (p<0.05).

Simultaneous measures of vessel volume, density and spacing were also obtained in the thigh regions of the sham and ischemic limb volumes (Figure 13). The thigh vessel volume followed the same pattern as was observed in the ischemic calf where the increase in ischemic vascular volume of the Fgf2-/- was similar to WT limbs but the increase in the FGF2 HMW-only

muscles was higher than WT. The thigh vessel numbers in the HMW FGF2 expressing limbs were

also not different from WT. Surprisingly, the vascular density in the ischemic Fgf2-/- thighs was

decreased compared to WT despite similar vessel volumes. The FGF2 LMW-only thighs had reduced vessel densities relative to WT which would be expected given the lower vessel volumes that was also observed. The percent change in the thigh intervessel spacing of the WT, Fgf2-/- and

FGF2 HMW-only limbs were not different from each other but they were significantly greater than

the spacing of the FGF2 LMW-only thighs (p<0.05).

106

Taken together, these micro-CT data demonstrate that the HMW isoforms of FGF2 possess a greater capacity for promoting collateral arteriogenesis in both ischemic calf and thigh muscles.

107

FGF2 FGF2 WT Fgf2-/- LMW-only HMW-only

SHAM

HLI

Figure 12. Representative micro-computed tomography (micro-CT) images of sham (top row) and ischemic (bottom row) hindlimbs

(HLI) from WT, Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only mice at 42 days of ischemia. Perfused femoral arteries (red arrows)

are present only in the sham limbs.

108

A. 100 WT Fgf2 -/- FGF2 LMW-only 80 FGF2 HMW-only *

60 * † *

40 Vessel Volume Vessel

% % ChangeSham from * † 20

0 CALF THIGH

B. 100

80

* 60 * * †

40 Vessel Density Vessel

% Change from Sham from Change% * † 20

0 CALF THIGH

109

C. WT -100 Fgf2 -/- FGF2 LMW-only -80 FGF2 HMW-only

-60

*† -40 Vessel Spacing Vessel % Change from Sham from Change% -20 *† *

0 CALF THIGH

Figure 13. Vessel volume (normalized for tissue volume) (A), vessel density (B) and vessel spacing (C) was calculated from 3-D micro-CT images of WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars) limbs. Volume, density and spacing of the ischemic limb vasculature in the calf or thigh are expressed as change (%) from sham limb. n=4 to 6. *p< 0.05 vs. WT, †p<0.05 vs. FGF2 HMW-only.

110

Expression of only HMW FGF2 increases the number of small and medium-sized collaterals in the post-ischemic vascular network

In order to provide additional characterization of the HMW or LMW FGF2-induced vascular network, the range of vessel diameters present in sham and ischemic hindlimbs was determined. 3-D limb volumes generated by micro-CT were analyzed and quantified for the distribution of vessels thicknesses (diameter). Histograms of vessel sizes present in the mouse skeletal muscle showed vessels that ranged from 34µm to 408µm. 34µm was the smallest vessel able to be visualized and quantified (i.e. the highest resolution possible for this study) and 408 µm was the upper limit above which the vascular data proved to be unreliable for analysis.

A similar distribution of vessels was observed in the (sham) non-ischemic limbs of

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice relative to WT (Figure 14). Also, the volume of the vessels at each thickness category (vessel frequency) was not different from WT. In

WT and FGF2 HMW-only ischemic limbs, the vessels on the lower half of the thickness range

(34-200µm, smaller collaterals) were increased in volume compared to their respective sham limbs

(Figure 15A, p<0.05). Additionally, the number of 34-200μm-sized collaterals was significantly higher in the FGF2 HMW-only group than WT (p<0.05). Conversely, the 34-200µm vessel volume of FGF2 LMW-only ischemic limbs was unchanged from the sham limb and significantly lower than WT or FGF2 HMW-only volumes (p<0.05).

Volumes of medium-sized (200-400μm) collateral vessels in WT and FGF2 HMW-only

ischemic limbs were also increased compared to the contralateral sham limb (Figure 15B, p<0.05).

Expression of only HMW FGF2 resulted in significantly higher volumes of 200-400μm vessels

relative to WT (p<0.05), while the presence of only the LMW isoform did not change the post-

111

ischemic development of the medium-sized collaterals. As was observed with the smaller ≤200µm

vessels, medium sized vessels in FGF2 LMW-only muscles were significantly decreased

compared to WT and HMW FGF2-only muscles (p<0.05). Overall, these results suggest that the

development of small (34-200µm) and medium (200-400µm, to a lesser degree) sized vessels are

responsible for the increased vascular volume present in ischemic HMW FGF2-expressing muscle.

A. 30 Vessel Diameter (µm)

# WT Sham WT HLI # # # 20 voxels) 4 4

10 Vessel Volume (x10 Volume Vessel

0 34 68 102 136 170 204 238 272 306 340 374 408

112

B.

30 Fgf2 -/- Sham

Fgf2 -/- HLI

voxels) 20 4 # #

10 Vessel Volume (x10 Volume Vessel

0 34 68 102 136 170 204 238 272 306 340 374 408 Vessel Diameter (µm)

C. 30 FGF2 LMW-only Sham

FGF2 LMW-only HLI

20 voxels) 4

10 Vessel Volume (x10 Volume Vessel

0 34 68 102 136 170 204 238 272 306 340 374 408 Vessel Diameter (µm) 113

D. 40 FGF2 HMW-only Sham # # # # FGF2 HMW-only HLI 30 voxels) 4 20 #

10 Vessel Volume (x10 Volume Vessel 0 34 68 102 136 170 204 238 272 306 340 374 408 Vessel Diameter (µm)

Figure 14. Histogram of vessel thickness distribution in sham (white bars) and ischemic

(black bars) limbs of WT (A), Fgf2 -/- (B), FGF2 LMW-only (C), and FGF2 HMW-only (D)

mice analyzed from micro-CT images at 42 days of ischemia. Data are presented as mean ±

SEM, n=4 to 6. #p<0.05 vs. SHAM cohort.

114

200 34-200μm diameter vessels A. WT 160 Fgf2 -/- FGF2 LMW only # * FGF2 HMW only voxels)

4 120 # † 80

40 Vessel Volume (x10 Volume Vessel

0 SHAM HLI B.

100 200 - 400µm diameter vessels

WT # * 80 Fgf2 -/- FGF2 LMW only #

voxels) FGF2 HMW only 4 60

* † 40

Vessel Volume (x10 Volume Vessel 20

0 SHAM HLI

115

Figure 15. The volume of small-sized (34 -200μm) vessels (A) and medium-sized (200-400μm) vessels (B) analyzed from micro-CT images of WT (white bars), Fgf2-/- (light gray bars),

FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars) limbs. Data are presented as mean ± SEM, n=4 to 6. #p<0.05 vs. SHAM cohort, *p< 0.05 vs. WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

116

Loss of LMW FGF2 perturbs while HMW FGF2 augments capillary growth after 14 days

of ischemia

To elucidate the effect of altered FGF2 isoform expression on the initial vascular response to ischemic injury, skeletal muscles from sham-operated or ischemic WT, Fgf2-/-, FGF2 LMW- only and FGF2 HMW-only mice were harvested at 14 days of ischemia. Immunohistological analysis was performed to determine the level of post-ischemic microvessel growth. Angiogenesis, referring solely to capillary vessel development, was measured by quantitative analysis of muscle cross-sections stained with Griffionia Simplicifolia Isolectin B4 (GSI Lectin). GSI Lectin-positive

endothelial cells (ECs) were present on the inner layer (endothelium) of all blood vessels (Figure

17). Capillary vessels were identified as vessels with an outer diameter ≤10 µm and positive

staining for ECs. Capillary vessel densities were determined as the ratio of the total number of

capillary vessels to myofiber numbers. This was done to address the possibility of under or

overestimation of vessel numbers due to ischemia-related or atrophy.

Capillary density was measured in muscle regions of the mouse hindlimb that are proximal

(calf) or distal (thigh) to the excised portions of the femoral artery and vein pair (Figure 16). After

14 days of ischemia, the capillary vessel density ratio (ischemic/sham) of WT, Fgf2-/- and FGF2

HMW-only limbs were greater than 1 indicating the presence of increased vessel numbers in the

ischemic muscles. FGF2 LMW-only ratios were scarcely above 1 at (1.07 ± 0.05) and (1.01 ±

0.06) for calf and thigh muscles respectively, and were decreased compared to WTs. Fgf2-/-

capillary density ratios were similar to WT ratios in either the calf or thigh. In FGF2 HMW-only

limbs, the ratio of capillary vessels of the thigh muscles was significantly greater than WT while

the calf muscle ratio had a trend towards an increase. Additionally, FGF2 HMW-only capillary

ratios were elevated relative to the FGF2 LMW-only ratios in both calf and thigh muscles. These

117 results suggest that modulation of FGF2 isoform expression alters the vascular response to ischemia. Furthermore, the data show the opposing roles for HMW FGF2 (stimulation) and LMW

FGF2 (suppression) in angiogenesis (capillary growth).

Capillary Density WT 3 Fgf2-/- FGF2 LMW-only 2.5 FGF2 HMW-only

2 *

1.5 * † * † 1

0.5 Vessel Density Ratio (Ischemic/Sham) Ratio Density Vessel

0 CALF THIGH

Figure 16. Capillary density (capillary number/myofiber number) in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars) calf or thigh expressed as a ratio of ischemic to sham muscle vessel density after 14 days of ischemia. Data are presented as mean ratio ± SEM, n= 4 to 5. *p< 0.05 vs. WT,

†p<0.05 vs. FGF2 HMW-only.

118

LMW FGF2 isoform did not increase post-ischemic vascular (capillary) growth while HMW

FGF2 improved capillary vascularization after 42 days of ischemia

To further elucidate the effect of altered FGF2 isoform expression on the vascular response to ischemic injury, skeletal muscles from sham-operated or ischemic WT, Fgf2-/-, FGF2 LMW- only and FGF2 HMW-only mice were also harvested at 42 days of ischemia. No significant differences in capillary density were observed in non-ischemic (sham) muscles between WT and

Fgf2-/- hindlimbs (Figure 18). Limbs with FGF2 HMW-only expression were also similar to WTs while the FGF2 LMW-only limbs had decreased capillaries in the sham muscles. Capillary density in the WT, Fgf2-/-, and FGF2 HMW-only ischemic muscles were significantly increased relative to their respective sham limbs (p<0.05). FGF2 LMW-only ischemic limbs did not have any change capillary density compared to their non-ischemic, sham limbs. Moreover, the proportional increase in capillary density was highest in the HMW FGF2 expressing muscles and the density of vessels in the ischemic FGF2 HMW-only limbs was greater than that of the WT, Fgf2-/-, or FGF2 LMW- only ischemic limbs (p<0.05).

To determine the contributions of the specific calf and thigh muscle groups to the overall changes in ischemic vascular density, capillary density was evaluated in the sham and ischemic gracilis, semi-membranosus, semi-tendinosus and gastrocnemius. The latter muscle makes up the calf muscle and the former three groups are found in the thigh. Increased vascular density in the gracilis and gastrocnemius muscles had the greatest contribution to the capillary growth (Figure

19 A, D). In WT, Fgf2-/- and FGF2 HMW-only gracilis and gastrocnemius muscles, capillary density was significantly higher than the contralateral sham muscles (p<0.05). FGF2 LMW-only ischemic muscles had no change in the vascular density relative to the sham limbs in the gracilis, semi-membranosus, semi-tendinosus or gastrocnemius muscles (Figure 19 A-D). Furthermore, the

119

capillary densities in WT or FGF2 HMW-only ischemic muscles were consistently greater than in

FGF2 LMW-only muscles (p<0.05). Overall these data, for the first time, indicate that HMW

FGF2 isoform expression is important for capillary growth after ischemia and supports the

anatomical vascular density results obtained with micro-CT.

120

WT Fgf2-/- FGF2 LMW-only FGF2 HMW-only

SHAM

HLI

Figure 17. Representative photomicrographs of GSI-Lectin stained vessels (all vessels) in WT, Fgf2-/-, FGF2 LMW-only, and

FGF2 HMW-only sham and ischemic muscle sections harvested at 42 days of hindlimb ischemia. Scale bar: 10μm, arrows

(capillaries), arrowheads (non-capillary GSI-Lectin+ vessels).

121

Capillary Density 1 WT # * Fgf2 -/- 0.8 FGF2 LMW-only FGF2 HMW-only # 0.6 #

*† 0.4 Vessel Myofiber# / # Vessel 0.2

0 SHAM HLI

Figure 18. Capillary density (expressed as capillary number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars),

FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs. WT HLI, †p<0.05 vs. FGF2

HMW-only HLI.

122

A. B. Semi-membranosus 1

0.8

0.6

0.4 Vessel # / Myofiber # Vessel 0.2

0 SHAM HLI

C. Semi-tendinosus D. Gastrocnemius 1 1 WT Fgf2 -/- WT FGF2 LMW-only 0.8 Fgf2 -/- 0.8 FGF2 HMW-only FGF2 LMW-only * FGF2 HMW-only * * 0.6 * 0.6 †

0.4 0.4

# # Vessel / Myofiber Vessel # Vessel / # Myofiber 0.2 0.2

0 0 SHAM HLI SHAM HLI

Figure 19. (A) Gracilis muscle capillary density (expressed as vessel number/myofiber

number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF HMW-only (black

123

bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs.

WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(B) Semi-membranosus muscle capillary density (expressed as vessel number/myofiber

number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black

bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs.

WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(C) Semi-tendinosus muscle capillary density (expressed as vessel number/myofiber number)

of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light

gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs. WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(D) Gastrocnemius muscle capillary density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs. WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

124

Loss of LMW FGF2 affects the early arteriogenic response to chronic ischemia

To address the function of selective FGF2 isoform expression in ischemia-induced vascular

growth, the level of arteriole (arteriogenesis) and venule growth were examined. To detect

arteriolar or venule vessels, α-smooth muscle actin (α-SMA) antibody was used to perform

immunohistological labeling of serial sections of sham-operated or ischemic WT, Fgf2-/-, FGF2

LMW-only and FGF2 HMW-only mice. α-SMA is a marker for pericytes and vascular smooth

muscle cells (vSMCs) present in the walls of venules, veins, arterioles and arteries (Skalli et al.,

1989). Arterioles/arteries were defined as α-SMA positive vessels with ≥20 µm outer diameter.

They had several layers of positive staining and a round lumen. Veins/venules were

morphologically distinguishable from arterioles/arteries by their collapsed luminal structure and

thinner layers of smooth muscle (Figure 21).

Arteriole density was determined as a ratio of the number of myofibers. Arteriole density

at 14 days of ischemia was evaluated in sham or ischemic muscle and with the ischemic density

expressed as a ratio of the later (Figure 20). The ratios of WT, Fgf2-/-, FGF2 LMW-only, FGF2

HMW-only limbs were higher than 1 indicating the presence of increased vessel numbers in the ischemic muscles. However, in the presence of only LMW FGF2, ratios were only slightly greater than 1 at (1.16 ± 0.06) and (1.04 ± 0.08) for calf and thigh muscles respectively. These ratios were decreased compared to WT or FGF2 HMW-only limbs. WT, Fgf2-/- and FGF2 HMW-only limbs all had similar arteriole density ratios (calf or thigh). These data suggest that unlike the HMW

FGF2 isoforms, LMW FGF2 does not stimulate collateral vessel remodeling (arteriogenesis) in response to ischemia.

125

Arteriole Density 3 WT Fgf2-/- 2.5 FGF2 LMW-only FGF2 HMW-only 2

1.5 * † * † 1

0.5 Vessel Density Ratio (Ischemic/Sham) Ratio Density Vessel

0 CALF THIGH

Figure 20. Arteriole density (arteriole number/myofiber number) in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars) calf or thigh expressed as a ratio of ischemic to sham muscle vessel density after 14 days of ischemia. Data are presented as mean ratio ± SEM, n= 4 to 5. *p< 0.05 vs. WT,

†p<0.05 vs. FGF2 HMW-only.

126

Expression of only HMW FGF2 has greater impact on ischemic arteriolization

(collateralization) 42 days after ischemia compared to WT or LMW isoform expression

Arteriole or venule density was also determined at 42 days and expressed as a ratio of the

number of myofibers (Figure 21). Venule or arteriole density was evaluated in sham or ischemic

muscles. No differences in venule or arteriole density were detected in non-ischemic (sham)

muscles between WT and Fgf2-/- hindlimbs (Figure 22A). Similarly, FGF2 LMW-only or FGF2

HMW-only sham muscles were not different from WT. After 42 days of ischemia, venule densities

in ischemic limbs of all groups were significantly increased compared to their respective sham

limbs (Figure 22A, p<0.05). To identify the changes in specific in calf (gracilis, semi-

membranosus, and semi-tendinosus) or thigh (gastrocnemius) muscles, the density of venule

vessels in the respective muscle layers was also determined. In the thigh, gracilis venule density

was increased above sham levels in all mouse groups except FGF2 LMW-only (Figure 23A,

p<0.05). The gastrocnemius muscle had elevated venule densities in all mouse groups (Figure 23D,

p<0.05). However, the venule densities in the ischemic gracilis, semi-membranosus, semi- tendinosus or gastrocnemius muscles were not different.

Arteriole density was significantly increased in ischemic muscles of WT, Fgf2-/-, FGF2

LMW-only or FGF2 HMW-only mice (Figure 22B, p<0.05). While expression of only LMW

FGF2 in ischemic hindlimbs induced a similar increase in arteriole density as the WT or Fgf2-/- limbs, expression of only the HMW isoforms promoted an even greater amount of revascularization compared to the three other groups (p<0.05). Ischemic FGF2 HMW-only arteriole density was elevated nearly five-fold relative to non-ischemic limbs but WT, Fgf2-/- and

FGF2 LMW-only muscles had only a two-fold increase. To assess the roles of individual calf and thigh skeletal muscles in the arteriole density changes observed during ischemia, vessel density in

127 gracilis, semi-membranosus, semi-tendinosus, or gastrocnemius muscle alone was calculated.

Unlike the angiogenesis results where the ischemia-induced capillary growth could be attributed to only the gracilis and gastrocnemius muscles, arteriogenesis was observed in all calf and thigh muscles analyzed (Figure 24A-D). Additionally, the expression of only the HMW FGF2 isoforms produced consistently higher levels of arterioles across all the muscle groups (p<0.05). Taken together, these results provide additional evidence that arteriogenesis is compromised in the LMW

FGF2-expressing limbs and confirms the increased of capacity of the HMW isoforms for inducing collateral growth.

In order to eliminate the likelihood that vessel density measurements were over- or under- estimated due to ischemia-related edema (tissue swelling) or atrophy (muscle shrinking), the size of the sham and ischemic muscles was determined. Muscles from the calf (gastrocnemius) or thigh

(semi-membranosus, semi-tendinosus, or gracilis) were weighed when harvested at the end of the study (Day 42 of ischemia). In general, the masses of the ischemic muscles were decreased relative to the sham limb but the differences were not significant. For all genotypes, the weights of the ischemic gracilis, semi-membranosus, semi-tendinosus, or gastrocnemius muscle was not different from the sham group (Figure 25).

128

WT Fgf2-/- FGF2 LMW-only FGF2 HMW-only

SHAM

HLI

Figure 21. Representative photomicrographs of α-smooth muscle actin (α-SMA) stained vessels (venules or arterioles) in WT,

Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only sham and ischemic muscle sections harvested at 42 days of hindlimb ischemia.

Scale bar: 10μm.

129

A. Venule Density 0.05 WT 0.04 Fgf2-/- FGF2 LMW-only FGF2 HMW-only 0.03

0.02 * * Vessel # / # Myofiber Vessel * * 0.01

0 SHAM HLI

B. Arteriole Density

0.05 WT Fgf2-/- *# 0.04 FGF2 LMW-only FGF2 HMW-only 0.03

*† * * 0.02

Vessel # / Myofiber # # Myofiber / Vessel 0.01

0 SHAM HLI

130

Figure 22. (A) Venule density (expressed as vessel number/myofiber number) of sham and

ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars),

FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort.

(B) Arteriole density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW- only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ±

SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs. WT HLI, †p<0.05 vs. FGF2 HMW- only HLI.

131

A. B. Gracilis Semi-membranosus 0.05 0.05 WT Fgf2-/- 0.04 0.04 FGF2 LMW-only FGF2 HMW-only 0.03 0.03 * 0.02 0.02 * * Vessel # / Myofiber # Vessel Vessel # / Myofiber # Vessel 0.01 0.01

0 0 SHAM HLI SHAM HLI

C. D. Semi-tendinosus Gastrocnemius 0.05 0.05

0.04 0.04

0.03 0.03 * 0.02 0.02 * * * Vessel # / Myofiber # Vessel

Vessel # / Myofiber # Vessel * 0.01 0.01

0 0 SHAM HLI SHAM HLI

Figure 23. (A) Gracilis muscle venule density (expressed as vessel number/myofiber number)

of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light

gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are

presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort.

132

(B) Semi-membranosus muscle venule density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8.

(C) Semi-tendinosus muscle venule density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort.

(D) Gastrocnemius muscle venule density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort.

133

A. Gracilis B. Semi-membranosus

0.05 0.05 *# WT Fgf2 -/- 0.04 * 0.04 * # FGF2 LMW-only FGF2 HMW-only 0.03 0.03 † * * # 0.02 0.02 * † * * Vessel # / Myofiber # Vessel Vessel # / Myofiber # Vessel 0.01 0.01

0 0 SHAM HLI SHAM HLI

C. D. Gastrocnemius Semi-tendinosus 0.05 0.05 WT * # Fgf2 -/- *# 0.04 FGF2 LMW-only 0.04 FGF2 HMW-only * 0.03 * * 0.03 * † * * 0.02 0.02 Vessel # / Myofiber # Vessel

0.01 # / Myofiber # Vessel 0.01

0 0 SHAM HLI SHAM HLI

Figure 24. (A) Gracilis muscle arteriole density (expressed as vessel number/myofiber

number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black

134

bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs.

WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(B) Semi-membranosus muscle arteriole density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs.

WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(C) Semi-tendinosus muscle arteriole density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars),

Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs.

WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

(D) Gastrocnemius muscle arteriole density (expressed as vessel number/myofiber number) of sham and ischemic muscles after 42 days of ischemia in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars). Data are presented as mean ± SEM, n= 6 to 8. *p<0.05 vs. SHAM cohort, #p<0.05 vs. WT HLI, †p<0.05 vs. FGF2 HMW-only HLI.

135

G racilis Semi-membranosus 100 200

75 150

50 100

25 50 W et m uscle w eight (m g) W et m uscle w eight (m g) 0 0 SHAM HLI SHAM HLI

Semi-tendinosus G astrocnem ius WT 250 200 Fgf2 -/- FG F2 LM W -only FG F2 H M W -only 200 150

150 100 100

50 50

W m et weights uscle (m g) W m et weights uscle (m g) 0 0 SHAM HLI SHAM HLI

Figure 25. Wet weights of sham and ischemic gracilis (A), semi-membranosus (B), semi-

tendinosus (C) and gastrocnemius (D) muscles. Sham or ischemic WT, Fgf2-/-, FGF2 LMW- only and FGF2 HMW-only muscles were collected at 42 days of ischemia.

136

Recovery of ischemic foot perfusion is unaffected by targeted deletion of LMW or HMW

FGF2

To determine the role of differential levels of ischemia-induced vascularization on limb perfusion, blood flow in the sham and ischemic limbs was analyzed over time. The non-invasive nature of the laser Doppler imaging allowed for baseline measurements before surgery and several serial measurements up to 42 days after hindlimb ischemia surgery from the non-ischemic (sham) and ischemic lower limbs (paws) (Leahy, de Mul, Nilsson, & Maniewski, 1999; Wårdell,

Jakobsson, & Nilsson, 1993). The lower hindlimbs (paws, and below the site of ischemic injury) of anesthetized WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice were scanned in triplicates and color-coded images displaying the resting microcirculation flow distribution were generated (Figure 26A). Perfusion (an index of blood flow) was calculated from images and expressed as the ratio of the mean perfusion values of the ischemic to non-ischemic (sham) feet.

Basal resting perfusion ratios for all the mouse groups were similar and approximately equal to 1 (Figure 26B). This confirms the histological evidence presented previously that ablation of LMW, HMW or all FGF2 alone does not alter basal vascularization (capillaries, specifically) in skeletal muscle. After induction of hindlimb ischemia (day 1), perfusion ratios were reduced drastically to between 6 and 14% of pre-ischemia values. Perfusion improved gradually over time and at day 14 of ischemia, up to 46% of baseline recovery had been recovered. At day 42 of ischemia, all groups had similar recovery of perfusion with perfusion ratios between 52 and 56% of baseline values. No significant differences were observed between the perfusion ratios of

Fgf2-/-, FGF2 LMW-only or FGF2 HMW-only mice relative to WT at either day 1, 14 or 42 of ischemia. In conclusion, these results illustrate that the LDPI method was not able to detect any

137 changes in post-ischemic resting microcirculatory flow in the presence of only HMW FGF2, only

LMW FGF2 or no FGF2.

138

POST ISCHEMIA SURGERY A. BASELINE DAY 1 DAY 14 DAY 42

WT

Fgf2-/-

FGF2 LMW- only

FGF2 HMW- only

HLI SHAM

PERFUSION DISTRIBUTION

MAX MIN

139

B. WT 1.2 Fgf2 -/- FGF2 LMW-only 1.0 FGF2 HMW-only

0.8

0.6 Perfusion Ratio Ratio Perfusion (Ischemic/Sham) 0.4

0.2

0.0 Baseline Day 1 Day 14 Day 42

Figure 26. (A) Representative laser Doppler perfusion images of lower limbs (i.e., footpad) of WT, Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only mice before (Baseline) and serially

(Days 1, 14 and 42) after surgery. The ischemic (HLI) limb is pictured on the left in each image and the non-ischemic limb (SHAM) is on the right. In each image, maximum (high) perfusion is signified by the red colors, yellow and green depict the mid-range and the blues indicate minimum (low) perfusion. (B) Recovery of footpad perfusion during chronic ischemia of WT (blue), Fgf2-/- (red), FGF2 LMW-only (green), and FGF2 HMW-only

(black) mice. Quantitative analysis of foot perfusion over time computed from color-coded images. Perfusion ratio is average perfusion value of the ischemic to sham (non-ischemic) limbs. Data are presented as mean ± SEM, n=4.

140

DISCUSSION I

The results presented in this dissertation demonstrate, for the first time, that the endogenous

HMW FGF2 isoforms are important and necessary for the protection of skeletal muscle from

chronic ischemia. In the presence of only the HMW isoforms, limb impairment and necrosis were

abolished in the ischemic mouse hindlimb. The salvage of the ischemic muscles was associated,

in part, with enhanced FGF2 expression and revascularization (angiogenesis, arteriogenesis and

collateralization). Early preservation of ischemic limb function occurred in parallel with increased

remodeling/development of small and medium sized collateral vessels which suggests a protection

of skeletal muscle that was independent of the recovery of muscle perfusion (Figure 27).

Research into therapeutic interventions with the potential to ameliorate/treat the morbidity

and poor quality of life associated with coronary or peripheral arterial disease has necessitated the

use of pre-clinical models. In vivo experimental models with a three-dimensional representation

of the whole organ (muscle and vasculature) that could reproducibly recapitulate symptoms of the human disease conditions are the ideal choice (Hoefer, van Royen, & Jost, 2006; Limbourg et al.,

2009; Lotfi et al., 2013). The hindlimb ischemia (HLI) model has been used to successfully investigate the tissue response to PAD and to assess the efficacy of candidate therapies (Dragneva et al., 2013; Madeddu et al., 2006; Waters et al., 2004). This is a murine model where unilateral

ligations and/or excisions are performed on the common femoral, common iliac or external iliac

arteries that perfuse the lower periphery (Hoefer et al., 2006; Lotfi et al., 2013; Madeddu et al.,

2006; Waters et al., 2004). In this dissertation, dual ligatures were inserted at proximal and distal

ends of superficial femoral artery/vein pair. The proximal suture was placed at the origin of the

deep femoral branch while the distal ligation occurred at the bifurcation of the saphenous and

popliteal branches (Figure 4). This approach routinely results in a reduction in foot perfusion (as

141

low as 20% of pre-surgery flow) of the ligated limb (Limbourg et al., 2009; Peng et al., 2011;

Sullivan et al., 2002; Westvik et al., 2009). The unligated contralateral side served as an internal

control for comparison. The immediate responses to the dramatic loss of blood supply in this model

range from tissue necrosis, ischemic limb impairment, muscle atrophy to auto-amputation of toes

and/or feet. In order to prevent sufferance of the animals, mice with spontaneous auto-amputation

were euthanized and excluded from the rest of the study.

This dissertation for the first time demonstrated that the endogenous FGF2 isoforms play

opposing functions in protecting ischemic skeletal muscle from necrosis and preserving functional

capacity. Necrosis was largely confined to the lower limbs and comprised of (dis)coloration and/or

loss of one or more toe or toenails. Mice experiencing auto-amputation of toes and/or entire limbs

were excluded from the analysis of necrosis and all other endpoints. Deficiency of HMW FGF2

(expression of only LMW FGF2) had significant necrosis of ischemic limbs (Figure 11). While

extent of necrosis was not different compared with WT limbs (22.6% vs 17.1%), it was

significantly increased compared with HMW FGF2 expressing limbs which had no incidences of

necrosis. Total loss of FGF2 expression (Fgf2-/-) also produced no changes in macroscopic

necrosis relative to WT. Functional recovery of the ischemic hindlimbs was assessed with

treadmill testing. This measure of forced exercise performance was performed at baseline before

the induction of hindlimb ischemia and at 14 days after ischemia surgery (Figure 12). Mice were

tested at a 7o incline using an exercise protocol that began with an initial speed of 7 meters/min

and ended with a maximal speed of 20 meters/min. In mice where only LMW FGF2 or HMW

FGF2 was expressed, exercise performance was not different from WT at baseline or during

ischemia recovery. In the absence of any FGF2 expression (Fgf2-/-), exercise capacity was decreased at baseline and during ischemia recovery (at 14 days).These results are preliminary due

142

to the small sample sizes (≤ 2 mice per group). Nevertheless, they provide some evidence that Fgf2

is important for maintaining hindlimb functional capacity in the presence or absence or ischemia.

As an additional measure of the FGF2-dependent functional response to ischemia, mice

were graded for the spontaneous or stimulated (tail traction) use of their ischemic limbs. Movement

of the ischemic limbs was restricted immediately following surgical procedure and continued for

the first three days following the ischemia surgery where mice gait was altered and presented as

dragging and no weight bearing of the ischemic foot (Table 12). This was observed in the ischemic feet of all WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice (Figure 9). HMW FGF2 expression resulted in a more rapid (~3x) recovery of active foot and limb movement (plantar flexion with flexing of toes and full weight bearing) in contrast to WT mice (10 vs. 28 days). In the presence of only LMW FGF2, mice had a sustained impairment of ischemic limbs and were delayed in regaining full use of their ischemic limbs until 42 days after ischemia surgery. Mice with a complete loss of Fgf2 gene expression (Fgf2-/-) only had a partial recovery of ischemic limb function by the end of the study.

These results suggest that the HMW FGF2 isoforms have a pronounced effect on preserving tissue viability and limb salvage. These data support the hypothesis that HMW FGF2 promotes cell protection and/or survival pathways in injured tissue. Evidence for this supposition was first observed in mice with global loss of Fgf2 expression. In addition to an altered cardiac

phenotype, these mice had delayed wound healing and neuronal deficits (Dono et al., 1998; Miller

et al., 2000; Ortega et al., 1998). The deficits included changes in the architecture of the neocortex

and a decrease in the neuronal density of the motor cortex layers. Additionally, exogenous 21 or

23kDa HMW FGF2 mediated the survival and neurite promoting activities of cultured

mesenphalic dopaminergic neurons and protected them from 6- hydroxydopamine (6-OHDA)-

143 induced neurotoxicity, a model of Parkinson’s disease (Grothe, Schulze, et al., 2000). When

Schwann cells were transfected to overexpress the HMW FGF2 isoforms, they stimulated the survival and functional recovery of co-transplanted dopaminergic micrografts (Timmer et al.,

2004). Also, apomorphine-induced survival of midbrain dopaminergic neurons was associated with increased trafficking of HMW FGF2 (A. Li et al., 2006). HMW FGF2 also displayed survival promoting properties in non-neuronal cell types. Specifically, overexpression of the HMW isoforms and not 18kDa has been found to reduce the cytotoxic effect of lentiviral gene transfer in corneal endothelial cells (Valtink et al., 2012). Transfection of 24kDa FGF2 confers an increased resistance to radiation-induced cell death in NIH3T3 fibroblasts and HeLa cells (Cohen-Jonathan et al., 1997; Delrieu, Arnaud, Ferjoux, Bayard, & Faye, 1998). 24kDa FGF2 also enhances the survival of rat bladder carcinoma cells (Thomas-Mudge et al., 2004).

The presence of only HMW FGF2 isoforms is necessary to protect chronically ischemic hindlimb muscle from necrosis. In the heart, however, a requirement of both LMW and HMW

FGF2 for protection from cell death has been observed in response to acute cardiac ischemia- reperfusion injury (Liao et al., 2007, 2009, 2010). Correspondingly, hearts overexpressing all of the human FGF2 isoforms were protected from myocardial infarction (MI) (House et al., 2003).

Similar decreases of infarct sizes were detected at 24 hours post-MI in hearts treated with exogenous LMW or HMW FGF2 (Jiang et al., 2007). Taken together, these results paint a picture of a tissue-dependent (cardiac vs. skeletal muscle) and stimulus-dependent (acute vs. chronic ischemia) regulation by the FGF2 isoforms. Examples of the context-dependent properties of the

FGF2 isoforms include the differing (opposing) functions in peripheral never repair, bone homeostasis and fracture repair (Haastert et al., 2006; Homer-Bouthiette, Doetschman, Xiao, &

Hurley, 2014; Xiao et al., 2014).

144

One caveat to note when comparing the results of one study with other rodent models of

ischemia is the dependence of the level of necrosis and recovery of limb use on the age, gender

and strain of the animals utilized. Investigators often utilize only young, male mice in their studies.

However, age and gender have been shown to significantly alter the response to hindlimb ischemia

in rodent models. Aged mice (≥13 months) subjected to ischemia have a slower recovery of foot

movement and increased necrosis relative to younger mice (3 months) (Bosch-Marce et al., 2007;

Faber et al., 2011; Rivard et al., 1999; Westvik et al., 2009). At 7 days after ischemia surgery, male mice have been shown to have improved hindlimb use and similar extent of necrosis compared to their females cohorts (Peng et al., 2011). All the hindlimb ischemic surgeries for the experiments in this dissertation were performed on age-matched (~3 months) male and female mice. The results of the necrosis experiments were stratified by gender and compared for differences (Figures 10B,

C). Within each genotype, no differences in necrosis were present between male and female mice.

Genetic backgrounds should also be considered during comparisons of ischemic mouse models (Limbourg et al., 2009; Lotfi et al., 2013). Several studies have shown that differences in inbred strains influence the severity and incidence of tissue necrosis and ischemic limb recovery.

When comparing the two most commonly utilized strains in limb ischemia studies (C57BL/6 and

BALB/c), the Schaper group observed that wildtype C57BL/6 mice had a more rapid recovery of hindlimb use function compared to the latter strain (Helisch et al., 2006). These findings were independently confirmed in two other studies and the latter group identified a similar pattern of susceptibility in the appearance of ischemic damage (Chalothorn, Clayton, Zhang, Pomp, & Faber,

2007; Dokun et al., 2008). This group also observed that A/J strain mice had an increased impairment of limb use and necrosis than BALB/c mice (Chalothorn & Faber, 2010). Overall, the

C57BL/6 and BALB/c strains can be considered to represent two extremes of the strain-dependent

145

responses to chronic hindlimb ischemia where the C57BL6 strain are ischemia-protected and the

BALB/C mice are ischemia-vulnerable. The 129/Sv strain is less widely utilized in ischemia studies; Helisch and colleagues compared this strain to C57BL/6 and BALB/c mice and found their level of tissue necrosis to be intermediate between the two extremes (Helisch et al., 2006).

This serves as a template for comparison of the mouse models used in this dissertation which were

bred on a mixed background of 50% 129/Sv and 50% Black Swiss (BS) (Azhar et al., 2009;

Garmy-Susini et al., 2004; Zhou et al., 1998). Several ischemia studies have been performed in mice bred on a 50% 129/Sv and 50% C57BL/6 background. The average incidence of ischemic damage in WT hindlimbs across these studies was ~14% (Chalothorn, Zhang, Clayton, Thomas,

& Faber, 2005; Lizotte et al., 2013; Rey et al., 2009). The levels ranged from a complete absence of necrosis to an incidence as high as 33%. Recovery from functional impairment of ischemic limbs was also observed in 129 mice with normal use of the ischemic limb returning at 28 days after ischemic surgery (Bosch-Marce et al., 2007; Rey et al., 2009). These incidences of necrosis

are consistent with the results observed in the WT 129SV/BS mice studied for this dissertation.

Moreover, the complete recovery of functional use of the WT ischemic limbs occurs within a

similar time frame as that observed by others in 129SV mixed background mice.

The impairment and necrosis of the ischemic limbs examined serve as clinically relevant

endpoints for assessing the role of the FGF2 isoforms in limb salvage and function. These

observations in the mouse ischemic limb are similar to those observed in human patients of PAD

whose symptoms present clinically as either intermittent claudication (IC) or critical limb ischemic

(CLI) (Lotfi et al., 2013; Rutherford et al., 1997; Waters et al., 2004). Another relevant endpoint

is the quantitative evaluation of ischemia-induced of vascular adaptation and regeneration.

Vascular growth in response to arterial obstruction occurs via two distinct processes that differ in

146

their vessel type and triggering stimulus (Limbourg et al., 2009; Schaper, 2009). Angiogenesis,

the growth (sprouting or intussception) of capillary networks from pre-existing capillaries is initiated by hypoxia in ischemic tissues while arteriogenesis, the enlargement (lumen area and wall thickness) of pre-existing collateral arterioles into larger conductance vessels is stimulated by the increase of shear forces (Fischer et al., 2006; Heil et al., 2006). The type(s) of vascular growth processes that occurs during hindlimb ischemia is highly dependent on the nature of the vascular injury. The surgical ligation (proximal and distal ligatures) and excision of the intervening segments of the femoral artery as performed for this study induces simultaneous occurrence of angiogenesis and arteriogenesis in the ischemic hindlimb (Hoefer et al., 2006; Lotfi et al., 2013;

Waters et al., 2004). In light of this, ischemic muscle tissue from the thighs where arteriogenesis

predominates and tissue from distal limb (calf) where angiogenesis occurs were collected and

analyzed (Hoefer et al., 2006; Limbourg et al., 2009). Muscles of the hindlimb from the pelvic

region (near the inguinal ligament) to the knee were defined as thigh muscles and muscles from

the knee to the ankle were designated as calf muscles (Hedrich & Bullock, 2004; Suckow,

Weisbroth, & Franklin, 2006). The gracilis, semitendinosus, and semimembranosus muscles from

the thigh and the gastrocnemius/soleus bundle from the calf were isolated from each ischemic and

sham-operated contralateral limb. In calf or thigh muscles, the extent of angiogenesis and arteriogenesis was quantified by histomorphometric assessment of tissues isolated at 42 days after

the ischemia surgery and presented as the density of capillary and arteriolar vessels respectively.

Results presented in Chapter 1 correlate with studies in the literature where a combination of

angiogenesis and arteriogenesis was present in the ischemic hindlimbs. While capillary vessel

growth was not entirely confined to the gastrocnemius (calf) muscle, greater increases from sham

densities were observed in ischemic distal limb muscles. Likewise, ischemia-induced changes in

147

arteriole densities were higher in the thigh muscles. Mouse hindlimb models have shown that

increases in vascular density of ischemic limbs occur as early 7 days (Couffinhal et al., 1998;

Scholz et al., 2002; Sullivan et al., 2002) and persist up to days 28, 35, (Boerckel et al., 2014;

Meisner, Song, Annex, & Price, 2013; Ruifrok et al., 2009) and 42 (J. Zhang et al., 2009). The vascular densities presented here were calculated from skeletal muscles isolated at 14 and 42 days postoperatively to allow for stabilization of vascular remodeling in the ischemic tissues.

These data, for the first time, represent novel evidence of a beneficial role for HMW FGF2 in stimulating vascular growth, repair and remodeling during chronic ischemia. Mice with expression of only HMW FGF2 had an increased level of post-ischemic capillary density suggesting that the HMW isoforms have angiogenic properties in response to ischemic injury

(Figures 16, 18). The effect of endogenous HMW FGF2 on angiogenesis has rarely been examined.

In fact, the sole report on the role of endogenous HMW FGF2 in angiogenesis is from the Arnal group who showed that the endogenous HMW isoforms mediate the effect of estradiol (E2) on

angiogenesis in vitro (Garmy-Susini et al., 2004). Specifically, HMW FGF2 was necessary for the

stimulating effects of E2 on endothelial cell migration. This important study was limited in its use

of in vitro angiogenesis models as well as its focus on the indirect effects of HMW FGF2. The effect of exogenous recombinant HMW FGF2 on in vitro angiogenesis has also been examined.

Administration of HMW FGF2 (24kDa) to cultured endothelial cells was shown to promote cell

proliferation and inhibit migration (Piotrowicz et al., 1999). This dissertation serves to address the

paucity of research of the direct and in vivo roles of the endogenous HMW isoforms of FGF2.

The expression of only LMW FGF2 in ischemic limbs did not improve capillary growth.

In fact, the density of capillaries in FGF2 LMW-only limbs were similar compared to control

(sham) level while the FGF2 HMW-only muscles had the greatest proportional increase of

148

capillary vessel numbers. The observed angiogenic response to LMW FGF2 in ischemic hindlimbs

is not consistent with some previous research showing that when given exogenously, the LMW

isoform promotes angiogenesis in vitro and in in vivo models of ischemia. Vascular cells (EC and

vSMCs) respond to treatment with the recombinant LMW isoform in several angiogenesis assays

including cell proliferation, migration or tube formation (Davis et al., 1997; Piotrowicz et al., 1999;

Presta et al., 2005, 2007; Shing et al., 1985). In in vivo models of peripheral artery ligation,

infusion of exogenous FGF2 (low molecular weight form) into ischemic muscle has been shown

to induce capillary growth (Asahara et al., 1995; Baffour et al., 1992; Bush et al., 1998; H.-T. Yang

et al., 1996). In contrast, no angiogenesis was detected in FGF2 LMW-only ischemic limbs. This

corresponds well with data from intra-arterial infusion of recombinant LMW FGF2 to ischemic

rabbit limbs and/or injection into infarcted porcine hearts where angiogenesis was inhibited by the

growth factor (Scholz et al., 2002; Watanabe et al., 1998). The incompatibility between the results presented here and those that show increased capillary growth with LMW FGF2 treatment could be due to several reasons. First, it cannot be overstated that the cellular response to a growth factor

in cell culture is vastly different from the conditions that exist in vivo where cell phenotype is

dependent upon a coordinated physiological response. Cell culture experiments may likely

overestimated the functions of the growth factor. Second, exogenous LMW FGF2 was freely

administered to the ischemic limbs in these studies via osmotic mini-pumps over a period of 7-10

days while the endogenous expression of FGF2 is likely under a large degree of gene regulation

and/or protein sequestration that is only lessened in response to specific stimuli. Third, the level

of endogenous protein released in response to the ischemic stress may not be comparable to the

pharmacologic (supra-physiological) levels of the exogenous LMW FGF2. Fourth, it is worth

noting that while recombinant LMW FGF2 is capable of inducing angiogenesis in vitro and in

149

vivo, endogenous LMW FGF2 may not possess the potent “angiogenic” ability that has been

assigned to it. The field of recombinant protein production is well established and a plethora of

studies have been undertaken that demonstrate recombinant proteins behaving in a similar manner

as their endogenous “equivalents” by showing comparable receptor binding, activation and

downstream signaling in target cells. In spite of this, it is still conceivable that endogenous LMW

FGF2 does not share the angiogenic potential of its exogenous counterpart because of the presence

of in vivo mechanisms that the recombinant proteins are not subjected to. Some evidence for this

discrepancy may exist in the fact that recombinant growth factors including LMW FGF2 failed to

show efficacious benefit in the treatment of human ischemic disease. Promising results in

multitude of small and large animal models of peripheral and coronary artery diseases (PAD,

CAD) propelled the investigation of LMW FGF2 in several placebo-controlled clinical trials

(Laham, Chronos, et al., 2000; Lederman et al., 2002; Molin & Post, 2007). Thus far, the use of

recombinant LMW FGF2 protein for the treatment of CAD or PAD has not been successful (Y.

Cao, Hong, Schulten, & Post, 2005; van Royen, Piek, Schaper, & Fulton, 2009). While there may

be a number of reasons for the failure of these trials including the insufficiency of single growth

factor administration, the mode of and route of growth factor delivery, the presence of

comorbidities in advanced age patients and the selection of appropriate clinical endpoints (Y. Cao,

2009; Dragneva et al., 2013), it is not out of the realm of possibility that the recombinant proteins

employed in these studies are the true culprits. Lastly, the differences in species, surgical ischemia

models and data interpretation may also explain the conflicting results. This dissertation provides

novel evidence of the in vivo roles of endogenous LMW FGF2 isoforms in ischemia-induced angiogenesis.

150

When there was no expression of FGF2 (in Fgf2-/- mice), similar increases in capillary

numbers were detected in response to ischemia (Figures 16 and18). These results are in agreement

with a similar study by Sullivan and colleagues where the absence of Fgf2 gene did not affect post-

ischemic capillary density (Sullivan et al., 2002). Normal injury-induced angiogenesis was also

observed in the eye when Fgf2-/- mice were subjected to ischemic retinopathy or choroidal injury

(Ozaki et al., 1998; Tobe et al., 1998). Postnatal cardiac capillary growth is also affected by loss

of the Fgf2 gene. Complete (Fgf2-/-) or partial (Fgf2+/-) disruption of FGF2 gene expression

impairs adult myocardial capillarogenesis (Amann et al., 2006). On the other hand, our laboratory

and others detected no alterations to baseline capillary development in Fgf2-/- cardiac or skeletal

muscles (House et al., 2003; Liao et al., 2009; Sullivan et al., 2002). The present study also did not discover any changes to capillary vessel development in non-ischemic skeletal muscle (Figures

16, 19). Compensation for the loss of Fgf2 by FGF1, the most closely related FGF was thought account for the absence of a phenotypic response to the genetic deletion of FGF2. However, when

Fgf2-/- mice were crossed with Fgf1 knockouts, the resultant double knockout offspring displayed similar phenotypes as the Fgf2-/- mice (Miller et al., 2000). By the same token, Fgf5/Fgf6,

Fgf6/Fgf7, and Fgf5/fgf6/Fgf7 double and triple mutants did not exhibit any defects beyond that observed in Fgf6-/- skeletal muscles that could account for the residual regeneration ability of Fgf6 mutant myotubes (Armand et al., 2005; Floss, Arnold, & Braun, 1997; Neuhaus et al., 2003). The advent of mouse models in which total Fgf2 expression is not merely disrupted but where specific expression of only one class of its protein isoforms are deleted allows for a more nuanced examination of the functions of FGF2.

In addition to its role in ischemia-induced capillary growth, the effect of expression of

HMW FGF2 on arteriole density was also examined (Figures 20, 22B). The arteriole density of

151

ischemic HMW FGF2-expressing muscles was nearly two-fold greater than WT ischemic levels

(Figure 22B). This indicates that in the presence of shear and ischemic stress, the HMW isoforms

have a profound stimulatory effect on arteriogenesis and represents a previously unrecognized

function of endogenous HMW FGF2. While no other direct evidence of the role of HMW isoforms

on ischemia-induced revascularization exists, findings from Jiang and coworkers showed the effect

of exogenous HMW FGF2 administration in infarcted rat hearts (Jiang et al., 2007). Specifically,

recombinant 23kDa produced no changes to arteriole vessel density in the infarct area. The

differential outcomes provide additional proof to distinguish the properties of endogenously

produced FGF2 from the exogenous isoforms. Nonetheless, the possibility that the conflicting

results could be because of differences in ischemia models and species remains. The effect of

expression of only the LMW isoforms on arteriogenesis was somewhat intriguing in light of the

total lack of angiogenic response observed in our model. LMW FGF2 stimulated an increase

arteriole and venule density in ischemic muscle (Figure 22A). The degree of arterial growth was

identical to that observed in WT limbs where arteriole numbers were elevated almost four-fold in the presence of ischemia. These results are consistent with other hindlimb ischemia studies performed in mice (Scholz et al., 2002) and rats (J. Zhang et al., 2009) where LMW FGF2

treatment stimulated arteriolar growth. Porcine (Watanabe et al., 1998) and murine (Jiang et al.,

2007) hearts undergoing chronic MI also had increased arteriole density in response to LMW FGF2

administration. It is worth noting that while the LMW FGF2 was capable of promoting

arteriogenesis in ischemic limbs, the HMW isoforms produced a greater effect on collateral vessel

growth. Resultantly, the number of arteriolar vessels in the ischemic FGF2 HMW-only muscles

was nearly two-fold greater than in the LMW FGF2-only cohorts.

152

As previously mentioned, ischemic limbs of Fgf2-/- mice had no deficits in angiogenesis

(capillary density). Arteriogenesis was also found to be normal in these mice (Figures 20, 22B).

These findings paralleled work by Sullivan et al. whose study also revealed similar arteriole

density in Fgf2-/- ischemic muscle as in WT (Sullivan et al., 2002). Conversely, Virag and co- workers have reported reduced density of smooth muscle actin containing-vessels in Fgf2-/- hearts after 4 weeks permanent coronary artery ligation (Virag et al., 2007). The discrepancies in these outcomes could be related differences in the site of ischemia induction or animal strain. Overall, these results might be surprising when they are viewed in the context of some of the earlier studies performed that identified FGF2 as a “potent angiogenic” molecule. Furthermore, the absence of vascular abnormalities/changes in Fgf2-/- mice at birth or in response to hindlimb ischemia led to questioning of the earlier in vitro work establishing FGF2 in this role. In fact, it suggested that the functions of FGF2 cannot always be gleamed from the use of Fgf2-/- where all of the isoforms are absent as this mode presumes that all the isoforms have similar functions. The availability of animal models in which total Fgf2 expression is not only deleted, but where selective expression of only HMW or LMW FGF2 is present will provide a clear picture of the functional roles of

FGF2. When only LMW FGF2 (in FGF2 HMW-only mice) or only HMW FGF2 (in FGF2 LMW- only mice) is deleted, there appears to be greater vascular physiology effect than in Fgf2-/- mice.

In the presence of only the HMW isoforms, FGF2 has a positive effect on revascularization and the opposite occurs in presence of only the LMW isoform. The results presented in this dissertation serve as evidence for the antithetical roles of the FGF2 isoforms in ischemia-induced angiogenesis or arteriogenesis.

In light of the previously addressed differences that are observed in the response to ischemia by mice of different background strains, the angiogenic and arteriogenic responses of the

153

WT mice in this study were compared to other studies of hindlimb ischemia. This also ensured that all changes observed were due only to selective or total deletion of FGF2 expression and not based on differences or inconsistencies in the surgical ischemia method. In five studies utilizing the surgical hindlimb ischemia method in WT mice of similar strain (50% 129SV), the average capillary and arteriole density in ischemic muscles was increased by 1.4- and 1.8-fold respectively

(Chalothorn et al., 2005; Lizotte et al., 2013; Scholz et al., 2002; Sullivan et al., 2002; Tilan et al.,

2013). The changes in WT capillary and arteriole density observed in this dissertation were 1.3- and 4.0-fold, respectively. The source of the differences in arteriole density between our work and published results may lie in the subjective criteria each group uses to defined arterioles. Positive staining of a smooth muscle cell marker (α-smooth muscle actin) and morphology was used in the identification of arterioles/arteries as well as to distinguish them from venules/veins. This classification is not always performed by others and the arteriole density reported is actually a measure of all vessels that are positive for smooth muscle. Another technical detail that might confound comparisons with other published work is the choice of ischemic muscle bed chosen for histological assessment of vascular growth. The importance of distinctions between calf and thigh vessel growth type has been addressed previously. Overall, the level of angiogenesis and arteriogenesis present in the WT mice were consistent with similar hindlimb ischemia studies.

In the absence of ischemia, the level of vessel density in FGF2 LMW-only and FGF2

HMW-only limbs were similar as WT (as seen in sham-operated skeletal muscle). The lack of baseline differences was also present in hearts with specific FGF2 isoform deletions (Liao et al.,

2007, 2010). Similarly, mice with chronic overexpression of all FGF2 isoforms or only HMW

FGF2 had no differences in baseline cardiac vessel density (House et al., 2003; Liao et al., 2010;

Virag et al., 2007). However, the Kardami group detected an increase (20%) in the capillary

154 density of LMW FGF2-overexpressing hearts (Sheikh et al., 2001). On the whole, perturbations in

FGF2 expression alone does not appear to produce vascular growth defects. This dissertation along with a study by the Murry group where Fgf2 overexpression induced increased vessel density during chronic ischemia, has shown that the effect of FGF2 isoforms on vascular growth is dormant at baseline and manifests in response to chronic vascular injury (Virag et al., 2007). This suggests that a latent phenotype of the FGF2 isoforms exists and is only activated during tissue pathophysiology.

The central pathophysiology to manifest in human diseases of arterial occlusion is the loss or reduction of tissue perfusion. Accordingly, preclinical studies modeled after these human conditions are set up to identify efficacious treatments that will increase perfusion to ischemic regions. In the mouse hindlimb ischemia model, adaptive or pharmacologically-induced vascular growth/remodeling via angiogenesis, arteriogenesis or collateral growth can all contribute to the restoration or augmentation of muscle blood flow. To establish the role of the FGF2 isoforms in the recovery of limb flow, perfusion of ischemic skeletal muscle was determined using laser

Doppler perfusion imaging (LDPI). Serial perfusion of the sham and ischemic feet (paws) was measured as a surrogate for skeletal muscle blood flow (Figure 26). In FGF2 HMW-only mice, recovery of resting foot perfusion was unchanged from WT. No differences from WT in relative foot perfusion (ischemic-to-sham ratio) were detected at baseline (absence of ischemia), immediately after (day 1 of ischemia), at 2 weeks or at 6 weeks after ischemia surgery. In FGF2

LMW-only mice, relative perfusion was normal in the absence of ischemia. After ischemia surgery, the trend of perfusion recovery described for the HMW FGF2 feet was also present in

LMW FGF2-expressing limbs (Figure 26B). The apparent lack of effect of LMW FGF2 is in alignment with a study where rats receiving LMW FGF2 treatment had no changes in foot

155

perfusion recovery after 6 weeks of ischemia (J. Zhang et al., 2009). Conversely, several studies

in myocardial (Laham, Rezaee, et al., 2000; Lazarous et al., 1995; Rajanayagam et al., 2000; Unger

et al., 1994) and hindlimb (Rakue et al., 1998; Scholz et al., 2002; H.-T. Yang et al., 1996) ischemia

models have reported improved regional blood flow after administration of recombinant LMW

FGF2.

The presence of species-to-species variations in blood flow recovery after myocardial or

peripheral arterial occlusion has been previously reported (Maxwell et al., 1987; Teunissen et al.,

2012; Waters et al., 2004). Similarly strain-to-strain variations exist within the same species and

have been detected in several mouse genetic strains (Scholz et al., 2002). To address changes in perfusion that could be strain dependent, the level of foot perfusion recovery in WT mice was compared to similar reports in 129SV WT mice. Perfusion ratios typically reached a plateau at ~28 days and the recovery of resting perfusion in the ischemic feet was as low as 50% (Landázuri,

Joseph, Guldberg, & Taylor, 2012; Rey et al., 2009; Tirziu et al., 2005) and as high as 90%

(Sullivan et al., 2002) of baseline levels. When averaged across several published studies, the recovery of resting perfusion in 129SV WT was ~60% (Bosch-Marce et al., 2007; Chalothorn et al., 2005; Heil et al., 2002; Helisch et al., 2006; Tilan et al., 2013). A similar recovery of pre- ischemia perfusion levels was present in the WT mice studied in this dissertation (Figure 26). An interesting observation was the similarity between WT, FGF2 LMW-only and FGF2 HMW-only mice in resting perfusion recovery of ischemic limbs despite their differential levels of angio- and/or arteriogenesis. The prevailing paradigm in the study of peripheral ischemia using experimental models suggests that increases or decreases in ischemic muscle vascular growth present should lead to improved or reduced limb perfusion, respectively (Couffinhal, Dufourcq,

Barandon, Leroux, & Duplàa, 2009; Limbourg et al., 2009; Silvestre et al., 2013; Waters et al.,

156

2004). While this holds true for some studies, there are published reports where changes in limb

perfusion do not always follow enhanced vascularization (L. S. Brevetti et al., 2001; Chalothorn et al., 2005; Shireman & Quinones, 2005; J. Zhang et al., 2009). The apparent discrepancy between

muscle vascular density and perfusion could be attributed to several factors including the effect of

ischemia on blood flow control, regulation of vasomotor tone by FGF2 expression, and the

perfusion measurement methodology.

Chronic ischemia alone has been shown to negatively impact blood flow control. This

phenomenon is present in animal models as well human patients of arterial occlusive disease

(Dragneva et al., 2013). Intermittent claudication patients have normal muscle perfusion at rest but

can experience ischemic pain during walking or exercise (Annex, 2013; Brass, 2013; Muir, 2009).

The impaired flow control manifests as altered vascular tone and reactivity where the

repairing/remodeled vasculature exhibits reduced vasodilation and/or enhanced vasoconstriction

(Colleran, Li, Yang, Laughlin, & Terjung, 2010; Kelsall, Brown, & Hudlicka, 2001; LeBlanc et

al., 2012). The presence of endothelial and smooth muscle dysfunction in ischemic muscles has

been confirmed by decreased reactivity to acetylcholine or sodium nitroprusside (Cardinal et al.,

2011; Kelsall, Brown, Kent, Kloehn, & Hudlicka, 2004). Another possible confounding variable

to consider when interpreting the results of the perfusion measurements is the effect of FGF2 itself

on regulation of vasomotor tone. Administration of LMW FGF2 to isolated arterioles of cerebral,

(Kajita, Takayasu, Yoshida, Dietrich, & Dacey, 2001), coronary (Tiefenbacher & Chilian, 1997)

or skeletal muscle (H. M. Wu et al., 1996) origins and infarcted porcine hearts (Unger et al., 1994)

induced arterial vasodilation. This effect was dependent on the presence of an intact endothelium,

nitric oxide production (NO) and activation of ATP-sensitive potassium channels (Tiefenbacher

& Chilian, 1997). When the Fgf2-/- mice were first produced to characterize the role of the Fgf2

157 gene in vascular physiology, a significant phenotype present in this mice was decreased vascular tone (Zhou et al., 1998) and basal hypotension (Dono et al., 1998; Ortega et al., 1998; Zhou et al.,

1998). The vascular tone control exerted by the presence of Fgf2 (via LMW FGF2) could account for the lack of differences observed in the recovery of ischemic perfusion of FGF2 LMW-only mice compared to WT. This would suggest that increased muscle perfusion in these mice is not reflective of changes in vascularization effects but due to the positive vasodilatory effects of LMW

FGF2. Studies by Sullivan and colleagues in Fgf2-/- mice provide some evidence to support this theory. They observed that despite normal resting perfusion in Fgf2-/- ischemic limbs, the reactive hyperemic response to acute iliac artery occlusion (vasodilation followed by a transient blood flow increase) was diminished in Fgf2-/- ischemic limbs at 14 days of ischemia (Sullivan et al., 2002;

Sullivan, 2002). In light of this, the effect of the specific FGF2 isoforms on vasomotor tone at baseline and during the ischemia repair process would be very important. Potential differences in vasomotor tone control due to the presence of ischemia and/or altered FGF2 expression could be addressed by repeating the LDPI measurements after brief periods of acute ischemic (via balloon cuff inflation) or bouts of exercise to induce reactive or functional hyperemia, respectively. Both challenges sidestep the low resting blood flow of skeletal muscle and serve to increase the demand for O2 in the ischemic tissues and requires the maximal dilation of the ischemic vasculature (L. S.

Brevetti et al., 2001; Ziegler et al., 2010). They can test the reactivity and capacity of the FGF2

LMW or HMW-only repairing vasculature (both the endothelium and smooth muscle) to respond to increased metabolic demand. Treadmill exercise, electrical nerve stimulation and acute ischemia have been shown to expose deficits in regional blood flow that are not present in resting (non- exercised) ischemic skeletal muscle (L. S. Brevetti et al., 2001; Cardinal & Hoying, 2007; Helisch et al., 2006; Hudlicka, Brown, Egginton, & Dawson, 1994). Treadmill exercise testing as

158

previously described was performed in a small cohort of mice in this dissertation (Figure 10).

While these experiments were performed to determine limb functional capacity and did not include perfusion measurements, it is noteworthy that in the absence of Fgf2 expression, ischemic limb

function is impaired.

There are several alternative strategies to the present method for determining blood flow

changes in ischemia models including arterial flow probes, X-ray angiography, magnetic

resonance imaging, and radio- or fluorescently-labelled microspheres (de Lussanet et al., 2007;

Lotfi et al., 2013; Madeddu et al., 2006; Waters et al., 2004). The use of microspheres is considered the gold standard for the evaluation of blood flow in hindlimb ischemia models (Lotfi et al., 2013;

Madeddu et al., 2006; Waters et al., 2004). The method utilizes labelled microspheres (usually

15µm) that are injected intra-arterially or into the left ventricle. Regional blood flow is directly

proportional to the number of microspores that becomes lodged in the tissue of interest (Deveci &

Egginton, 1999; Prinzen & Bassingthwaighte, 2000). This approach is highly invasive and

technically challenging in smaller sized animal models as it requires both arterial catheterization

and withdrawal of a reference blood sample at a fixed rate. This blood sampling can lead to

hypotension due to small blood volumes and limits the use of microspheres in longitudinal studies

(Lotfi et al., 2013). Laser Doppler Perfusion Imaging (LDPI) provides a measure of tissue perfusion based on the magnitude and frequency distribution of the Doppler shift in the laser light after it is scattered by moving erythrocytes (Wårdell et al., 1993). The reflected Doppler signal is linearly proportional to the underlying surface tissue perfusion (Leahy et al., 1999). Disadvantages of this method include the intrinsic depth of penetration of the scanning laser (1-2 mm) which limits the flow measured to superficial skin circulation only (Madeddu et al., 2006) and the use of

this cutaneous perfusion as a surrogate measure of blood flow. Another important technical

159

limitation of LDPI is the region/portion of the mouse hindlimbs that are scanned and analyzed.

Animal fur and skin pigmentation of the upper hindlimb can interfere with LDPI signal and

eliminates the upper thigh area of the hindlimb as a suitable region for measurement (Hoefer et al.,

2006; Limbourg et al., 2009). Therefore, perfusion data reported in this dissertation and in other

studies with the mouse HLI model were quantified from the plantar sole (feet) of the ischemic and

sham legs (Figure 26). Another important caveat to consider when evaluating perfusion in rodents

using LDPI is the animals’ core body temperature. Animals should be monitored during scanning

to ensure that no fluctuation in core body temperature occurs (≤ 0.5oC). To that end, all animals

were placed on a warming pad and equipped with a rectal temperature probe to monitor body

temperature during the perfusion experiments. Core body temperature for the mice scanned in this

study was an average of 36.8oC.

Despite the outlined shortfalls of the laser Doppler technique for the assessment of skeletal

muscle perfusion, it is still widely accepted as an appropriate tool for testing the vascular/tissue

responsiveness to hindlimb ischemia. It has the advantage of being a non-invasive and repeatable technique. Furthermore, longitudinal experiments can be carried to measure perfusion recovery over a long period (weeks-months). It is extremely useful for identifying blood flow deficits in the ischemic feet relative to the non-ischemic contralateral foot immediately after the ischemia surgery

(1-24 hours) which confirms successful ligation of the femoral artery (Figure 26, Day 1 perfusion ratios) (Madeddu et al., 2006; Waters et al., 2004). However, this method may be less sensitive in detecting subtler changes in flow and this could account for the uncoupling of the changes in vascular density quantified by histology and muscle perfusion.

An alternative quantitative assessment of FGF2 isoform-induced revascularization was performed with the use of micro-computed tomography (micro-CT). This imaging modality

160

addresses the limitations inherent within the previously mentioned techniques for quantifying

vascular growth including the two-dimensional (2-D) nature and small sampling of both

histological and LDPI measurements, the subjectivity of histology and the superficial and resting

perfusion analyzed by the latter (Duvall, Taylor, Weiss, & Guldberg, 2004; Lotfi et al., 2013;

Madeddu et al., 2006; Ziegler et al., 2010). While these other methods are still more widely

utilized, micro-CT is advancing the field of microvascular imaging by providing a means for

obtaining high resolution, objective and volumetric analysis of vascular anatomy (A. S. Lin et al.,

2007; Zagorchev et al., 2010). Unlike bone which has high inherent radiopacity, micro-CT imaging

of soft tissues like blood vessels require the use of contrast agents to enhance the x-ray attenuation of the ischemic vascular network. Several preliminary studies were performed with the perfusion of non-ischemic hindlimbs WT with three different contrast agents; barium sulfate, lead oxide and lead chromate (Figures 5-7). Barium sulfate (suspended in 5% gelatin) did not perfuse vessels well and while lead oxide produced images that were not suitable for quantitative analysis due to the extremely high x-ray attenuation of the perfused vessels. Lead chromate suspended in a silicon polymer solution was selected based on its ease of vascular perfusion, viscosity, and homogeneity.

The hindlimbs were scanned 42 days after ischemia surgery to produce three-dimensional

(3-D) representations of the vasculature in sham and ischemic calf or thigh regions (Figure 12).

The definition of separate regions of interest (ROI) within each 3-D volume allows for evaluation

of angiogenesis and arteriogenesis in the upper limb defined as thigh and lower limb as calf,

respectively. For each ischemic calf or thigh ROI, vessels were analyzed for the relative (% change

from sham) vessel volume, number (density) and spacing. In FGF2 HMW-only mice, vascular volumes in the ischemic calves and thighs were greater than in WT cohorts (Figure 13A). This is in agreement with the results of the immunohistological labeling of vessels described previously

161

where capillary and arteriole density was greatly increased in the presence of FGF2 HMW. The

volume of vessels in the FGF2 LMW-only calf or thigh also corroborated the 2-D histological

data, showing that the presence of only LMW FGF2 inhibited production of vessels in response to

ischemia. By the same token, changes in vessel density and vessel spacing due to ischemia were

both decreased (Figures 13B, C). The spacing parameter represents the vessel separation and is a

measure of the volume of the scanned limbs not occupied by vessels (Bouxsein et al., 2010a;

Duvall et al., 2004; A. S. Lin et al., 2007). This means that as the vascular volume increases, there

is a corresponding decrease in the amount of vessel separation.

In Fgf2-/- calf and thigh regions, only the changes in vascular volume were in agreement

with the trend observed in the histology results were the loss of Fgf2 expression did not alter

vascular growth. Vessel density and spacing in the Fgf2-/- limbs were both reduced relative to

WT. This highlights the advantage of the micro-CT method for uncovering alterations in vascular

growth that are beyond the ability of immunohistochemistry. The loss of vascular reactivity and

presence of endothelial dysfunction described previously for ischemic Fgf2-/- muscles could be the reason for these differences. The micro-CT technique is dependent on the utility of contrast agents to enhance the X-ray attenuation of hydrated soft tissues (Bouxsein et al., 2010b; A. S. Lin et al., 2007; Zagorchev et al., 2010). The contrast material is perfused at constant pressure into

vessels previously flushed with a cocktail of endothelium and smooth muscle-responsive vasodilators to produce maximal vasodilatation. The altered vascular tone and reactivity present in Fgf2-/- under hyperemic conditions would suggest that response to this cocktail were limited and produced the observed changes in the 3-D volumes (Cardinal et al., 2011; Sullivan, 2002).

The use of a contrast agent in vessels for micro-CT imaging of vessels highlights both a strength and drawback of this technique. Incomplete perfusion of vessels, artifacts from high or

162

inconstant perfusion pressure or vascular leakage can lead to inhomogeneous attenuation and

erroneous quantitative analysis. An additional limitation is the post-mortem nature of the tissue preparation (Bouxsein et al., 2010a; Zagorchev et al., 2010). This prohibits its use for longitudinal studies and increases the cost and number of animals required for each study. The time point of 42 days after the ischemia surgery was chosen to allow for comparison with the histological counts of vascular density. While micro-CT is not as widely used as LDPI for the analysis of the mouse models of ischemic vasculature due to its cost and/or technical difficulty, there have been a number of reports where this methodology was successfully employed (Duvall et al., 2008, 2004; W. Li et al., 2006; Oses et al., 2009). These studies largely focused on the distribution of vessel thickness present in the ischemic hindlimbs. This evaluation provides additional information on trends in the development of the post-ischemic vasculature that may not be evident from the averaged assessments of vascular morphology (volume, number or spacing). To further characterize the observed changes in vascular volume response to specific FGF2 isoform expression, size distribution analysis of the vasculature was performed. Specific isoform deletion or genetic knockout of FGF2 alone did not affect the distribution of vessels in the hindlimb (Figure 14). In the presence of ischemia, expression of the HMW isoforms alone stimulated an increase in the development of small (36-200µm) and medium (200-400µm)-sized collateral vessels.

Contrastingly, the LMW FGF2-expressing limbs had no change in the level of small or medium collateral vessels after 42 days of ischemia. These results closely mirror the responses observed in the histological staining of vascular structures. WT muscles also had greater numbers of collateral vessels in the ischemic muscles but to a lesser degree than the FGF2 LMW-only limbs. Several hindlimb ischemia studies in WT mice have detected similar increases in the volumes of vessels between 50 and 200µm in diameter (Duvall et al., 2004; Landázuri et al., 2012; W. Li et al., 2006;

163

Morrison et al., 2014). Likewise, enhancement of vessels ranging in size from 200 to 420µm has been reported by others (Duvall et al., 2004; Landázuri et al., 2012; Westvik et al., 2009).

164

Figure 27: Schematic outlining the FGF2 isoform-mediated tissue response to hindlimb ischemia. The nuclear localized HMW FGF2 is protective against necrosis and promotes recovery of limb function. On the other hand, LMW FGF2 is located in the cytosolic and nuclear compartments and may be secreted during ischemia. LMW FGF2 induced arteriogenesis and had no effect on angiogenesis. HMW FGF2 isoforms stimulated a greater degree of arteriogenesis as well as angiogenesis.

Inhibit Activate

165

Chapter 2: Potential mechanisms of the FGF2 isoforms in adaptive revascularization

during chronic ischemia

RESULTS II

Loss of HMW FGF2 leads to increased expression of Fgf2 mRNA in non-ischemic muscle.

The absence of HMW FGF2 expression in non-ischemic hearts results in increased LMW

FGF2 protein expression (Liao et al., 2010). It has yet to be determined if ablation of LMW FGF2 or HMW FGF2 affects overall Fgf2 gene and/or protein expression in skeletal muscle. To determine Fgf2 transcript levels, RNA was isolated from non-ischemic WT, Fgf2-/-, FGF2 LMW- only and FGF2 HMW-only thigh and calf muscles. First strand cDNA was amplified by quantitative real-time PCR (qRT-PCR) using primers for Fgf2 and an internal control gene, 18S.

No Fgf2 mRNA was detected in Fgf2-/- mice which confirmed the deletion of the FGF2 gene in skeletal muscle (Figure 28). Fgf2 mRNA was significantly increased by about three-fold in FGF2

LMW-only muscles relative to non-ischemic WTs (p<0.05). Similarly, FGF2 HMW-only mice had upregulated Fgf2 expression, nearly four times greater than WT (p<0.05). mRNA levels were not different between FGF2 LMW-only and FGF2 HMW-only muscles. Overall, deletion of the

LMW or HMW isoform resulted in elevated Fgf2 gene transcription. This would suggest that there is transcriptional regulation of Fgf2 by the presence of either class of FGF2 isoforms.

166

Figure 28. Fgf2 mRNA expression in non-ischemic muscles. Quantitative RT-PCR analysis

for Fgf2 mRNA expression in non-ischemic skeletal muscle relative to WT. Fgf2-/-, FGF2

LMW-only or FGF2 HMW-only transcript expression levels are represented as fold- changes relative to WT non-ischemic muscles. Data are presented as mean ± SEM. n=4 for each mouse group. *p<0.05 vs. WT.

167

Loss of HMW FGF2 does not affect LMW isoform expression in non-ischemic skeletal

muscle

Genome and proteome-wide studies have shown that mRNA levels in mammalian cells do

not always correlate well with protein abundance (Greenbaum, Colangelo, Williams, & Gerstein,

2003; Vogel & Marcotte, 2012). To determine if selective deletion of either the LMW or HMW

isoforms induced changes in skeletal muscle FGF2 expression at the protein level, Western

blotting was performed on non-ischemic skeletal muscles. As there is currently no specific

antibody for 18kDa, 21kDa or 22kDa FGF2, analysis for the LMW and HMW protein isoforms

was performed simultaneously (Figure 3). Quantitation of the Western blot bands were performed

by densitometry analysis and are presented in Figures 29 and 30. As expected, neither Fgf2-/- nor

FGF2 HMW-only muscles expressed the LMW isoform (Figure 29). There was expression of

18kDa FGF2 in WT and FGF2 LMW-only limbs. There was a trend towards increased expression

of the LMW isoform in the FGF2 LMW-only group relative to WT.

Absence of the LMW isoform leads to increased expression of HMW FGF2 in non-ischemic

muscle

Ablation of the 18kDa LMW FGF2 isoform was confirmed in non-ischemic Fgf2-/- and

FGF2 HMW-only skeletal muscles. WT tissues expressed similar amounts of the 21kDa or 22kDa isoforms. This pattern was also present in FGF2 HMW-only limbs (Figure 30). When compared to WT, expression of each HMW isoform was significantly increased in the FGF2 HMW-only muscles (p<0.05). HMW FGF2 isoform levels were three times higher when LMW is absent than in WT. This would suggest that the FGF2-dependent regulation that is present in transcription of

Fgf2 extends to isoform specific translation.

168

FGF2 isoform expression (Non-ischemic skeletal muscle)

4 WT

Fgf2-/- 3 N.S FGF2 LMW-only

2

1

units) (Arbitrary

expression) protein FGF2 0 18kDa 21kDa 22kDa

-1

Figure 29. FGF2 protein isoform expression in non-ischemic skeletal muscles. Quantitative analysis of LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT (white bar), Fgf2-/- (light gray bar), and FGF2 LMW-only (dark gray bar) non-ischemic muscles.

Data are presented as mean ± SEM. n=4 for each mouse group at each time point. N.S (Not

significant).

169

FGF2 isoform expression (non-ischemic muscle) 8 WT

7 * FGF2 HMW-only 6

5 * 4 3 Arbitrary units 2

FGF2 protein expression FGF2 protein 1

0 18kDa 21kDa 22kDa

Figure 30. FGF2 protein isoform expression in non-ischemic skeletal muscles. Quantitative

analysis of LMW (18kDa) and HMW (21kDa, 22kDa) FGF2 isoform expression in WT (white

bar) and FGF2 HMW-only (black bar) non-ischemic muscles. Data are presented as mean ±

SEM. n=4 for each mouse group at each time point. *p<0.05 vs. WT for each HMW isoform.

170

Loss of HMW FGF2 or ischemia did not affect Fgf2 transcription

To assess the effect of hindlimb ischemia on the expression of Fgf2, calf and thigh skeletal muscles were isolated from sham-operated and ischemic limbs at different time points of ischemia.

RT-PCR was used to determine Fgf2 transcript levels at 3, 7 and 14 days after the ischemia surgery.

Fgf2 mRNA in Fgf2-/- and FGF2 LMW-only muscles was analyzed relative to WT expression

(Figure 31). As expected, no Fgf2 mRNA was detected in Fgf2-/- limbs. In FGF2 LMW-only

muscles, mRNA expression in ischemic limbs was unchanged compared to WT at day 3, elevated

at day 7 and returning to normal at day 14 of ischemia. However, ischemic transcript levels, at any

time point of ischemia, were not different from the sham cohorts.

Expression of only the HMW isoforms increases Fgf2 mRNA during ischemia

In the presence of only HMW FGF2, Fgf2 mRNA was significantly increased in response to ischemia. Fgf2 transcript levels were increased in ischemic FGF2 HMW-only limbs (compared to sham cohorts or WT). Fgf2 mRNA expression was elevated from non-ischemic levels at day 3 and day 7 of ischemia and returned to normal levels at 14 days (p<0.05, Figure 32). Overall, there appears to be a FGF2 protein (HMW)-dependent upregulation of Fgf2 expression that occurs in

the absence of LMW FGF2 and the presence of ischemia.

171

Fgf2-/- SHAM Fgf2 mRNA Expression (ischemic muscle) Fgf2 -/- HLI FGF2 LMW-only SHAM 10 FGF2 LMW-only HLI WT

8

6

4 (Arbitrary Units)

2 mRNA Expression mormalized to WT to mormalized Expression mRNA 0 3d 7d 14d

Figure 31. Fgf2 mRNA expression in sham and ischemic skeletal muscles. Time course of

Fgf2 mRNA expression relative to WT. Quantitative RT-PCR analysis was performed on sham (SHAM) and ischemic (HLI) muscles at days 3, 7, and 14 of hindlimb ischemia. For each time point of ischemia, sham (open) or ischemic (filled) fold-changes in Fgf2-/- (square) or FGF2 LMW-only (circle) muscles are expressed as a ratio of the corresponding WT (solid line) transcript levels. Data are presented as mean ± SEM. n=4 for each mouse group at each time point.

172

Fgf2 mRNA Expression (ischemic muscle)

# *

# *

*

Figure 32. Fgf2 mRNA expression in sham and ischemic skeletal muscles. Time course of

Fgf2 mRNA expression relative to WT. Quantitative RT-PCR analysis was performed on

sham (SHAM) and ischemic (HLI) muscles at days 3, 7, and 14 of hindlimb ischemia. For

each time point of ischemia, sham (open) or ischemic (filled) fold-changes in FGF2 HMW- only (circle) are expressed as a ratio of the corresponding WT (solid line) transcript levels.

Data are presented as mean ± SEM. n=4 for each mouse group at each time point. # p<0.05 vs. SHAM cohort, *p<0.05 vs. WT HLI.

173

Ablation of HMW FGF2 did not affect LMW isoform expression during ischemia

To elucidate the effect of hindlimb ischemia on the expression of FGF2 protein isoforms,

calf and thigh skeletal muscles were evaluated with Western immunoblotting at 3, 7 and 14 days

of ischemia. There was no expression of the LMW 18kDa isoform in Fgf2-/- muscles. In ischemic

WT tissues, the 18 kDa protein level was not different from sham expression at day 3, day 7 or day 14 (Figures 33A and 34A). A similar pattern was also observed in FGF2 LMW-only limbs.

This suggests the absence of FGF2 LMW-dependent regulation of FGF2 translation during

ischemia.

A. 18kDa isoform expression in ischemic muscle

4.5 4 3.5 3 2.5 2

Arbitrary Units Units Arbitrary 1.5 1 0.5 0 -0.5 3d 7d 14d

174

B. 21kDa isoform expression in ischemic muscle 3 2.5 2 1.5 1 0.5 Arbitrary Units 0 -0.5 -1 3d 7d 14d

C. 22kDa isoform expression in ischemic muscle

2

1.5

1

0.5

Arbitrary Units 0

-0.5

-1 3d 7d 14d

Figure 33. FGF2 protein isoform expression in sham and ischemic skeletal muscle.

Quantitative analysis of 18kDa (A), 21kDa (B) and 22kDa (C) FGF2 expression in WT,

Fgf2-/-, and FGF2 LMW-only sham (dotted line) and ischemic (solid line) muscles at days 3,

7, and 14 of hindlimb ischemia. Data are presented as mean ± SEM. n=4 for each mouse group at each time point.

175

Loss of LMW FGF2 altered HMW FGF2 protein expression

The expression of the HMW isoforms in the absence of LMW FGF2 was also examined

during ischemic injury. The trend of 21kDa or 22kDa FGF2 expression in WT muscles mirrored

that of the 18kDa isoform with increased expression of either protein at day 3 and continued

elevation at day 14 (Figure 33-34B and C). In FGF2 HMW-only limbs, isoform expression at 3 days of ischemia had decreased relative to non-ischemic HMW expression (data not shown).

Expression had begun to rise by day 7, remained increased at day 14 to return to baseline (non- ischemic) levels. Despite the reduction in HMW FGF2, expression was still slightly increased relative to WT (for 21/22 kDa, Figure 34B and C). This could indicate that ischemia does not affect

FGF2 protein isoform expression but the absence of one class of FGF2 protein (LMW) is compensated for by increased expression of HMW FGF2 at baseline.

A. 18kDa isoform WT SHAM

10 WT HLI

8

6

4 Arbitrary Units 2

0 3d 7d 14d

176

WT SHAM

21kDa isoform WT HLI B. FGF2 HMW-only SHAM 7 FGF2 HMW-only HLI 6 5

4 * 3

Arbitrary Units 2 1 0 3d 7d 14d C. 22kDa isoform

4

3 *

2

Arbitrary Units 1

0 3d 7d 14d

Figure 34. FGF2 protein isoform expression in sham and ischemic skeletal muscle.

Quantitative analysis of 18kDa (A) 21kDa (B) and 22kDa (C) FGF2 expression in WT and

FGF2 HMW-only sham (dotted line) and ischemic (solid line) muscles at days 3, 7, and 14 of

hindlimb ischemia. Data are presented as mean ± SEM. n=4 for each mouse group at each

time point. *p<0.05 vs. WT HLI.

177

Role of FGF2 protein isoforms in FGFR expression

In response to ischemia-reperfusion injury in the heart, the LMW isoform of FGF2 protects against cardiac dysfunction via activation of the FGFR1 receptor tyrosine kinase (Liao et al.,

2010). Additionally, overexpression of HMW FGF2 depresses basal FGFR1 activation (Liao et al., 2010). Yet, it is unknown if ablation of LMW FGF2 or HMW FGF2 alters the expression profile of FGF receptors in skeletal muscle basally or under stress such as ischemia. To determine the gene expression levels of the murine FGF receptors, RNA was isolated from non-ischemic and ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only limbs. First strand cDNA was amplified by quantitative real-time PCR (qRT-PCR) using primers for the seven FGF receptors,

Fgfr1b, -1c, -2b, -2c, -3b, -3c, and -4 and an internal control gene (18S).

HMW FGF2 expression influences Fgfr1 gene expression at baseline and during ischemia

No Fgfr2b or Fgfr2c mRNA was detected in skeletal muscle. Fgfr1b and Fgfr1c mRNA expression were both unchanged in Fgf2-/- or FGF2 LMW-only mice muscles relative to non- ischemic WTs (Figure 35A). In FGF2 HMW-only limbs, Fgfr1b transcript levels were significantly decreased (p<0.05 vs. WT) while Fgfr1c expression was unchanged.

To characterize the effect of FGF2 isoform and ischemia on Fgfr1 expression, qRT-PCR was used to determine Fgfr1 mRNA expression at 1, 3, 7 and 14 days after the induction of ischemia. Transcript levels are presented as a ratio of sham muscles to address any changes related to the hindlimb ischemia surgery. In all groups, Fgfr1b expression increased with time, was highest after 7 days and at a decline by 14 days of ischemic (Figure 35B). At 3 days, Fgfr1b and Fgfr1c had a trend towards or was significantly increased in Fgf2-/- and FGF2 HMW-only muscles

178 relative to WT (p<0.05 vs. WT; Figures 35B and C). Fgfr1b and Fgfr1c mRNA had a similar expression profile in FGF2 LMW-only tissues relative to WT.

179

Fgfr1 expression in non-ischemic muscle A. 2.5 WT Fgf2-/- 2 FGF2 LMW-only FGF2 HMW-only

1.5

1 (Arbitrary Units)

# $‡ 0.5 mRNA Expression mormalized WT to mRNA

0 Fgfr1b Fgfr1c

B. Fgfr1b expression (ischemic muscle) 5 WT Fgf2-/- 4 FGF2 LMW-only FGF2 HMW-only

3

2 (Arbitrary units)

1 mRNA Expression relative to sham to relative Expression mRNA

0 1 3 7 14 Days after ischemia surgery 180

Fgfr1c expression (ischemic muscle) C. 5 WT Fgf2-/- FGF2 LMW-only 4 FGF2 HMW-only

3 # ‡

2 # ‡ (arbitrary units) (arbitrary

1

sham to relative Expression mRNA

0 1 3 7 14 Days after ischemia surgery

Figure 35. Fgfr1 mRNA expression in non-ischemic, sham and ischemic skeletal muscles. (A)

Quantitative RT-PCR analysis for Fgfr1b and 1c gene expression in non-ischemic skeletal

muscle relative to WT (white bar). Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar)

or FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles. Time course of Fgfr1b (B) and Fgfr1c (C) mRNA

expression. For each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA

expression in ischemic muscles are expressed relative to corresponding sham levels in WT

(blue), Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are

presented as mean ± SEM. n=4 for each mouse group at each timepoint. #p<0.05 vs. WT,

$p<0.05 vs. Fgf2-/-, ‡p<0.05 vs. FGF2 LMW-only, †p<0.05 vs. FGF2 HMW-only.

181

Alterations in FGF2 protein expression did not affect Fgfr3 transcription

Fgfr3 expression in non-ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only

skeletal muscles was also examined. Fgfr3b and Fgfr3c mRNA expression were both unchanged

in Fgf2-/- mice relative to non-ischemic WTs (Figure 36A). FGF2 LMW-only and FGF2 HMW-

only muscles also had similar levels of Fgfr3b or Fgfr3c mRNA as WTs.

To describe the effect of specific FGF2 isoform expression and hindlimb ischemia on Fgfr3

transcript levels, sham and ischemic muscles were isolated and analyzed for Fgfr3b and Fgfr3c

mRNA expression. In WT, FGF2 LMW-only, and FGF2 HMW-only limbs, no changes in expression of either transcript were detected at 1, 3, 7 or 14 days of ischemia (Figures 36B and C).

While the Fgf2-/- tissues had increased Fgfr3b expression at 7 days compared to WT, this change was not significant (Figure 36B). In all groups, ischemic Fgfr3c expression increased with time, was highest at 7 days and returned to normal at 14 days (Figure 36C).

182

Fgfr3 expression in non-ischemic muscle

A. 3 WT Fgf2-/- 2.5 FGF2 LMW-only FGF2 HMW-only 2

1.5 (Arbitrary Units)

1 mRNA Expression mormalized WT to mRNA 0.5

0 Fgfr3b Fgfr3c

B. Fgfr3b expression (ischemic muscle) 6 WT Fgf2-/- 5 FGF2 LMW-only FGF2 HMW-only 4

3

2 (arbitrary units) (arbitrary

1 mRNA Expression relative sham to relative Expression mRNA

0 1 3 7 14 183 Days after ischemia surgery Fgfr3c expression (ischemic muscle) C. 7 WT Fgf2-/- 6 FGF2 LMW-only

5 FGF2 HMW-only

4

3 (arbitrary units) (arbitrary 2

mRNA Expression relative sham to relative Expression mRNA 1

0 1 3 7 14 Days after ischemia surgery

Figure 36. Fgfr3 mRNA expression in non-ischemic, sham and ischemic skeletal muscles. (A)

Quantitative RT-PCR analysis for Fgfr3b and 3c gene expression in non-ischemic skeletal

muscle relative to WT (white bar). Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar)

or FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles. Time course of Fgfr3b (B) and Fgfr3c (C) mRNA

expression. For each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA

expression in ischemic muscles are expressed relative to corresponding sham expression

levels in WT (blue), Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black)

limbs. Data are presented as mean ± SEM. n=4 for each mouse group at each timepoint.

184

Expression of only LMW FGF2 affects Fgfr4 expression during ischemia

Fgfr4 expression in non-ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only skeletal muscles was also examined. For all mouse groups, Fgfr4 mRNA expression was not different from WT (Figure 37A).

To determine the role of FGF2 isoform expression and ischemia on Fgfr4 gene expression, transcript levels were measured at 1, 3, 7 or 14 days of ischemia. FGF2 LMW-only muscles had increased levels at 7 days relative to WTs (p < 0.05, Figure 37B) and returned to normal expression at 14 days. Fgf2-/- and FGF2 HMW-only also trended towards an elevated Fgfr4 expression at 7 days. Both groups had an Fgfr4 expression profile that increased slightly with time, was highest at 7 days and returned to normal at 14 days (Figure 37B).

185

Fgfr4 expression (non-ischemic muscle) A. 2.5 WT Fgf2-/- 2 FGF2 LMW-only FGF2 HMW-only

1.5

1 (arbitrary units) (arbitrary

# ‡† Relative mRNA expression expression mRNA Relative 0.5

0

B. Fgfr4 expression (ischemic muscle) 6 WT Fgf2-/- 5 FGF2 LMW-only FGF2 HMW-only 4 #

3 †

* 2 (arbitrary units) (arbitrary

1 mRNA Expression relative sham to relative Expression mRNA

0 1 3 7 14 186 Days after ischemia surgery Figure 37. Fgfr4 mRNA expression in non-ischemic, sham and ischemic skeletal muscles. (A)

Quantitative RT-PCR analysis for Fgfr4 gene expression in non-ischemic skeletal muscle relative to WT (white bar). Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or

FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes relative to WT non-ischemic muscles. (B) Time course of Fgfr4 mRNA expression. For each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic muscles are expressed relative to corresponding sham levels in WT (blue), Fgf2-/- (red),

FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are presented as mean ±

SEM. n=4 for each mouse group at each timepoint. *p<0.05 vs. Day 1 cohort, #p<0.05 vs.

WT, †p<0.05 vs. FGF2 HMW-only.

187

Basal FGFR activation is not affected by selective FGF2 isoform expression

The FGFRs are typical receptor tyrosine kinases whose activation in response to ligand binding and receptor dimerization is driven by a sequence of tyrosine autophosphorylation and results in intracellular signaling and response (Eswarakumar et al., 2005). To determine whether the trends in Fgfr1, Fgfr3 or Fgfr4 gene expression observed in response to selective deletion of

LMW or HMW FGF2 affected FGFR expression and/or activation, Western blotting was performed on non-ischemic skeletal muscles, sham and ischemic muscles.

As there are currently no specific antibodies for the “b” or “c” forms of the FGFRs, immunoblotting results for FGFR1 actually represents activation and expression of FGFR1b and

FGFR1c protein translation products (Figures 38-41). The level of activation of FGFR1, 3 or 4 was expressed as the ratio of the amount of each phosphorylated receptor to total level of the receptor. In line with the qRT-PCR analysis, FGFR2 expression or activation was not observed in non-ischemic skeletal muscles. Neither deletion of the Fgf2 gene (Fgf2-/-) nor selective ablation of FGF2 isoforms (FGF2 LMW-only or FGF2 HMW-only) affected the level of FGFR1 activation

(Figure 38A). Similarly, FGFR3 and FGFR4 activation were unchanged with any alterations of

FGF2 expression.

Total protein expression of the FGFR1, FGFR3 and FGFR4 in non-ischemic skeletal muscles was not different between the mouse groups (Figure 38B). Overall, manipulation of FGF2 gene or protein isoform expression had no effect on FGF receptor protein or activation under baseline conditions.

188

FGFR Activation

A. WT 3 Fgf2-/- FGF2 LMW-only FGF2 HMW-only 2 FGFR/total FGFR - 1 phospho

0 FGFR1 FGFR3 FGFR4

30 B.

25

20

15

10 (Arbitrary Unit)

Total FGFR expression FGFR expression Total 5

0 FGFR1 FGFR3 FGFR4

Figure 38. FGFR protein expression and activation (phosphorylation) in non-ischemic

skeletal muscle. Quantitative analysis of FGFR1, FGFR3 and FGFR4 activation (A) and total

expression (B) in WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar),

and FGF2 HMW-only (black bars) non-ischemic muscles. Data are presented as mean ±

SEM. n=4 for each mouse group.

189

FGFR1 activation is elevated during ischemic injury in the absence of all FGF2 expression

LMW FGF2 protects against cardiac dysfunction following acute ischemia-reperfusion

injury. This response is non-mitogenic and dependent on activation of the FGF receptor, FGFR1

(Liao et al., 2010). To determine if the vascular response to endogenous LMW or HMW FGF2

expression during chronic ischemia was associated with changes in FGFR activation and/or

expression, WT, Fgf2-/-, FGF2 LMW-only or FGF2 HMW-only mice were subjected to hindlimb

ischemia for 3 or 7 days. Skeletal muscles were isolated from ischemic limbs (HLI) as well as the

sham-operated, contralateral limbs (SHAM). To address variations in protein expression related only to the ischemia surgery procedure, receptor activation and expression in ischemic limbs were presented as a ratio of the corresponding sham limb expression.

In the presence of only HMW FGF2, activation (phosphorylation) of FGFR1 was unchanged from WT after 3 days of ischemia (Figure 39A). Expression (total protein) of the

receptor was also not different at this timepoint (Figure 39B). FGF2 LMW-only or Fgf2-/- muscles also had a similar pattern of FGFR1 activation and expression.

A higher degree of receptor phosphorylation was observed in Fgf2-/- muscles at day 7 of ischemia (p<0.05 vs. WT, Figure 39A). Conversely, expression of only the LMW or only the

HMW isoforms of FGF2 did not affect FGFR activation after a week of hindlimb ischemia. FGFR1 protein expression was also similar to WT in isoform-specific knock out skeletal muscles.

190

FGFR1 activation (ischemic muscle)

A. 4 WT Fgf2-/- # ‡ † FGF2 LMW-only 3 FGF2 HMW-only SHAM/totalSHAM) - 2

1 * FGFR activation relative to sham HLI/total HLI)/(p HLI/total -

(p 0 3d 7d

FGFR1 expression (ischemic muscle) B. 2

1.5

1

(total HLI/total SHAM) SHAM) (total HLI/total 0.5 FGFR expression relative to sham relative expression FGFR

0 3d 7d Figure 39. FGFR1 activation (phosphorylation) and expression at 3 or 7 days of ischemia.

Phosphorylation (ratio vs. total protein) (A) and expression (B) in ischemic WT, Fgf2-/-,

FGF2 LMW-only, and FGF2 HMW-only muscles at relative to sham tissues. Data are presented as mean ± SEM. n=4 for each mouse group. *p<0.05 vs. Day 7 cohort, #p<0.05 vs.

WT, ‡p<0.05 vs. FGF2 LMW-only, †p<0.05 vs. FGF2 HMW-only.

191

Ablation of FGF2 and hindlimb ischemia does not affect FGFR3 activation

The degree of FGFR3 activation and protein expression in WT, Fgf2-/-, FGF2 LMW-only

and FGF2 HMW-only skeletal muscles was also examined at 3 or 7 days of ischemia. While the

complete loss of FGF2 expression (Fgf2-/-) produced a trend towards decreased phosphorylation

of FGFR3 at day 3, this change was not significant (Figure 40A). Total protein expression of

FGFR3 was increased in the Fgf2-/- muscles (p<0.05, Figure 40B). In the presence of only LMW

or HMW FGF2, the level of FGFR3 activation at 3 days of ischemia was unchanged. Similarly,

protein expression of the receptor was not different from WT in the isoform knockout mice.

Activation of FGFR3 at 7 days of ischemia in Fgf2-/-, FGF2 LMW-only and FGF2 HMW- only limbs was not significantly different from WT (Figure 40A). Likewise, total protein expression of the receptor was similar among the mouse groups (Figure 40B).

192

FGFR3 activation (ischemic muscle)

3 WT Fgf2-/- FGF2 LMW-only FGF2 HMW-only 2 SHAM/totalSHAM) -

1 FGFR activation relative to sham HLI/total HLI)/(p HLI/total -

(p 0 3d 7d

FGFR3 expression (ischemic muscle) 4

# ‡ † 3

2

(total HLI/total SHAM) 1 FGFR expression relative shamto relative expression FGFR

0 3d 7d

Figure 40. FGFR3 activation (phosphorylation) and expression at 3 or 7 days of ischemia.

Phosphorylation (ratio vs. total protein) (A) and total protein expression (B) in ischemic WT,

Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only muscles at relative to sham tissues. Data are presented as mean ± SEM. n=4 for each mouse group. #p<0.05 vs. WT, ‡p<0.05 vs. FGF2

LMW-only, †p<0.05 vs. FGF2 HMW-only.

193

Activation of FGFR4 does not influence FGF2-mediated response to chronic ischemia

To elucidate the effect of hindlimb ischemia and selective FGF2 expression on the

activation of FGFR4, expression of all and only phosphorylated forms of the receptor were

evaluated with Western immunoblotting at 3 and 7 days of ischemia (Figure 41A). In FGF2 LMW-

only or FGF2 HMW-only muscles, FGFR4 activation was not changed from WT at 3 days of

ischemia. FGFR4 protein expression was not different from WT expression (Figure 41B). Fgf2-/-

limbs followed a similar pattern of FGFR4 activation and expression.

By 7 days of ischemia, activation of FGFR4 in Fgf2-/-, FGF2 LMW-only and FGF2 HMW-

only limbs was unchanged relative to WT tissues (Figure 41A). Correspondingly, total protein

expression of the receptor was similar among the mouse groups (Figure 41B).

The absence of changes in FGFR activation and expression in these results confirm

previous data from our laboratory where FGFR1 protein levels were unchanged in non-ischemic

FGF2 HMWKO hearts but contradict data from mice overexpressing HMW FGF2 where basal

FGFR1 activation was depressed (Liao et al., 2010; Manning et al., 2013). Furthermore, FGFR1 activation was similarly activated by LMW FGF2 or HMW FGF2 in skeletal muscles but is upregulated in Fgf2-/- limbs. The elevated receptor activation in Fgf2-/- muscles is most likely an attempt to increased FGF2 signaling to compensate for the loss of the ligand expression. These findings suggest that there is an FGF2-mediated regulation of FGFR in skeletal muscle that is independent of the class of FGF2 protein isoform.

194

FGFR4 activation (ischemic muscle)

A. WT 3 Fgf2-/- FGF2 LMW-only FGF2 HMW-only

2 SHAM/totalSHAM) -

1 FGFR activation relative to sham HLI/total HLI)/(p HLI/total - (p 0 3d 7d

B. FGFR4 expression (ischemic muscle) 2

1 (total HLI/total SHAM) FGFR expression relative shamto relative expression FGFR

0 3d 7d Figure 41. FGFR4 activation (phosphorylation) and expression at 3 and 7 days of ischemia.

Phosphorylation (ratio vs. total protein) (A) and expression (B) in ischemic WT, Fgf2-/-,

FGF2 LMW-only, and FGF2 HMW-only muscles at relative to sham tissues. Data are

presented as mean ± SEM. n=4 for each mouse group.

195

Effect of Fgf2 gene or protein isoform deletion on angiogenesis-related protein expression

FGF2 signaling plays a role in the regulation of normal and adaptive vascular development

(Annex & Simons, 2005; Potente, Gerhardt, & Carmeliet, 2011; Presta et al., 2005). There is also recent evidence to suggest that, in addition to its own direct effect on vascular growth, FGF2 has an indirect role in neo-/re-vascularization by promoting the expression of other angiogenic growth factors, cytokines adhesion molecules, and chemokines (Murakami & Simons, 2008; Presta et al.,

2009; X. Yang et al., 2015). Furthermore, the different autocrine, paracrine and intracrine signaling pathways mediated by the FGF2 isoforms suggests that in response to ischemic injury, HMW

FGF2 might induce the expression of a repertoire of angiogenesis-related proteins distinct from that of LMW FGF2 and vice-versa.

To identify the expression profile of angiogenic factors, cytokines and chemokines regulated by the individual FGF2 isoforms in ischemic-induced revascularization of the mouse hindlimb, the level of several angiogenesis-related proteins was determined using a proteome antibody array. The array containing antibodies against 53 different proteins spotted in duplicates was hybridized with equal amounts of protein isolated from sham and ischemic skeletal muscles after 7 days of ischemia. This was chosen because the ischemic-induced upregulation of FGF2 isoform expression is maximal 7 days after surgery as previously described in Chapter 1 of this dissertation. The array dots were quantified by densitometric analysis and the ischemic expression of each protein (after compensation for differences in background intensity) was normalized to its expression in the contralateral (sham) limb as presented in Table 13.

A majority of the dots (antibodies) showed similar intensities between WT, Fgf2-/-, FGF2

LMW-only and FGF2 HMW-only ischemic muscles but changes were observed in 7 different antibodies. Among these was MMP-9, a proteolytic enzyme involved in the remodeling of the

196

basement membrane and extracellular matrix (ECM) during neovascularization to allow for

outward migration of ECs and vSMCs (Q. Chen et al., 2013; P.-H. Huang et al., 2009). In the

absence of any FGF2 expression (Fgf2-/-), MMP-9 expression was significantly increased compared to WT after 7 days of ischemia (p<0.05). Serpin E1 (PAI-1), a serine protease inhibitor, was also elevated in Fgf2-/- ischemic muscles (p<0.05 vs. WT). PAI-1 is the principal inhibitor of the urokinase-type plasminogen activator (uPA) system. u-PA catalyzes the conversion of plasminogen to plasmin, another serine protease whose proteolytic activity during vascular growth mirrors that of MMP9 (Medcalf, 2007; Van De Craen, Declerck, & Gils, 2012). Plasmin can also convert inactive MMPs (pro-MMPs) zymogen into their active forms (Q. Chen et al., 2013).

When only HMW FGF2 is expressed, insulin-like growth factor binding protein-3 (IGFBP-

3) protein expression was significantly increased compared to WT (p<0.05). Classically, IGFBPs

are members of the endocrine IGF signaling pathway, where they serve as carriers of IGF and

regulate the turn-over of circulating IGFs (Baxter, 2013; Yamada & Lee, 2009). However, IGF- independent activities on cell growth have also been attributed to these proteins. Expression of

IGFBP-3, in particular has an inhibitory effect on cell proliferation and tumor growth (Oh, Gucev,

Ng, Müller, & Rosenfeld, 1995; Yamada & Lee, 2009). IGFBP-3 can also play a more ambiguous role in cell proliferation. When endogenous IGFBP-3 expression is reduced, EGF-induced proliferation of breast epithelial or cancer cells is also decreased (Baxter, 2015; Butt et al., 2004;

McIntosh et al., 2010). EGF, epidermal growth factor is the prototypic member of the EGFR family characterized by promotion of epidermal cell regeneration and healing of acute or chronic dermal wounds. (Bhora et al., 1995; Peplow & Chatterjee, 2013; Seeger & Paller, 2015). EGF-stimulated wound healing involves the proliferation and migration of keratinocytes and fibroblasts (Haase,

197

Evans, Pofahl, & Watt, 2003; Seeger & Paller, 2015). In Fgf2-/- muscles, EGF was significantly

decreased relative to FGF2 LMW-only and trended towards lower expression compared to WT.

FGF7, another member of the FGF family is associated with wound healing. It is

exclusively expressed by mesenchymal cells and regulates wound healing via proliferation and

migration of epithelial cells (keratinocytes) and activation of the alternatively spliced variant of

the FGF receptor, FGFR2IIIb (Maretzky et al., 2011; Niu et al., 2007; Tsuboi et al., 1993; Yen,

Thao, & Thuoc, 2014). FGF7 had increased expression in FGF2 HMW-only muscles that trended towards significance.

Another IGF binding protein, IGFBP-10 (CCN1) was also analyzed using the antibody array. There was a trend towards increased of expression CCN1 in FGF2 HMW-only ischemic

muscles. IGFBP-10 (CCN1) is not a classical member of the IGF family but is categorized as an

IGFBP based on shared structural homology with IGFBPs1-7 (Malik, Liszewska, & Jaworski,

2015). CCN1 is a secreted matricellular protein of the CCN family that exist at the border of cells and the ECM (Jun & Lau, 2011). While not involved in structural integrity of the ECM, they modulate surface receptors (integrins and HSPGs) to promote pro-angiogenic activity, vascular cell adhesion, migration and proliferation (C.-C. Chen & Lau, 2009; Fataccioli et al., 2002; C. G.

Lin et al., 2003). CCNs can also function as transcription factors under hypoxia conditions to regulate expression of other angiogenic factors including VEGFs and FGFs (Kubota & Takigawa,

2007; Lau, 2011; Malik et al., 2015).

An antibody to a protein (CX3CL1/fractalkine) with both angiogenic and inflammatory roles was also detected on the array. CX3CL1, the sole chemokine (chemotactic ) in the

CX3C subfamily of chemokines is expressed primarily by vascular cells and macrophages and signals through its receptor, CX3CR1 to induce EC proliferation, migration and tube formation

198

(Imai et al., 1997; S.-J. Lee et al., 2006; Shireman, 2007). In addition, the CX3CL1/CX3CR1

interaction mediates the chemotaxis and survival of inflammatory leukocytes such as monocytes/

macrophages (Owen & Mohamadzadeh, 2013; White & Greaves, 2009). There was a trend

towards increased expression of CX3CL1 in FGF2 HMW-only ischemic muscles (p=0.06 vs WT).

Overall, the proteome profile in ischemic skeletal muscle suggests that complete loss of

FGF2 expression (in Fgf2-/- limbs) stimulates increased expression of proteins involved in ECM

remodeling and activation. These ECM remodelers might serve to promote the release and

trafficking of other growth factors that are typically bound to the structural proteins that make up

the ECM including proteoglycans and fibronectin.

Functional Group Fold change in pixel intensity

Growth Factors WT Fgf2-/- FGF2 LMW-only FGF2 HMW-only FGF1 0.94 ± 0.13 0.88 ± 0.09 0.85 ± 0.16 1.09 ± 0.24 FGF2 0.84 ± 0.07 0.89 ± 0.09 0.93 ± 0.16 0.87 ± 0.28 FGF7 0.87 ± 0.06 0.91 ± 0.04 0.99 ± 0.11 1.15 ± 0.04 VEGF-A 1.08 ± 0.12 0.83 ± 0.13 0.99 ± 0.08 1.33 ± 0.41 VEGF-B 0.92 ± 0.05 1.13 ± 0.11 1.02 ± 0.12 1.29 ± 0.40 PIGF-2 0.85 ± 0.08 1.05 ± 0.08 1.10 ± 0.13 0.95 ± 0.13 PDGF-AA 0.92 ± 0.09 0.90 ± 0.02 1.04 ± 0.17 1.17 ± 0.09 PDGF-BB 0.91 ± 0.06 0.98 ± 0.14 1.11 ± 0.18 1.11 ± 0.20 HGF 0.99 ± 0.19 1.12 ± 0.06 1.29 ± 0.24 1.06 ± 0.20 EGF 0.87 ± 0.03 0.80 ± 0.07 1.07 ± 0.03 0.98 ± 0.08 HB-EGF 0.84 ± 0.03 0.99 ± 0.10 1.05 ± 0.09 1.05 ± 0.22

199

PD-ECGF 0.90 ± 0.19 1.17 ± 0.06 1.00 ± 0.07 1.08 ± 0.01 Angiogenesis Angiogenin 1.07 ± 0.12 1.04 ± 0.02 1.29 ± 0.23 1.03 ± 0.17 Angiopoietin-1 0.89 ± 0.16 1.13 ± 0.15 1.38 ± 0.33 1.01 ± 0.39 Angiopoietin-3 0.93 ± 0.07 0.87 ± 0.05 1.11 ± 0.16 0.89 ± 0.16 CCN1/IGFBP-10 0.91 ± 0.12 1.98 ± 0.30 1.61 ± 0.79 2.05 ± 0.42# CCN3/IGFBP-9 0.82 ± 0.27 3.55 ± 0.63 1.46 ± 0.48 2.38 ± 0.71 IGFBP-2 0.86 ± 0.17 2.03 ± 0.25 1.75 ± 0.66 1.03 ± 0.16 Endothelin-1 0.86 ± 0.03 1.03 ± 0.05 1.10 ± 0.17 1.11 ± 0.32 Leptin/OB 0.90 ± 0.04 0.86 ± 0.05 1.09 ± 0.19 1.01 ± 0.22 Osteopontin 1.18 ± 0.45 1.69 ± 0.34 1.15 ± 0.35 1.13 ± 0.59 Prolactin 0.94 ± 0.12 1.01 ± 0.06 0.99 ± 0.11 1.26 ± 0.41 Proliferin 0.75 ± 0.10 1.05 ± 0.02 0.90 ± 0.05 1.38 ± 0.44 Angiostatic DLL4 0.89 ± 0.10 0.99 ± 0.05 1.17 ± 0.16 1.03 ± 0.07 DPPIV/CD26 1.06 ± 0.34 0.98 ± 0.18 1.09 ± 0.20 1.11 ± 0.31 Endostatin 0.95 ± 0.08 1.02 ± 0.12 0.96 ± 0.15 1.09 ± 0.13 IGFBP-1 0.84 ± 0.10 1.02 ± 0.10 1.11 ± 0.07 0.86 ± 0.22 IGFBP-3 1.19 ± 0.39 3.25 ± 0.34 2.29 ± 0.55 3.9 ± 0.86 # TIMP-1 0.76 ± 0.15 2.97 ± 1.31 2.03 ± 0.92 2.22 ± 0.96 TIMP-4 0.86 ± 0.07 0.91 ± 0.04 0.92 ± 0.13 1.01 ± 0.19 Pentraxin-3 0.80 ± 0.32 1.97 ± 0.18 1.80 ± 0.77 2.84 ± 1.31 Serpin F1/PEDF 1.02 ± 0.19 1.35 ± 0.17 1.29 ± 0.27 1.44 ± 0.26 Thrombospondin-2 0.78 ± 0.17 1.63 ± 0.21 1.37 ± 0.35 1.5 ± 0.41 ECM Remodeling MMP-3 1.10 ± 0.28 2.43 ± 0.38 2.29 ± 0.84 1.74 ± 0.56 MMP-8 0.84 ± 0.05 1.19 ± 0.03 1.07 ± 0.14 0.92 ± 0.21 MMP-9 0.69 ± 0.03 4.471 ± 0.57 #‡† 0.88 ± 0.25 1.92 ± 0.74 ADAMTS1 0.92 ± 0.07 0.93 ± 0.05 1.12 ± 0.23 1.17 ± 0.06 Serpin E1/PAI-1 0.68 ± 0.23 6.43 ± 0.56 #‡† 2.64 ± 0.93 1.38 ± 0.52

200

Chemokines MCP-1 0.56 ± 0.17 2.23 ± 0.74 1.38 ± 0.34 1.25 ± 0.29 MIP-1α 0.92 ± 0.12 1.05 ± 0.05 1.12 ± 0.11 0.87 ± 0.16 CXCL1 0.82 ± 0.08 0.90 ± 0.11 1.07 ± 0.16 0.98 ± 0.05 CXCL4/PF-4 0.92 ± 0.10 0.98 ± 0.05 1.09 ± 0.14 1.02 ± 0.08 CX3CL1 0.83 ± 0.04 1.10 ± 0.15 1.05 ± 0.09 1.26 ± 0.08 (Fractalkine) CXCL16 0.83 ± 0.11 1.93 ± 0.39 2.13 ± 0.88 2.25 ± 0.96 CXCL10 0.78 ± 0.07 0.99 ± 0.11 1.34 ± 0.31 1.07 ± 0.03 CXCL12/SDF-1 0.96 ± 0.26 1.81 ± 0.17 1.77 ± 0.20 2.62 ± 0.98 GM-CSF 0.94 ± 0.12 0.87 ± 0.08 0.93 ± 0.04 0.98 ± 0.08 Inflammation IL-1α 0.84 ± 0.12 1.04 ± 0.08 0.82 ± 0.07 1.03 ± 0.17 IL-10 (anti-) 0.89 ± 0.18 1.12 ± 0.05 1.19 ± 0.14 1.12 ± 0.10 Amphiregulin 0.97 ± 0.02 0.84 ± 0.06 0.95 ± 0.09 1.03 ± 0.17 Tissue Factor 0.89 ± 0.22 1.50 ± 0.25 1.51 ± 0.35 1.45 ± 0.48 Endoglin/CD105 0.86 ± 0.16 1.36 ± 0.05 1.17 ± 0.19 1.15 ± 0.41

Table 13: Expression levels of various angiogenic-related proteins were determined in WT,

Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only muscles at 7 days of ischemia using a

protein antibody array. For each group, the fold change in ischemic muscle protein

expression is presented as a ratio of the contralateral (sham) limb. Data are presented as

mean ± SEM. n=4 for each mouse group. Fold changes that appear in red indicate a trend

towards or significant difference. Statistical significance is marked by #p<0.05 vs. WT,

‡p<0.05 vs. FGF2 LMW-only, †p<0.05 vs. FGF2 HMW-only.

201

HMW FGF2 promotes neutrophil infiltration during ischemia

In addition to vascular cells, ischemia-induced revascularization also triggers the activation and mobilization of bone marrow derived-mononuclear cells (BM-MNCs) through increased expression of chemokines and adhesion molecules (Frangogiannis et al., 2002; Keeley, Mehrad,

& Strieter, 2008). These cells of myeloid lineage are the first responders and their trafficking starts within minutes following the ischemic injury and can last up to two weeks (Meisner & Price, 2010;

Tidball, 2005). Initially, neutrophils predominate the BM-MNCs that are recruited to the ischemic muscles (Frangogiannis & Entman, 2005; Tidball, 2005). Their numbers have been observed to increase as early as two hours after the injury and are undetectable 7 days later (Hoefer et al., 2005;

Silvestre et al., 2013). Many studies have described a role for inflammatory cell response in FGF2- mediated neovascularization (Andrés et al., 2009; Jih et al., 2001; Leali et al., 2003; Presta et al.,

2007, 2009).

To identify the role of the FGF2 isoforms in the tissue inflammatory response to chronic hindlimb ischemia, neutrophil recruitment in non-ischemic, sham-operated or ischemic calf and thigh was assessed (Figure 42). To that end, neutrophils were identified as myeloperoxidase

(MPO) positive cells. MPO is a peroxidase enzyme and the major component of neutrophil azurophilic granules (Bradley, Priebat, Christensen, & Rothstein, 1982; Oklu, Albadawi, Jones,

Yoo, & Watkins, 2013). Calf (gastrocnemius) and thigh (gracilis, semi-membranosus and semi- tendinosus) muscles were harvested at baseline or 3 days of ischemia from WT, Fgf2-/-, FGF2

LMW-only and FGF2 HMW-only mice. After immunodetection of MPO, no significant differences between genotypes in the number of neutrophils were present in non-ischemic muscles

(Figure 41A). Three days of ischemia showed a significantly higher number of neutrophils in ischemic muscles from Fgf2-/- and FGF2 HMW-only mice relative to their respective sham

202 muscles (p<0.05 vs HLI, Figure 42B). Surprisingly, neutrophil levels in WT and FGF2 LMW-only ischemic muscles were unchanged from their contralateral controls. Furthermore, FGF2 HMW- only muscles and Fgf2-/- had significantly increased neutrophil density relative to WT (p<0.05).

This suggests that while ischemia-induced neutrophil accumulation was inhibited by expression of all the FGF2 isoforms, HMW FGF2 alone can sufficiently mount an inflammatory response.

A. Neutrophil density (non-ischemic muscle)

0.1 WT Fgf2-/- FGF2 LMW-only 0.075 FGF2 HMW-only

0.05

0.025 Number ofcells/field MPO+ Number

0 1

203

Neutrophil density (ischemic muscle) B.

5 WT # * ‡ Fgf2-/- 4 FGF2 LMW-only FGF2 HMW-only 3 # *

2

1 Number ofcells/field MPO+ Number

0 SHAM HLI

Figure 42. Neutrophil density expressed as the number of myeloperoxidase (MPO)-positive cells/field in WT (white bars), Fgf2-/- (light gray bars), FGF2 LMW-only (dark gray bars), and FGF2 HMW-only (black bars) non-ischemic (A) and ischemic (B) skeletal muscle after

3 days of ischemia. Data are presented as mean ratio ± SEM, n= 4. *p<0.05 vs. SHAM cohort,

#p< 0.05 vs. WT, ‡p< 0.05 vs. FGF2 LMW-only.

204

Selective ablation of FGF2 isoform expression does not alter monocyte infiltration

Subsequent to the accumulation of neutrophils at the site of ischemic injury, a different

population BM-MNCs are recruited (Meisner & Price, 2010). These late arriving BM-MNCs (3

days after injury) are predominantly monocytes/macrophages and they further amplify the

inflammatory response (J. Lu, Pompili, & Das, 2013; Nguyen & Tidball, 2003). Neutrophils also

secrete cytokines and chemokines that attract and activate macrophages. The initial neutrophil

inflammatory infiltrate is replaced by macrophages and this is important for the removal of

necrotic tissue and resolution of inflammation (Serhan & Savill, 2005; Silvestre et al., 2008;

Tidball, 2005). Macrophages also secrete a repertoire of pro-angiogenic growth factors (including

FGF2), cytokines and ECM remodeling proteins that propagate the revascularization process

(Fung & Helisch, 2012; Silvestre et al., 2008, 2013). FGF2 exerts a direct chemotactic effect on

monocytes as well as upregulates the expression of other chemoattractants important for the

adhesion and recruitment of monocytes/macrophages (Andrés et al., 2009; Leali et al., 2003; Presta

et al., 2005, 2009).

To further examine the role of the FGF2 isoforms in the inflammatory response to chronic

hindlimb ischemia, macrophage infiltration was quantified in non-ischemic, sham-operated or

ischemic skeletal muscles. Macrophages/monocytes were identified as Mac-3-positive cells

(Koestler, Rieman, Muirhead, Greig, & Poste, 1984; Meisner et al., 2014; Shireman, Contreras-

Shannon, Reyes-Reyna, Robinson, & McManus, 2006). Calf (gastrocnemius) and thigh (gracilis, semi-membranosus and semi-tendinosus) muscles were harvested at baseline, 3 or 7 days of ischemia from WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only mice. No significant differences between genotypes in the density of macrophages were present in non-ischemic muscles (Figure 43A). By 3 days of ischemia, there was significantly increased macrophage

205

infiltration in the ischemic muscles of WT, Fgf2-/-, and FGF2 LMW-only mice compared to their respective sham muscles (p<0.05 vs HLI, Figure 43B). This differed from ischemic FGF2 HMW- only muscles whose macrophages numbers were similar to their sham limb levels.

Monocyte/macrophage infiltration was also evaluated after 7 days of ischemia (Figure

43B). The number of Mac-3-positive cells in the ischemic muscles was significantly increased compared to sham muscles for all mouse groups (p<0.05 vs HLI). Expression of only LMW FGF2 or only HMW FGF2 did not change the level of ischemia-induced macrophage recruitment relative to WT. However, complete absence of FGF2 protein expression (Fgf2-/-) affected the amount of macrophage infiltrate in response to ischemia. At 3 or 7 days of ischemia, Fgf2-/- muscles had higher numbers of macrophages. There was a nearly 3-fold increase in infiltration in the Fgf2-/-

ischemic limbs compared to WT, FGF2 LMW-only or FGF2 HMW-only limbs.

The data presented here suggest that expression of FGF2 protein (either LMW or HMW isoforms) is necessary to induce normal macrophage infiltration during ischemia and loss of total

FGF2 expression results in an exaggerated and potentially harmful inflammatory response.

206

Monocyte density (non-ischemic muscle)

A. 0.5 WT Fgf2-/- 0.4 FGF2 LMW-only FGF2 HMW-only 3+ cells/field - 0.3

0.2

0.1 Number of Mac Number

0 1

B. Monocyte infiltration (ischemic muscle) 20 * #‡ †

15 *# ‡ †

10

* * * 5 * *

Number of Mac - 3+ cells/field Number

0 3d SHAM 3d HLI 7d SHAM 7d HLI

Figure 43. Monocyte density expressed as the number of Mac-3-positive cells/field in WT

(white), Fgf2-/- (light gray), FGF2 LMW-only (dark gray), and FGF2 HMW-only (black)

non-ischemic (A) and ischemic (B) skeletal muscle after 3 or 7 days of ischemia. Data are

presented as mean ratio ± SEM, n = 2-4. *p<0.05 vs. SHAM cohort, #p< 0.05 vs. WT, ‡p<0.05

vs. FGF2 LMW-only, †p<0.05 vs. FGF2 HMW-only.

207

DISCUSSION II

The results presented in this chapter of the dissertation demonstrate that the endogenous

HMW FGF2 isoforms have an important role in protecting skeletal muscle from ischemia-induced damage. This effect of the HMW isoforms is possibly mediated by FGF receptor-independent signaling, leading to gene expression changes in angiogenic pathways (IGFBPs and CX3CL1) and in conjunction with an altered inflammatory cell response (Figure 44).

Since its discovery in the late 1980s, FGF2 has been classified as a “classic angiogenic factor” (Presta et al., 2007; Ribatti, 2014). In vitro experiments showed a mitogenic effect of FGF2 on numerous cell types including endothelial cells, smooth muscle cells, cells, fibroblasts, neuronal cells, and osteoblasts (Bikfalvi et al., 1995; Grothe, Meisinger, Holzschuh,

Wewetzer, & Cattini, 1998; Pasumarthi, Kardami, & Cattini, 1996; Piotrowicz, Maher, & Levin,

1999; Xiao et al., 2003). Moreover, exogenous FGF2 is a potent inducer of proliferation, migration and/or differentiation of endothelial and vascular smooth muscle cells; cellular activities which are critical for vascular growth (Davis et al., 1997; Pintucci et al., 2005; Piotrowicz et al., 2001,

1999). The function of endogenous FGF2 has also been investigated in angiogenesis and other cellular processes. The distinction between the roles of the endogenous or exogenous forms is important in the study of many proteins but is particularly significant when considering FGF2. The existence of several protein isoforms produced from a single Fgf2 gene coupled with discrete cellular localizations of those isoforms prompted the use of the isoform-specific mouse models when deciphering the functions of the classes of FGF2 isoforms.

In the present study, mice with a targeted deletion of the LMW FGF2 or HMW FGF2 were utilized to determine the endogenous roles of the FGF2 isoforms in revascularization of the hindlimb. Characterization of normal skeletal muscle was performed to eliminate the potential of

208

confounding results from altered baseline FGF2 expression. Expression of Fgf2 mRNA and all protein isoforms was present in non-ischemic WT muscles (Figures 28 & 29). This is in accordance with the patterning of FGF expression levels in adult mouse tissues that found moderate expression of Fgf2 mRNA in the structural system (skin, bone and skeletal muscle) (Fon Tacer et al., 2010).

When Fgf2 mRNA was examined in tibiae of HMWKO mice (i.e. FGF2 LMW only), expression was significantly decreased (Homer-Bouthiette et al., 2014). In normal skeletal muscle, however,

Fgf2 mRNA expression is significantly increased when only HMW FGF2 or only LMW FGF2 is present. Ablation of either the LMW or HMW protein isoforms also altered FGF2 protein expression. Similar to the changes in mRNA, loss of HMW or LMW FGF2 expression in skeletal muscles resulted in increased FGF2 protein expression. In the case of global deletion of HMW

FGF2, the LMW isoform was significantly increased and vice versa. This phenotype is partly recapitulated in non-ischemic HMWKO hearts, where the LMW isoform is increased in the absence of the HMW isoforms (Liao et al., 2010). Conversely, in LMWKO (i.e. FGF2 HMW only) hearts, expression of the HMW isoforms (21 or 22kDa) remains unchanged (Liao et al., 2007).

These changes in FGF2 expression suggest that under baseline conditions, there exists a FGF2- mediated regulation of FGF2 expression at the transcriptional and translational levels. Previous work by Jimenez had provided evidence that FGF2 regulates its own promoter in cardiac myocytes

(Jimenez et al., 2004). Furthermore, this regulation is selective based on the FGF2 isoform(s) expressed and in which tissues they are expressed (Coffin et al., 1995). The increased expression of the HMW isoforms takes on more importance because the 18kDa LMW FGF2 isoform is the predominant FGF2 that is expressed in both wildtype and FGF2 transgenic skeletal muscle (Coffin et al., 1995). However, it is important to note that there may be a baseline level of FGF2 protein expression irrespective of isoform identity that is required for cellular functions in skeletal

209

myocytes. The possibility that this is the driver of the increased FGF2 isoform or mRNA

expression in the absence of LMW or HMW FGF2 cannot be excluded.

In light of the changes to baseline levels of Fgf2 and its protein products when a specific class of protein isoform is deleted, the effect of chronic ischemia on FGF2 expression was also addressed. This dissertation demonstrated that induction of chronic ischemia in the mouse hindlimb produced changes in FGF2 transcript or protein expression. The presence of only LMW

FGF2 significantly increased skeletal muscle Fgf2 mRNA compared to wildtype at day 7 of

ischemia (Figure 31). Similarly in the absence of the LMW isoform (only HMW FGF2 present),

Fgf2 mRNA was significantly elevated. This increase persisted over the course of the tissue repair

and remodeling phases (1-14 days) that characterize the physiological response to ischemia in

muscle (Silvestre et al., 2013; Smith et al., 2008; Tidball, 2011). Studies of ischemia in the rabbit

hindlimb have also reported a similar a pattern of FGF2 transcript expression (Bush et al., 1998;

Walgenbach, Gratas, Shestak, & Becker, 1995). Deindl and colleagues however, did not observe

any changes in FGF mRNA (both FGF1 and FGF2) at any time points of ischemia in the same

model.

When only the HMW isoforms were present, FGF2 expression (21 or 22kDa) was also

increased compared to wildtype expression at day 7 of ischemia (Figures 33B, C). However, the

level of FGF2 protein in ischemic muscles did not rise above that of baseline expression (FGF2

HMW-only non-ischemic) muscles and was similar to expression in sham tissues. This suggests that while HMW FGF2 isoform expression might be changed when the LMW isoform is ablated; there is no additional ischemia-dependent promotion of FGF2 translation in ischemic skeletal muscle. Increased expression of FGF2 protein (all isoforms) has been detected in the urine, serum or ischemic tissues of patients with myocardial or limb ischemia (Campuzano et al., 2002; Cuevas

210

et al., 1997; Gu, Santiago, Olowe, & Weinberger, 1997; Hasdai et al., 1997; Rohovsky et al., 1996).

Enhanced FGF2 expression has also been observed in animal models of acute cardiac

ischemia/reperfusion injury or hindlimb ischemia (Arras et al., 1998; Bush et al., 1998; Chleboun

& Martins, 1994; Cohen, Vernon, Yaghdjian, & Hatcher, 1994; X. Liu & Zhu, 1999; Sullivan et

al., 2002; Walgenbach et al., 1995).

The upregulation of FGF2 expression in ischemia models is not limited to cardiac or

skeletal myocytes alone. Rodent models of focal cerebral ischemia have been shown to induce

Fgf2 mRNA and/or protein expression within 24h of the injury (T. N. Lin, Te, Lee, Sun, & Hsu,

1997; Speliotes et al., 1996; X.-C. Zhao et al., 2013). Additionally, optic or peripheral nerve

axotomy initiates release of FGF2 from astrocytes or glial cells (Duprey-Díaz, Blagburn, &

Blanco, 2012; Grothe et al., 1996; Meisinger & Grothe, 1997b). Drug-induced neurotoxicity that

produce lesions of the nigrostratal pathway is another injury condition where increases in Fgf2

mRNA and/or protein are present in astrocytes and microglia (Chadi, Cao, Pettersson, & Fuxe,

1994; Chadi & Gomide, 2004; Leonard et al., 1993; Nakagawa & Schwartz, 2006; Noda et al.,

2014; Yurek, Fletcher, Peters, & Cass, 2010). Furthermore, cultured cortical astrocytes in an

ischemia-like state produce increased FGF2 protein expression that largely accumulates in nuclear bodies (X. Liu & Zhu, 1999). This spatial distribution of FGF2 may provide some evidence for

the identity of the FGF2 isoform that is preferentially upregulated in during brain injury. The

HMW FGF2 isoforms has been shown to be predominantly localized to the nucleus and not

secreted in various cell types including cardiomyocytes, neurons, fibroblasts, and endothelial cells

(Arese et al., 1999; Chadi & Fuxe, 1998; Claus et al., 2003; Garmy-Susini et al., 2004; Liao et al.,

2010).

211

Detection of FGF2 protein by Western blotting allows for quantification of the levels of

each isoform and represents a more informative method than histological methods. The currently available FGF2 antibodies do not distinguish between the isoforms types; therefore, immunohisto-

or immunocyto- chemistry techniques utilized in the previously described reports are only useful

for demonstrating changes in overall FGF2. In response to neurotoxin-induced degeneration in the

rat brain, expression of all three isoforms was analyzed for up to 4 weeks after the lesions (Claus,

Werner, et al., 2004). The LMW or HMW isoforms displayed similar expression patterns and were

not different from intact rat brains. Using a similar injury model where only the levels of the HMW

FGF2 was reported, the 21kDa was unchanged while 23kDa expression was elevates relative to

intact brains (Silva, Fuxe, & Chadi, 2009). The distinction made in these few studies between the

endogenous FGF2 proteins is very critical to further assign cellular functions to the HMW and

LMW isoforms. This dissertation, for the first time, demonstrated that in skeletal muscle,

endogenous LMW or HMW FGF2 can regulate overall FGF gene or protein expression under

baseline conditions. Furthermore, this isoform-specific control of FGF2 expression persists under chronically ischemic conditions and is present at both the level of gene transcription and protein translation.

FGFs are widely expressed across several physiological systems where they serve as

ligands to a family of high affinity FGF receptors (FGFRs 1, 2, 3 & 4). FGFR1, 2 and 3 have

additional splice variants of the immunoglobulin-like domain III (IIIb and IIIc) which are important for FGF ligand-binding specificity (Eswarakumar et al., 2005; Itoh & Ornitz, 2011). The

most closely related FGF to FGF2, FGF1 is known to bind and activate all the FGFR splice variants

(Ornitz & Itoh, 2015). FGF2 (the LMW isoform) activates all FGFRc isoforms, FGFR3b and

FGFR4 (summarized in Table 2). To establish whether the phenotypic changes observed in

212

response to ischemia and the presence of only LMW FGF2 or HMW FGF2 were associated with

FGF receptor levels and/or activation, receptor gene and protein expression and activation was

determined in normal or ischemic muscles.

Interestingly, neither Fgfr2 mRNA (IIIb or IIIc form) nor protein was detected in the any

of the muscles analyzed. This finding supports previous observations by others in which Fgfr2

transcript expression was below the level of detection or present in very low quantities in myogenic

cells isolated from wildtype mouse limbs (Fon Tacer et al., 2010; Yablonka-Reuveni, Danoviz,

Phelps, & Stuelsatz, 2015). Similarly, very low levels of Fgfr2 was present in adult rat flexor digitorum brevis and tibialis anterior (TA) muscles (Kästner, Elias, Rivera, & Yablonka-Reuveni,

2000). In a study by Yablonka-Reuveni and coworkers, muscle from wildtype mouse limbs were isolated and separated into non-myogenic and myogenic cell populations. Low levels of Fgfr2c

mRNA were detected in the non-myogenic cells but absent from the myogenic population

(Yablonka-Reuveni et al., 2015). These non-myogenic cells are likely to include cells of neuronal

origin as the CNS, including the spinal cord, has high expression of Fgfr2c transcripts and little or no Fgfr2b (Fon Tacer et al., 2010; Hensel et al., 2012; Ratzka, Baron, & Grothe, 2011). These

studies confirm that the RT-PCR and immuoblotting experiments performed in this dissertation

were specific for Fgfr expression in skeletal myocytes and the absence of expression is not a result

of unsuccessful primer design or other errors in experimentation.

Receptor activation assays have shown that FGF2 (LMW isoform) is selective for the IIIc

variant of FGFR1 and has lower binding to FGFR1b (Beenken & Mohammadi, 2009; Xiuqin

Zhang et al., 2006). In studies where the expression of the IIIb isoform of Fgfr1 was distinguished

from the IIIc form, skeletal muscle expression of Fgfr1c was predominant over Fgfr1b (Fon Tacer et al., 2010; Hensel et al., 2012; S. E. Hughes, 1997). Likewise in the CNS (spinal cord), the Fgfr1c

213 transcript is more abundant over Fgfr1b (Fon Tacer et al., 2010; Hensel et al., 2012; Ratzka et al.,

2011). Both splice variants of Fgfr1 were present in all skeletal muscles analyzed from WT or

FGF2 HMW-only mice (Figure 35A). In normal muscle expressing only the HMW isoforms,

Fgfr1b transcript levels were depressed while Fgfr1c was unchanged. While Fgfr1c is the more abundant variant in muscle and vascular cells, FGF2 can still bind and activate the protein product of the Fgfr1b transcript (S. E. Hughes, 1997; Ornitz & Itoh, 2015; Ornitz et al., 1996). The down- regulation of the receptor could be attributed to the presence of increased HMW FGF2 protein expression in non-ischemic muscle. Both variants of Fgfr3 were not differentially expressed from

WT in FGF2 HMW-only mice (Figure 36A). Similarly to Fgfr1, the “b” isoform of Fgfr3 was lower in expression or similar to the level of detection in WT muscle or spinal cord (Fon Tacer et al., 2010; Hensel et al., 2012; Ratzka et al., 2011). The presence of only HMW FGF2 did not alter

Fgfr4 mRNA levels in non-ischemic muscles (Figure 37A). Fgfr4 mRNA has been shown to be present in equal amounts to Fgfr1 transcripts in muscle, brain and bone cells (Fon Tacer et al.,

2010; Han, Xiao, & Quarles, 2015; Hensel et al., 2012; Yablonka-Reuveni et al., 2015).

In response to ischemia, Fgfr1b or Fgfr1c mRNA expression was highest by 7 days in WT muscle, declining to control levels at 14 days (Figures 35B & C). When only the HMW isoforms of FGF2 were present, a similar pattern of Fgfr1 expression was observed with a marked increase in transcript levels at day 3 of ischemia. This effect can be attributed to the loss of LMW FGF2 protein as Fgf2-/- muscles exhibited a comparable pattern of Fgfr1 transcript expression. The results from the WT mice are not consistent with a rabbit arteriogenesis study where increases to

Fgfr1 mRNA in skeletal muscle was rapid, peaked at 6 hours after the femoral artery occlusion, and decreased to control levels by 3 and 7 days (Deindl et al., 2003). The absence of qRT-PCR experiments from ischemic muscles isolated before 24 hours precludes a true comparison between

214

the results of this dissertation and the work by the Schaper group but it does suggest that there

might be a bi-phasic (early and late phases) effect on Fgfr1 expression during hindlimb ischemia

(Deindl et al., 2003). Experimental evidence for this phenomenon was present in ischemic rat

extensor digitorum longus (EDL) muscles where VEGF mRNA and protein was elevated within

3 days of the injury, returned to normal by 7 days and a secondary wave of VEGF protein was

observed at 14 days (Milkiewicz, Hudlicka, Shiner, Egginton, & Brown, 2006). Similar to Fgfr1,

two splice variants of Fgfr3 (Fgfr3IIIb and Fgfr3c) were detected in skeletal muscle. In the

presence of ischemia and only HMW FGF2 or only LMW FGF2 expression, neither Fgfr3

transcript was altered (Figures 36B & C). By the same token, the single Fgfr4 mRNA was

unchanged with the loss of either FGF2 protein isoform (Figure 37B). This suggests that the loss

of LMW isoform expression modulates ischemic Fgfr expression, particularly the Fgfr1 isoforms.

It is likely a compensatory measure during a stress condition like ischemia to boost autocrine

FGFR1 signaling in the absence of LMW FGF2. This is not surprising as FGFR1 is the

predominant receptor for FGF2 in skeletal muscle (Fon Tacer et al., 2010; Hensel et al., 2012;

Yablonka-Reuveni et al., 2015), heart (Liao, 2008) and it shares this role with FGFR2 in bone and neuronal development and repair/regeneration (Du et al., 2012; Fei et al., 2013; Grothe et al., 2006;

Masumura, Murayama, Inoue, & Ohno, 1996; Nakajima et al., 2001).

FGFR expression was also evaluated at the protein level along with FGFR activation (via

phosphorylation). Protein expression of FGFR1, 3 or 4 was not affected by the presence of only

the HMW FGF2 isoforms in non-ischemic muscles. Likewise, cultured calvarial osteoblasts with

an overexpression of HMW FGF2 had no change in either FGFR1 or FGFR2 protein expression

(Sabbieti et al., 2013). Overexpression of the human FGF2 isoforms (22, 23 and 24kDa), however,

resulted in amplified Fgfr1c and Fgfr3c mRNA expression as well as increased nuclear

215

accumulation of FGFR1 protein in adult osteoblasts or bone marrow stromal cells (Xiao et al.,

2010, 2013). This pattern is also observed in human neuroblastoma cells transfected with human

23kDa FGF2 (Dunham-Ems et al., 2009). The AR4-2J pancreatic cell line also produced increased

Fgfr1 mRNA three-fold in the presence of 22.5kDa FGF2 overexpression (Estival et al., 1996).

On the other hand, FGF receptor protein expression was not altered in cultured NIH 3T3 fibroblasts

in response to HMW FGF2 (22, 22.5 and 24kDa) overexpression (Bikfalvi et al., 1995). There

appears to be a tissue-selective up-regulation of FGFR transcript or protein, especially FGFR1 in

the presence of only HMW FGF2 expression. In this dissertation, activation (tyrosine

phosphorylation) of FGFRs 1, 3 and 4 was also unchanged by HMW FGF2 expression (Figure

38). Basal FGFR1 activation has been shown to be decreased in mouse hearts overexpressing

24kDa HMW FGF2 with no change in the receptor protein expression (Liao et al., 2010; Manning

et al., 2013). These results confirm that in skeletal muscle, the HMW FGF2 isoforms act in an

intracrine manner and are not secreted like their paracrine/autocrine LMW counterpart to interact

with the FGFRs. This joins the increasing evidence suggesting that HMW FGF2 is a nuclear

growth factor with putative activity (Chlebova et al., 2009; Liao et al., 2009;

Yu et al., 2007).

After the induction of ischemia (3 or 7 days), FGFR 1, 3, and 4 protein expression or

activation was again evaluated and found to be similar between WT and HMW FGF2 expressing-

limbs (Figures 39-41). The study discussed above by the Schaper group also assessed the

expression of FGFR1 protein in the ischemic rabbit hindlimb (Deindl et al., 2003). The authors

observed an early up-regulation of both FGFR1 phosphorylation (i.e., activation) and protein

expression at 6 and 12 hours, respectively. As with the transcript expression data, these results

propose that the effect of ischemia on FGFR1 expression/activation occurs at an early phase that

216

is excluded from the results of this dissertation. The other FGFRs, FGFR3 and 4 might also be

similarly induced in ischemic muscle. Nevertheless, the timeframe chosen for the mRNA and

protein expression analyses is in agreement with other published reports of receptor tyrosine kinase

expression/activation during ischemia. VEGF-R2 (Flk-1), a receptor that positively modulates vascular growth and homeostasis in response to the VEGFs, was elevated within 3 days of ischemia and declining to control levels by 7 - 10 days in rat and rabbit muscles after femoral or iliac artery ligation (Lloyd, Prior, Yang, & Terjung, 2003; Milkiewicz, Hudlicka, Verhaeg, Egginton, &

Brown, 2003; Rissanen et al., 2002). However, Prior and coauthors observed no changes to VEGF-

R2 mRNA at any timepoint (up to 3 weeks) (Prior et al., 2004). Another tyrosine kinase receptor,

PDGFRα was also upregulated in expression at 3 days of ischemic and declining after 10 days

(Moriya et al., 2013).

Further proof of the changes in FGFR expression has also been observed in other muscle injury models. Chronic low frequency stimulation which has been shown to increase capillary growth in fast twitch fibers (Hudlicka et al., 1994; Hudlicka & Brown, 2009; Shen et al., 2009)

amplified Fgfr1 and Fgfr4 mRNA almost two-fold in rat tibialis anterior (TA) muscles

(Düsterhöft, Putman, & Pette, 1999). However, Fgfr1 was shown to be unchanged in rat EDL stimulated for 5 days (Brown et al., 1998). Fgfr1 mRNA/protein and Fgfr4 mRNA were elevated in response to mouse hindlimb suspension and this prevented muscle disuse atrophy (Eash, Olsen,

Breur, Gerrard, & Hannon, 2007). Similarly, both Fgfr1 and Fgfr4 transcripts were up-regulated in a spinal muscle atrophy (SMA) model (Hensel et al., 2012). Transient ischemia in the rat eye or brain selectively induces Fgfr1 mRNA within hours of injury and persists for days (Endoh,

Pulsinelli, & Wagner, 1994; Masumura et al., 1996; Miyashiro et al., 1998). Similarly several

“ischemia-like” insults (serum-free, glucose-free or glutamate addition) in cultured astrocytes

217

increased FGFR1 immunoreactivity (X. Liu & Zhu, 1999). Nerve lesion injury also induces

expression of Fgfr1-3 transcripts and FGFR1 protein at the lesion site (Grothe, Meisinger, & Claus,

2001; Grothe et al., 1996). Temporal expression analyses of skeletal fractures revealed early

upregulation of Fgfrs 1 and 2 at 1 day after injury which increased over the healing period with

maximal expression at 14 days post-fracture (Nakajima et al., 2001; Schmid et al., 2009). In

contrast, expression of Fgfr3 was not observed until 10 days in a rat model of femoral fracture

(Rundle et al., 2002).

At baseline or in response to ischemia, ablation of the HMW FGF2 isoforms (in FGF2

LMW-only mice) did not affect Fgfr expression at the RNA or protein level with the exception of

Fgfr4 (Figures 35 - 41). The presence of LMW FGF2 alone increased the levels of Fgfr4 transcripts after 3 and 7 days of ischemia (Figure 37B) which did not produce any alterations of FGFR4 protein expression or phosphorylation. Consistent with the results presented here, expression of only LMW FGF2 in the heart (HMWKO) did not affect FGFR1 protein expression at baseline

(Liao et al., 2010; Manning et al., 2013). FGF receptor levels were decreased in NIH 3T3 cells overexpressing 18kDa FGF2 (Bikfalvi et al., 1995). Conversely, osteoblastic-specific LMW FGF2 overexpression increased basal Fgfr1 and Fgfr2 mRNA expression (Xiao et al., 2009, 2014). The

AR4-2J pancreatic cell line also produced increased Fgfr1 mRNA two-fold in the presence of

18kDa FGF2 overexpression (Estival et al., 1996). Cultured calvarial osteoblasts with an overexpression of LMW FGF2 had increased FGFR1 protein levels with no change in FGFR2 expression (Sabbieti et al., 2013). These studies suggest that shifting the FGF2 equilibrium towards increased LMW FGF2 results in upregulation of its dominant receptors in the respective tissues.

The existence of FGFR-independent signaling and responses induced by endogenous

HMW FGF2 adds a layer of complexity to the LMW/HMW/FGFR interactions. It is widely

218

accepted that of all endogenous FGF2 proteins, it is the 18kDa LMW FGF2 that interacts with the

cell surface FGF receptors to mediate autocrine and paracrine signaling responses (Chlebova et

al., 2009; Yu et al., 2007). This is due to the localization of this isoform to the cytoplasm, cell

surface and the extracellular matrix (Kardami et al., 2007; Presta et al., 2005). It can also be

secreted at times of cellular stress (Piotrowicz, Martin, Dillman, & Levin, 1997; Steringer, Müller,

& Nickel, 2014). On the other hand, the HMW isoforms with NH2 terminal nuclear localization

signals are present exclusively in the nuclear compartment and do not typically interact with the

membrane-bound FGF receptors (Claus et al., 2003; R Z Florkiewicz, Baird, & Gonzalez, 1991).

Studies with exogenous administration of recombinant HMW FGF2 have reported FGFR-

dependent cellular signaling on par with activation by LMW FGF2 (Gualandris et al., 1994; Jiang

et al., 2007; Moscatelli, Joseph-Silverstein, Manejias, & Rifkin, 1987). There is evidence to

suggest that there are differences between the cellular activities of exogenous and endogenous

HMW FGF2. Unlike LMW FGF2, endogenous HMW FGF2 is seldom secreted under basal or

stress conditions (Claus et al., 2003; House et al., 2003; Pintucci et al., 2005). Cardiac

myofibroblasts have been shown to accumulate and export HMW FGF2 after angiotensin II

stimulation (Santiago et al., 2011). The HMW isoforms share considerable sequence homology with LMW FGF2 including the binding domains and affinities for HSPG or FGFR1 so it is likely that the HMW FGF2/FGFR-dependent signaling observed in these studies is not representative of

actions of endogenous HMW FGF2 (Moosa Mohammadi et al., 2005).

In NIH 3T3 cells overexpressing all the HMW isoforms (22, 22.5 and 24kDa FGF2) or

only the LMW isoform (18kDa), similar levels of cell migration were observed and were

subsequently inhibited by the introduction of a dominant negative form of the FGF receptor in

only the cells with transfected with 18kDA cDNA (Bikfalvi et al., 1995). Likewise, HMW FGF2-

219

induced alteration of PKCδ, PKCε or ERK1/2 activation in cultured pancreatic cells were not

affected by dominant negative FGFR1 (Gaubert et al., 2001; Hortala, Estival, Pradayrol, Susini, &

Clemente, 2005). FGF2 neutralizing antibodies also did not revert cellular responses induced by

HMW FGF2 indicating that paracrine/autocrine signaling by secreted FGF2 was not responsible for the effect of HMW FGF2 (Estival et al., 1996; Gaubert et al., 2001; Ma et al., 2007; Nindl et al., 2004; Pasumarthi et al., 1996; Xiao et al., 2003). In response to acute I/R injury, Liao and

colleagues reported decreased PKCα activity (translocation), increased MAP kinase activation

(p38, ERK or JNK) and amplified Akt activation in hearts expressing only the HMW isoforms or

overexpressing the human 24kDa FGF2 (Liao, 2008; Liao et al., 2007). These cellular responses likely represent intracrine signaling by the nuclear localized HMW FGF2 that are not mediated by classical interaction with membrane FGFRs.

Integrative nuclear FGF receptor 1 signaling (INFS) is a novel mechanism that involves gene regulation by a nuclear regulatory complex of FGFR1 and its partner CREB binding protein

(CBP) (Fang, Stachowiak, Dunham-Ems, Klejbor, & Stachowiak, 2005; Stachowiak, Maher, &

Stachowiak, 2007). Nuclear FGFR1 (nFGFR1) is a soluble full-length protein that accumulates in the nucleus bypassing Golgi processing and does not represent translocation of cell-surface FGFR1

(Dunham-Ems et al., 2009; Stachowiak et al., 2015). Gene activation by nuclear FGFR1 is not blocked after inhibition of ligand-induced FGFR1 internalization from the cell surface and does not require tyrosine kinase activity which supports the theory of INFS (Stachowiak et al., 2015,

2007). Immunolocalization and co-immunoprecipitation experiments have shown that HMW

FGF2 (23kDa) interacts with nuclear FGFR1 and is suggested to be part of a chromatin-bound complex with the transcription factor co-activator, CBP (Baron et al., 2012; Dunham-Ems et al.,

2009; Stachowiak et al., 2015). In contrast, LMW FGF2 does not interact with nuclear FGFR1

220

(Dunham-Ems et al., 2009; Stachowiak et al., 2015). Evidence for INFS induction has been observed in differentiation of neuronal populations, neurogenesis and neurodevelopment, activation and invasion of pancreatic stellate cells (Förthmann et al., 2013; Kiyota, Ingraham,

Jacobsen, Xiong, & Ikezu, 2011; Woodbury & Ikezu, 2013). Nuclear FGFR1 is also involved in direct regulation of several processes genes including; Fgf2, Fgf23, Nurr1, c-Jun and Nur77

(Baron et al., 2012; Han et al., 2015; Stachowiak et al., 2015; Terranova et al., 2015).

An intriguing phenotype emerged from the expression experiments performed in Fgf2-/-

mice. While basal expression of Fgfr 1b, 1c, 3b, 3c or 4 mRNA was unchanged in skeletal muscles

with deletion of Fgf2, ischemia produced an increase of Fgfr1c transcript at 3 days of ischemia.

The level of Fgfr1 activation observed was in keeping with this change in gene expression and

elevated in response to ischemia after 7 days of ischemia. Fgfr3 expression was unaltered at the

mRNA level but protein expression was increased at 3 days of ischemia. No differences in mRNA

or protein expression of Fgfr4 were present in Fgf2-/- muscles. Characterization of FGF receptor

expression profile in Fgf2-deficient skeletal muscles has never been reported. However, others

have documented the effects of Fgf2 ablation on FGFR expression in other tissues. Ortega and

coworkers described similar levels of Fgfr1 and Fgfr2 mRNA in brain, heart, liver and testes of

Fgf2-/- mice as WT (Ortega et al., 1998). Likewise, the Grothe group detected no changes in any

Fgfr transcripts in Fgf2 knockout spinal cord, striatum and ventral mesencephalon (Ratzka et al.,

2011). However, changes were present when expression of the receptors was determined in either

nuclear or cytoplasmic compartments of dopaminergic (DA) neurons. Fgf2-deficient DA neurons

had increased nuclear Fgfr1 accumulation but similar cytoplasmic Fgfr1 (Ratzka, Baron,

Stachowiak, & Grothe, 2012). The over-activation of INFS signaling (measured by nuclear-

localized FGFR1) is reversed after treatment with recombinant 18kDa which suggests the presence

221

of selective FGF2 isoform-induced regulation of the equilibrium between intracrine and

cytoplasmic in FGF2-FGFR signaling.

The role of INFS in cellular responses of the endogenous HMW FGF2 isoforms has yet to

be studied in isoform-specific knockout mice. FGFR1 is currently the only FGF receptor known

to be present in the nucleus, where it is present in its full-length form and in association with HMW

FGF2 (Dunham-Ems et al., 2009; Stachowiak et al., 2003). The experiments to determine the

FGFR expression/activation in non-ischemic and ischemic skeletal muscles did not consider the

apparent compartmentalization of FGFR1 and the other FGFRs into cell-membrane or nuclear factions and represent overall cell/tissue expression of the receptors. Differences in gene regulation by membrane FGFRs and nuclear FGFR in ischemic limbs may represent a novel mechanism for mediating the FGF2 isoform-specific effect on ischemia-induced revascularization.

While whole muscle lysates were analyzed for the FGFR activation and expression results presented here, the levels of receptor phosphorylation and expression in other cells distinct from skeletal myocytes are also important. Skeletal muscle represents a mixture of myofibers and non- myogenic cell types including satellite cells, neurons, fibroblasts, and cells of the vascular compartment; endothelial (ECs) and smooth muscle (vSMCs). Immunodetection of FGFR1 in rodent muscle sections or rat myofiber cultures showed strong expression in myocytes, fibroblasts, pericytes, myogenic cells or vSMCs and no expression in ECs (Brown et al., 1998; Deindl et al.,

2003; Eash et al., 2007; Kästner et al., 2000). Conversely, positive FGFR4 staining was exclusively localized to myogenic cells and absent in non-myogenic cells (Cornelison, Filla, Stanley,

Rapraeger, & Olwin, 2001; Yablonka-Reuveni et al., 2015). Human ECs and vSMCs of aortic or umbilical origin in cultured singly or together in EC/vSMC co-culture spheroids strongly express

Fgfr1 mRNA and had lower expression of Fgfr2 or Fgfr4 (Antoine et al., 2005; Brogi et al., 1993).

222

In an extensive immunohistochemical study of adult human tissues, Hughes reported strong and

wide spread expression of Fgfr1-4 mRNAs in the endothelium and media of microvessels, veins

and arteries of epithelial and mesenchymal (S. E. Hughes, 1997).

The effect of isoform-specific FGF2 signaling on gene regulation has been investigated in

neovascularization processes. Findings suggest that FGF2 may exert an indirect control on

angiogenesis by promoting the expression of other angiogenic growth factors (Murakami &

Simons, 2008; Presta et al., 2005). This is proposed to occur in concert with its (FGF2) own direct

and potent modulation of angiogenesis. The strongest evidence for the indirect control of

angiogenesis by FGF2 is demonstrated by its effect on the VEGF system. FGF2-induced

angiogenesis in Matrigel implants coincided with increased expression of VEGF mRNA and

protein by infiltrating stromal cells and fibroblasts. Neutralizing VEGF antibodies suppressed the

FGF2-induced angiogenesis in this model (Claffey et al., 2001). Similarly, antibody blockade of

VEGFA and the VEGF receptor, Vegfr1 in murine brain ECs or HUVECs inhibited FGF2-

mediated capillary growth and morphogenesis (Kanda, Miyata, & Kanetake, 2004). In vivo, FGF2-

induced angiogenesis in the mouse cornea was associated with increased VEGF mRNA and

protein expression in endothelial cells of growing capillaries (Seghezzi et al., 1998). FGF2 gene

transfer to ischemic mouse hindlimbs targeted nonendothelial mesenchymal cells and induced

expression of angiogenic VEGF pathways (Fujii et al., 2006, 2008). Co-stimulation of cultured

endothelial cells with FGF2 and VEGF-A enhanced mural cell recruitment as well as endothelia-

mural interaction. Furthermore, neovessel formation required synergistic upregulation of Pdgfb nd

PDGF receptor β (Pdgfrβ) gene expression (Kano et al., 2005). FGF2 treatment with either PDGF-

A or PDGF-B synergistically promoted PDGFR and FGFR expression in ECs and vSMCs, respectively (Nissen et al., 2007). PDGF-A/FGF2 and PDGF-B/FGF2 combinations also

223 synergistically stimulated angiogenesis in mouse cornea and ischemic skeletal muscles (J. Zhang et al., 2009). PDGF, likewise plays a synergistic role in the FGF2-induced transcriptional and translation regulation of HGF expression in vitro in cultured cells and an in vivo ischemia model

(Onimaru et al., 2002). Overall, these studies suggest that FGF2 signaling lies upstream of enhanced growth factor signaling in ischemia-induced neovascularization. Interestingly, the

VEGF expression in ECs was similarly augmented in ECs transfected with LMW FGF2 or HMW

FGF2 cDNAs. However, when FGF2 neutralizing antibodies were administered, this upregulation is inhibited in LMW FGF2-expressing ECs but unchanged in HMW FGF2 cells (Seghezzi et al.,

1998). The intracrine effect of HMW FGF2 on VEGF expression follows the trend of FGFR- independent signaling by the HMW isoforms discussed previously.

The FGF2-isoform specific autocrine/intracrine signaling is proposed to underlie the selective transcriptional activation in response to overexpression of either only LMW FGF2

(18kDa) or HMW FGF2 (22, 22.5 and 24kDa) in NIH 3T3 cells. Quarto and colleagues performed a microarray analysis and reported that each class of FGF2 isoform differentially modulates a distinct profile of genes in stimulated cells (Quarto et al., 2005). In the HMW FGF2 -expressing cells, genes involved in growth arrest and tumor suppression were up-regulated while those associated with cell proliferation and tumor growth were down-regulated and unchanged in LMW

FGF2 cells. Using a proteome antibody array, the expression profile of angiogenesis-related proteins in HMW or LMW expressing skeletal muscle was investigated (Table 13). Protein isolated from sham or ischemic muscles after 7 days of ischemic was hybridized to an array spotted with antibodies to 53 different proteins. These proteins were classified based on their roles in angiogenic process into the following functional groups: growth factors, pro-angiogenic, angiostatic, ECM remodeling, chemokines and inflammatory mediators.

224

The array was spotted with several growth factors including FGFs, VEGFs, PDGFs, &

HGF which have been previously shown to be regulated by FGF2 (Murakami & Simons, 2008;

Presta et al., 2005). A novel role for FGF2 in the regulation of FGF7 protein expression was

observed. HMW FGF2 induced an increase in FGF7 expression during ischemia. The potential

role of FGF7 in ischemia-induced vascular growth is yet unknown as it has only been linked to the

proliferative phase of acute or chronic wound healing by promoting of keratinocyte proliferation

and migration (Maretzky et al., 2011; Niu et al., 2007; Yen et al., 2014). It is known that FGF2

autoregulates its own expression as described in the Chapter 1 of this dissertation as well as

modulating Fgf23 transcription and translation (Jimenez et al., 2004; Stachowiak et al., 2015; Xiao

et al., 2010). The presence of only the HMW isoforms up-regulated the expression of two IGF

binding proteins, IGFBP-3 and IGFBP-10, which modulate the activity of the IGF axis. IGFBP-3

is an angiostatic protein with inhibitory effects on tumor growth while IGFBP-10 promotes

angiogenesis by regulating angiogenic gene transcription in response to hypoxia (Kubota &

Takigawa, 2007; Malik et al., 2015; Yamada & Lee, 2009). Microarray studies detected increased

IGFBP3 gene expression within a day of hindlimb ischemia surgery and this change was

associated with regeneration of the ischemic muscle (Matsakas, Yadav, Lorca, Evans, & Narkar,

2012; Paoni et al., 2002). Two other genes in the insulin-like growth factor (IGF) axis, IGFBP-2

and IGF2 were both increased two-fold during 24kDa HMW FGF2-induced cell proliferation in

the AR4-2J cancer cell line (Hortala et al., 2005). Protein levels of CX3CL1 (fractalkine), a

chemokine was also elevated with expression of only HMW FGF2 and the presence of ischemia.

Comparison of the expression profile induced by selective HMW FGF2 expression in this

dissertation with the microarray analysis performed by Quarto and colleagues is not entirely

straightforward as the former analyzed expression of a specific group of proteins (angiogenesis-

225 related), while the later performed a whole genome expression survey (Quarto et al., 2005). The presence of selective FGF2 isoform expression occurred from birth in the muscle tissues analyzed while the fibroblasts were transfected with FGF2 cDNAs for only 48 hours. Nevertheless, it is important to note that like increased anti-angiogenic IGFBP-3, the microarray studies reported up- regulation of several genes with growth arrest and tumor inhibition activities in HMW FGF2 cells.

The proteome array identified epidermal growth factor (EGF) as elevated in the presence of LMW FGF2 (Table 13). Similarly to FGF7, EGF has been implicated in wound healing; particularly, in dermal wounds where it enhances proliferation and migration of fibroblasts and keratinocytes (Haase et al., 2003; Seeger & Paller, 2015). The pro-angiogenic factor angiopoietin- like 4 (Angptl4) was up-regulated in LMW FGF2 3T3 fibroblasts (Quarto et al., 2005). However, other members of the angiopoietin family (angiopoietins-1 and -3) analyzed on the array were not changed in the FGF2 LMW-only muscles. Angiopoetin-1 (Ang-1) is a potent angiogenic factor mainly expressed by perivascular cells of mature vessels (Eklund & Saharinen, 2013; Herbert &

Stainier, 2011). Ang-1-mediated activation of the EC-specific receptor Tie2 is important for the vessel stabilization and maturation (Eklund & Saharinen, 2013; Fagiani & Christofori, 2013).

Angiopoietin-3 also has positive angiogenic activity in human ECs and glioblastoma cells (H. J.

Lee et al., 2004; Morisada, Kubota, Urano, Suda, & Oike, 2006). This suggests that, in the context of chronic limb ischemia, there is an absence of LMW FGF2-induced expression of angiogenic factors.

The importance of FGF2-induced activation of angiogenic growth factor expression is further called into question when the expression profile of the angiogenic proteins is examined in

Fgf2-/- ischemic muscles. Like FGF2 LMW-only, expression of angiogenic factors like VEGF,

PDGF or HGF was not different from WT in Fgf2-/- mice. Two proteins of significant importance

226 during ECM remodeling were upregulated in Fgf2-/- muscles after 7 days of ischemia, MMP-9 and PAI-1. PAI-1 is the main inhibitor of the urokinase-type plasminogen activator (uPA) system which cleaves adhesion molecules, growth factors and their receptors as well as the ECM degradation required for endothelial and smooth muscle cell migration during angiogenesis

(Papetti & Herman, 2002; Van De Craen et al., 2012). The active enzyme of the uPA system is plasmin, a serine protease converted from the plasminogen proenzyme (Medcalf, 2007). In addition to the ECM, plasmin also converts latent pro-MMPs like MMP9 into active enzyme which in turn degrade the ECM (Arroyo & Iruela-Arispe, 2010; Q. Chen et al., 2013). The role of MMP-

9 is revascularization is as yet unclear since normal angiogenesis and arteriogenesis but increased necrosis was observed in MMP9-deficent ischemic limbs (Meisner et al., 2014). Pharmacologic inhibition of PAI-1 ameliorated tissue necrosis and enhanced angiogenesis in mice with hind limb ischemia (Tashiro et al., 2012). This improved angiogenesis was associated with increased immunolocalization of FGF2 in the PAI-1 inhibitor-treated muscles. This suggests that the presence of elevated PAI-1 and MMP-9 may underlie the reduced muscle viability observed in

Fgf2-/- mice (discussed in Chapter 1).

Surprisingly, no effect of selective FGF2 expression was observed in the expression of several inflammatory cytokines and chemokines. These chemokines/cytokines mediate the initial phase of wound healing/repair that occurs immediately following ischemic or other forms of tissue injury (Keeley et al., 2008; Silvestre et al., 2008). This is the inflammatory response that is critical for the eradication of infectious microbes, clearance of cellular waster or debris and eventually repair or resolution (Raman, Sobolik-Delmaire, & Richmond, 2011; Tidball, 2005, 2011).

Vascular growth and inflammation are closely linked and inter-dependent in a number of

(patho)physiological conditions including chronic ischemia (Keeley et al., 2008; Romagnani,

227

Lasagni, Annunziato, Serio, & Romagnani, 2004; Silvestre et al., 2008). The Presta group has

shown the necessity of inflammatory cytokines and chemokines for FGF2 induced-angiogenesis

in vitro and in vivo (Andrés et al., 2009; Presta et al., 2009). FGF2 stimulation of cultured

microvascular ECs up-regulated the expression of several pro-inflammatory genes including Ccl2,

Ccl7, Cxcl31, Cxcl1, Opn, IL-6. These cytokines and chemokines served as chemoattractants for

the recruitment of infiltrating leukocytes to vascularizing areas of FGF2-treated Matrigel plugs

(Andrés et al., 2009). Of these chemokines, CCL2 also known as MCP-1 is of particular

importance in FGF2-induced neovascularization. MCP-1, the most extensively studied of the CC

chemokine ligand family, regulates inflammation by recruiting leucocytes, such as monocytes, to

sites of tissue injury and infection (Capoccia, Gregory, & Link, 2008; Shireman, 2007). MCP-1

infusion potently induces angiogenesis and collateral growth in the ischemic rabbit hindlimb

(Arras et al., 1998; Deindl et al., 2003; Hoefer et al., 2001). Furthermore, MCP-1 mRNA and

protein is increased in response to FGF2 treatment in migrating BAECs and ischemic skeletal

muscle (Fujii et al., 2006; Jih et al., 2001; Onimaru et al., 2002; Wempe, Lindner, & Augustin,

1997). The time course of MCP-1 upregulation follows closely with the activation and trafficking

of monocytes/macrophages to the site of ischemic injury (Arras et al., 1998; Deindl et al., 2003).

The role of neutrophils, a subset of mononuclear leucocytes in FGF2-driven angiogenesis is not as

well understood.

Neutrophils predominate early in the inflammatory cell population that infiltrates injured/ischemic muscle where they may play a beneficial or negative role in the resolution of muscle injury (Tidball & Villalta, 2010; Tidball, 2011). The former is due to the phagocytic activity on necrotic debris and secretion of factors that attract and transactivate monocytes/macrophages while the latter is linked to the secondary damage from the generation of

228 free radicals (Smith et al., 2008; Tidball & Villalta, 2010; Tidball, 2005). Within 3-5 days of injury, they are replaced by monocytes/macrophages which is thought to signify the start of injury resolution (Smith et al., 2008; Tidball, 2005). Growth factors like FGF2 and other angiogenic cytokines are released by macrophages themselves or from activated ECs to mediate the angiogenesis process (Andrés et al., 2009; Presta et al., 2009; Shireman, 2007). The timing and extent of the inflammatory response in ischemic muscle is very critical to the regulation of ischemic revascularization by FGF2.

To account for differences due to selective FGF2 isoform expression, inflammatory cell

(neutrophil and monocyte) density was evaluated in non-ischemic skeletal muscle. Loss of LMW

FGF2 (FGF2 HMW-only muscles) did not significantly change neutrophil or monocyte density

(Figure 42A and 43A). Similarly, ablation of the LMW isoform had no effect on basal neutrophil or monocyte density (Figure 42B and 43B). This suggests that the form of FGF2 present in non- injured tissues does not affect the inflammatory cell population. In the presence of HMW FGF2, the expected increase in neutrophil recruitment was observed in the ischemic limbs (Figure 42B).

This was unlike the WT muscles where neutrophil density was unchanged from the contralateral

(sham-operated limb) after 3 days of ischemia. LMW FGF2-only mice also had no difference in neutrophil infiltration in response to ischemia.

Several studies have shown increased neutrophil accumulation in WT skeletal muscle after

3 days of chronic injury (Contreras-Shannon et al., 2007; Martinez et al., 2010; Meisner et al.,

2014; Shireman et al., 2006). The divergence between the results presented here and these published reports may be attributed to the use of immunohistochemistry. The use of flow cytometry with multiple markers to detect and quantify neutrophil cells might prove to be more conclusive. This method has been applied in a number muscle injury studies to analyze resident

229

and circulating inflammatory cells and they show early, increased accumulation of neutrophils in

ischemic muscle (Capoccia et al., 2008; Contreras-Shannon et al., 2007; Martinez et al., 2010;

Tashiro et al., 2012). Secondary measures of histological staining using neutrophil markers like

myeloperoxidase (MPO), granulocyte-1 marker, or Ly-6 were also performed. Another reason for the discrepancy in the WT ischemic muscle may be timing of the tissue sampling (Oklu et al.,

2013). While some have suggested that neutrophils have a relatively short life span and are not detectable 1-2 days after induction of skeletal muscle injury (Meisner & Price, 2010; Tidball,

2011), others have observed neutrophil accumulation as late as 3 days in injured muscle

(Contreras-Shannon et al., 2007; Martinez et al., 2010; Uaesoontrachoon, Wasgewatte Wijesinghe,

Mackie, & Pagel, 2013). Furthermore, histological detection of robust neutrophil accumulation in ischemic muscle has been observed as late as 14 days after the ischemic surgery (Tashiro et al.,

2012). The lack of additional timepoints earlier than day 3 at 12h or 24h after ischemia is a limitation of this study. It is entirely possible that the peak of neutrophil density in the WT ischemic limbs occurs early and the day 3 results capture a moment when the inflammatory cell infiltrate switches over from a largely neutrophil population to monocytic one. This shift is very critical for the resolution of the inflammatory response and tissue healing as the persistence of neutrophils can further exacerbate the tissue damage by elevating oxidative stress and monocytes mediate the clearance of neutrophils.

Strong inflammatory cell infiltration has been observed during angiogenesis in chick embryonic CAM and subcutaneous Matrigel plug assays treated with recombinant LMW FGF2

(Andrés et al., 2009). An abundance of the cell infiltrate was identified to be positive for F4/80, a cell marker for monocytes. Similarly, Zittermann and Issekutz reported that FGF2 stimulated monocyte recruitment to dorsal skin sites as well synergistically enhanced the chemotactic

230

response to other proinflammatory cytokines (Zittermann & Issekutz, 2006a, 2006b). Consistent

with these studies, LMW FGF2-expressing limbs had elevated Mac-3-positive cell (monocytes)

densities in response to ischemia (Figure 43B). Monocyte recruitment was assessed in the ischemic

muscles at 3 or 7 days and increases were observed at both timepoints. HMW FGF2 expression

had no significant effect on monocyte infiltration until 7 days after the induction of ischemia. The

ability of LMW or HMW FGF2 to mediate chemotactic movement of neutrophils during ischemia

was similar and not changed from WT FGF2 expression. The pattern and timing of recruitment in

the WT tissues is consistent with several studies where neutrophil infiltration in ischemic muscle

was increased at 3 and 7 days (Contreras-Shannon et al., 2007; Martinez et al., 2010; Shireman et

al., 2007, 2006; Tashiro et al., 2012). Other studies reported declining monocytes levels starting at

7 days of ischemia (Capoccia et al., 2008; McClung et al., 2012). In cardiotoxin-induced muscle

injury, macrophage recruitment continued for 14 days after induction of the insult (Martinez et al.,

2010).

231

Figure 44: Schematic outlining the cellular pathways involved in the FGF2 isoform-mediated response to hindlimb ischemia.

Inhibit Activate Unknown interaction

232

Chapter 3: FGF2 isoforms in the regulation of skeletal muscle repair and regeneration

RESULTS III

Chronic ischemia induces an injurious phenotype in the vascular compartment of ischemic

limbs but also affects the viability of the muscles. Therefore, effective compensation for vascular

obstruction in PAD requires not only the promotion of vascular growth via angio- or arteriogenesis

but also the repair of ischemic-injured muscle (McClung et al., 2012; Scholz et al., 2002). Skeletal

muscle is capable of endogenous repair and regeneration in response to ischemic damage due to

the presence of a resident stem cell population known as satellite cells (Chang & Rudnicki, 2014;

Kang & Krauss, 2010; Le Grand & Rudnicki, 2007). These satellite cells, located between the

sarcolemma and basal lamina, are quiescent in resting adult muscle but become activated during

injury (Ceafalan et al., 2014). The myogenic repair program includes satellite cell activation,

proliferation, and ends with differentiation to myoblasts and fusion with existing tubes or to one

another to form new myofibers (Dumont, Wang, & Rudnicki, 2015; Rudnicki, Le Grand,

McKinnell, & Kuang, 2008; Yin, Price, & Rudnicki, 2013).

Satellite cell activation and subsequent differentiation to newly generated myofibers is

regulated by a family of transcription factors known as the myogenic regulatory factors (MRFs)

(Figure 45). These MRFs include the paired domain transcription factor, paired box 7 (Pax7) which

is expressed by all satellite cells, myogenic factor 5 (Myf5) and myoblast differentiation protein

(MyoD) which are highly expressed by activated satellite cells and some myoblasts, and myogenin

and muscle-specific regulatory factor 4 (MRF4) which are upregulated during differentiation of

myoblasts (Dumont et al., 2015; Karalaki, Fili, Philippou, & Koutsilieris, 2009; Yablonka-

Reuveni, Day, Vine, & Shefer, 2008; Yin et al., 2013). Traditional angiogenic factors like VEGF, angiopoietins, MMPs, or FGFs may potentially serve as responders to skeletal muscle injury by

233 modulating myogenesis (Doukas et al., 2002; McClung et al., 2015; Meisner et al., 2014; Rissanen et al., 2002; Yablonka-Reuveni et al., 2015).

234

Figure 45: Adult myogenesis is coordinated by a group of transcription factors (MRFs) that

regulate satellite cell activation, differentiation/self-renewal, and fusion. Satellite cells exit quiescence and become activated in response to muscle injury (exercise or ischemia). The satellite stem cells (Myf-) and satellite progenitor cells (Myf+) are characterized by their expression of Pax7. Proliferating satellite cells and their progeny are often referred to as myogenic precursor cells or adult myoblasts. Myoblasts express both MyoD and Myf5.

Following proliferation, myoblasts begin differentiation by downregulating Pax7. Terminal differentiation and fusion begins with the expression of Myogenin, which together with

MyoD activate muscle specific structural and contractile genes. MRF4 is further required for hypertrophy of the new fibers. During regeneration, activated satellite cells have the capability to return to quiescence to maintain the satellite cell pool. Reprinted with permission from Elsevier (Le Grand & Rudnicki, 2007).

235

Absence of LMW FGF2 influences Pax7 expression during ischemia

To determine the gene expression levels of the MRFs, RNA was isolated from non-

ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only skeletal tissues. First strand cDNA was amplified by qRT-PCR using primers for the MRFs (Pax7, Myf5, MyoD, Myogenin and

MRF4) and an internal control gene (18S). Pax7 and Myf5 mRNA expression were both unchanged

in Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only muscles relative to non-ischemic WTs

(Figures 46A and 47A).

To characterize the effect of FGF2 isoform and ischemia on Pax7 and Myf5 expression, mRNA expression was analyzed at 1, 3, 7 and 14 days the induction of ischemia. Transcript levels were presented as a ratio of sham muscles to address any changes related to the hindlimb ischemia surgery. After 3 days of ischemia, Pax7 expression in all groups was decreased from day 1 and increasing by day 7 (Figure 46B). Expression was not changed from WT in the absence of all

FGF2 (Fgf2-/-) and in the absence of the HMW isoforms (FGF2 LMW-only) at all timepoints of ischemia. In the absence of LMW FGF2 (FGF2 HMW-only), Pax7 expression was significantly increased after 14 days of ischemia (p<0.05 vs. WT; Figure 46B).

WT and Fgf2-/- ischemic muscles had a similar profile of Myf5 expression during ischemia where expression after 3 days was increased from day 1 and decreased at after 7 days and returned to normal levels at day 14 (Figure 47B). Ischemic skeletal muscle from FGF2 LMW-only and

FGF2 HMW-only also shared a similar pattern where day 3 expression was decreased from day 1 and increasing by day 7 and returned to normal after 14 days (Figure 47B). For all mouse groups, no significant changes in Myf5 expression were observed relative to WTs.

236

Pax7 expression (non-ischemic muscle) A. 2 WT Fgf2-/- FGF2 LMW-only 1.5 FGF2 HMW-only

1 (arbitrary units) (arbitrary

0.5 Relative mRNA expression expression mRNA Relative

0

Pax7 expression (ischemic muscle) B. 6

WT Fgf2-/- 5 FGF2 LMW-only FGF2 HMW-only # * 4

3

(arbitrary units) (arbitrary 2

1 mRNA Expression relative sham to relative Expression mRNA

† † 0 1 3 7 14 Days after ischemia surgery 237

Figure 46. Pax7 mRNA expression in non-ischemic, sham and ischemic skeletal muscles. (A)

Quantitative RT-PCR analysis for Pax7 gene expression in non-ischemic skeletal muscle relative to WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or

FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles (B) Time course of Pax7 mRNA expression. For each

timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic

muscles are expressed relative to corresponding sham expression levels in WT (blue),

Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are

presented as mean ± SEM. n=4 for each mouse group at each timepoint. *p<0.05 vs. Day 1

cohort, #p< 0.05 vs. WT, †p<0.05 vs. FGF2 HMW-only.

238

A. Myf5 expression (non-ischemic muscle)

2 WT Fgf2-/- FGF2 LMW-only 1.5 FGF2 HMW-only

1 (arbitrary units) (arbitrary 0.5 Relative mRNA expression expression mRNA Relative

0

B. Myf5 expression (ischemic muscle) 12 WT Fgf2-/- FGF2 LMW-only 10 FGF2 HMW-only

8

6

4 (arbitrary units) (arbitrary

2 mRNA Expression relative sham to relative Expression mRNA

0 239 1 3 7 14

Days after ischemia surgery Figure 47. Myf5 mRNA expression in non-ischemic, sham and ischemic skeletal muscles. (A)

Quantitative RT-PCR analysis for Myf5 gene expression in non-ischemic skeletal muscle

relative to WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or

FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles (B) Time course of Myf5 mRNA expression. For each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic muscles are expressed relative to corresponding sham expression levels in WT (blue),

Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are presented as mean ± SEM. n=4 for each mouse group at each timepoint.

240

Loss of HMW FGF2 expression decreases basal levels of MyoD

MyoD and myogenin expression levels in non-ischemic WT, Fgf2-/-, FGF2 LMW-only and

FGF2 HMW-only skeletal muscles were also examined. MyoD mRNA expression was decreased

in both Fgf2-/- and FGF2 HMW-only mice (p<0.05 vs WT) while FGF2 LMW-only expression

was unchanged relative to non-ischemic WT (Figure 48A). Myogenin expression was unchanged

from WT in Fgf2-/-,FGF2 LMW-only or FGF2 HMW-only muscles (Figure 49A)

To describe the effect of specific FGF2 isoform expression and hindlimb ischemia on

MyoD and myogenin transcript levels, sham and ischemic muscles were isolated and analyzed for

MyoD and myogenin expression. In Fgf2-/-, FGF2 LMW-only, and FGF2 HMW-only limbs, no changes from WT expression levels of either MyoD (Figure 48B) or myogenin (Figure 49B) were detected at 1, 3, 7 or 14 days of ischemia.

A. MyoD expression (non-ischemic muscle)

2 WT Fgf2-/- FGF2 LMW-only 1.5 FGF2 HMW-only

1 # (arbitrary units) (arbitrary

0.5 # Relative mRNA expression expression mRNA Relative

241 0 B. MyoD expression (ischemic muscle) WT 9 Fgf2-/- FGF2 LMW-only 8 FGF2 HMW-only 7

6

5

4 (arbitrary units) (arbitrary 3

2 mRNA Expression relative sham to relative Expression mRNA 1

0 1 3 7 14 Days after ischemia surgery

Figure 48. MyoD mRNA expression in non-ischemic, sham and ischemic skeletal muscles.

(A) Quantitative RT-PCR analysis for MyoD gene expression in non-ischemic skeletal muscle

relative to WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or

FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles (B) Time course of MyoD mRNA expression. For each

timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic

muscles are expressed relative to corresponding sham expression levels in WT (blue),

Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are

presented as mean ± SEM. n=4 for each mouse group at each timepoint. #p< 0.05 vs. WT.

242

Myogenin expression (non-ischemic muscle)

A. 2 WT Fgf2-/- 1.5 FGF2 LMW-only FGF2 HMW-only

1

(arbitrary units) (arbitrary 0.5 Relative mRNA expression expression mRNA Relative

0

B. Myogenin expression (ischemic muscle) 8 WT 7 Fgf2-/- FGF2 LMW-only 6 FGF2 HMW-only

5

4

3 (arbitrary units) (arbitrary

2

mRNA Expression relative sham to relative Expression mRNA 1

0 1 3 7 14 Days after ischemia surgery 243

Figure 49. Myogenin mRNA expression in non-ischemic, sham and ischemic skeletal muscles.

(A) Quantitative RT-PCR analysis for Myogenin gene expression in non-ischemic skeletal muscle relative to WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes

relative to WT non-ischemic muscles (B) Time course of Myogenin mRNA expression. For

each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic

muscles are expressed relative to corresponding sham expression levels in WT (blue),

Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are presented as mean ± SEM. n=4 for each mouse group at each timepoint.

244

HMW FGF2 upregulates MRF4 expression in response to ischemia

MRF4 expression in non-ischemic WT, Fgf2-/-, FGF2 LMW-only and FGF2 HMW-only skeletal muscles was also examined. For all mouse groups, MRF4 mRNA expression was not different from WT (Figure 50A).

To determine the role of FGF2 isoform expression on MRF4 gene expression in response to ischemia, transcript levels were measured at 1, 3, 7 or 14 days of ischemia. Ischemic WT,

Fgf2-/-, and FGF2 LMW-only limbs had a similar pattern of expression in which MRF4 levels were significantly decreased by 3 days, increased at day 7 and returned to normal by day day14

(Figure 50B). Expression at 3 days of ischemia was significantly elevated in FGF2 HMW-only muscles compared to WT (p<0.05, Figure 50B). Furthermore, the pattern of expression over time in FGF2 HMW-only muscles differed from the other groups with highest expression of MRF4 occurring on 3 days of ischemia, decreasing levels by 7 days and a return to normal expression at

14 days.

Overall, the MRF expression data suggest that HMW FGF2 is important for maintaining a population in ischemic skeletal muscle. It also enhances the progression of the activated satellite cells to differentiated myoblasts and incorporation into repairing/regenerating muscle.

245

A. MRF4 expression (non-ischemic muscle) 2

WT Fgf2-/- 1.5 FGF2 LMW-only FGF2 HMW-only

1 (arbitrary units) (arbitrary Relative mRNA expression expression mRNA Relative 0.5

0 MRF4 expression (ischemic muscle) B. 5 WT Fgf2-/- FGF2 LMW-only 4 FGF2 HMW-only

3 # ‡

2 (arbitrary units) (arbitrary

1 mRNA Expression relative sham to relative Expression mRNA *

0 1 3 7 14 246 Days after ischemia surgery Figure 50. MRF4 mRNA expression in non-ischemic, sham and ischemic skeletal muscles.

(A) Quantitative RT-PCR analysis for MRF4 gene expression in non-ischemic skeletal muscle relative to WT (white bar), Fgf2-/- (light gray bar), FGF2 LMW-only (dark gray bar) or FGF2 HMW-only (black bar) transcript expression levels are represented as fold-changes relative to WT non-ischemic muscles (B) Time course of MRF4 mRNA expression. For each timepoint of ischemia (1, 3, 7, or 14 days), fold changes of mRNA expression in ischemic muscles are expressed relative to corresponding sham expression levels in WT (blue),

Fgf2-/- (red), FGF2 LMW-only (green), or FGF2 HMW-only (black) limbs. Data are presented as mean ± SEM. n=4 for each mouse group at each timepoint. *p<0.05 vs. Day 7 cohorts, #p<0.05 vs. WT, ‡p<0.05 vs. FGF2 LMW-only.

247

DISCUSSION III

The results presented in this chapter of the dissertation demonstrate that the protection of skeletal muscle from ischemia-induced damage by the endogenous HMW FGF2 isoforms is possibly mediated by muscle regeneration pathways (Figure 51).

Satellite cell activation, expansion and subsequent differentiation during embryogenesis or myogenic repair is tightly regulated by the Pax genes and the myogenic regulatory factors (MRFs)

(Buckingham & Relaix, 2007; Kang & Krauss, 2010; P. Zhao & Hoffman, 2004). To determine the role of the FGF2 isoforms in the regeneration potential of skeletal muscle during ischemia, the expression of the myogenic markers was determined in normal or ischemic muscles. The Pax genes, Pax3 and Pax7 are expressed by satellite cells (Chang & Rudnicki, 2014). Pax3 expression was not detected in any of the muscles analyzed. Pax3 is critical for embryonic myogenesis but is down-regulated in muscle before or at birth (Buckingham & Relaix, 2007; Rudnicki et al., 2008).

Pax3-/- mice die in utero (Buckingham & Relaix, 2007). Pax7 is dispensable for embryonic myogenesis but is essential after birth and is expressed by all quiescent satellite cells. Pax7-/- mice have complete loss of satellite cell pool, smaller-sized muscles and impaired skeletal muscle regeneration (Sambasivan et al., 2011; von Maltzahn, Jones, Parks, & Rudnicki, 2013). Pax7 was detected in FGF2 LMW-only and FGF2 HMW-only muscles and its expression was unchanged from WT in non-ischemic muscles (Figure 46A). This confirms that selective FGF2 isoform expression does not affect the satellite cell population. In response to ischemia, Pax7 expression in WT or LMW FGF2 muscles decreased from day 1 to undetectable levels at day 3 and was returned to normal level at day 14 (Figure 46B). In the presence of only HMW FGF2 however, the expression of Pax7 was elevated over four-fold. The population of Pax7-expressing cells at 14 days of ischemia represents satellite cell numbers in recovering muscle after cycles of self-renewal

248

and myogenic differentiation. This is possibly the true myogenic potential of the HMW FGF2

muscles which was enhanced by the presence of ischemia and muscle injury. Downstream of the

Pax transcription factors are the MRFs which are required for the commitment, progression and

differentiation of proliferating satellite cells and their progeny (myoblasts) into mature myofibers

(Kang & Krauss, 2010; Rudnicki et al., 2008; Yin et al., 2013). Pax3/Pax7 induce the expression

of two MRFS, Myf5 and MyoD by direct interaction with distal enhancer elements and proximal promoters of both genes.

Myf5 is another early myogenic marker expressed by satellite progenitor cells which also express Pax7 (Ceafalan et al., 2014; Yin et al., 2013). Expression of only LMW FGF2 or only

HMW FGF2 did not affect normal Myf5 expression (Figure 47A). After induction of ischemia,

Myf5 expression follows the pattern of Pax7 expression (Figure 47B). The increased myogenic potential of the HMW FGF2 muscles is evident as Myf5 expression is elevated and at a maximum by day 14. The addition of Myf5 expression marks early commitment of the activated satellite cells to the myogenic lineage; these cells leave the basal lamina niche and will eventually undergo differentiation (Cooper et al., 1999; Gayraud-Morel et al., 2007). Myf5 regulates the proliferation rate and homeostasis of satellite cells and is expressed by activated, proliferating myoblasts

(Gayraud-Morel et al., 2007; Rudnicki et al., 2008).

Similar to Myf5, MyoD expression is up-regulated in activated cells and expressed by proliferating myoblasts. Both genes have been termed myogenic determination factors (Cooper et al., 1999; Yin et al., 2013). HMW FGF2 lowered MyoD expression in non-ischemic muscle while

LMW FGF2 had no effect (Figure 48A). During ischemia, the MyoD is required for the differentiation potential of satellite cells. MyoD-/- mice have reduced muscle mass, myoblasts which continually self-renew, are differentiation-defective and do not fuse into myotubes

249

(Cornelison, Olwin, Rudnicki, & Wold, 2000; Le Grand & Rudnicki, 2007). The reduced MyoD

expression in normal muscle may be of limited importance as the absolute expression of MyoD in

myoblasts is less relevant than the Pax7-to-MyoD ratio. A high ratio is observed in non-activated

satellite cells and keep them in their quiescent state and cells with a low Pax7-to-MyoD ratio begin

to differentiate (Montarras, L’honoré, & Buckingham, 2013; Yin et al., 2013). This is evident in the ischemic expression of MyoD which is highest at Day 3 and coincides with the timepoint of lowest Pax7 expression (Figure 48B). The Pax7-to-MyoD ratio has shown to decrease further after activation of Myogenin (Chang & Rudnicki, 2014).

After limited rounds of proliferation, myoblasts withdraw from the cell cycle and terminal

differentiation is initiated by the expression of Myogenin and MRF4 which is prompted by the

transcriptional activity of MyoD (Kang & Krauss, 2010; Yin et al., 2013). Neither Myogenin nor

MRF4 was affected by expression of only HMW FGF2 or only LMW FGF2 in non-ischemic

muscle (Figures 47A and 48A). Myogenin and MRF4 drive the expression of several muscle-

specific structural and contractile proteins including sarcomeric myosin. The differentiating

myoblasts fuse to form elongated multinucleated myotubes and nascent myofibers. MRF4 also

appears to be necessary for hypertrophy of the new fibers. It is the only MRF that is expressed to a

high degree in mature myofibers and its increased expression is proposed to be associated with

downregulation of Myogenin (Gayraud-Morel et al., 2007; Yablonka-Reuveni, 2011). During

ischemia, Myogenin expression peaked at day 7 with a non-significant elevation at day 3 in the

HMW FGF2-expressing cells (Figure 49B). Ischemic expression of MRF4 followed a similar

pattern as Myogenin expression except for HMW FGF2 muscles whose expression peaked after 3

days and continued to decline (Figure 50B). These results suggest that the rate of myoblast

differentiation is much faster and more pronounced in the presence of HMW FGF2. This increased

250 myogenic potential of HMW FGF2 may underlie the complete protection from toe/limb necrosis observed in the FGF2 HMW-only mice as well as the rapid recovery from limb use impairment

(as discussed in Chapter 1). Another FGF, FGF6 plays an important in muscle repair and regeneration. Like FGF2, FGF6 is capable of stimulating rat satellite cell proliferation in vitro

(Sheehan & Allen, 1999). Fgf6-/- mice display impaired muscle regeneration linked to reduced numbers MyoD and Myogenin-expressing myoblasts and decreased migratory ability of the existing myoblasts (Floss et al., 1997; Neuhaus et al., 2003). The degree of migratory capability was further reduced in satellite cells-derived myoblasts isolated from Fgf2/Fgf6 double knockout mice (Neuhaus et al., 2003).

An important limitation to note is that the analysis of MRF gene expression was performed in whole muscle. This analysis while extremely useful has the potential to dilute the results and make subtle differences in expression of the transcriptions factors difficult to detect (Danoviz &

Yablonka-Reuveni, 2012; Yablonka-Reuveni et al., 2015). Satellite cells isolated from muscle tissue would be a useful alternative and may further highlight the effects of the HMW FGF2 isoforms on ischemia-induced muscle regeneration.

251

Figure 51. Schematic representation of the muscle regeneration pathway that may mediate the HMW FGF2 protection of ischemic muscle.

Inhibit

Activate Unknown interaction

252

CONCLUSIONS AND CLINICAL RELEVANCE

This dissertation provides significant advances in the understanding of the FGF2’s biological functions in vascular biology, in particular and for the first time, the in vivo actions of the HMW protein isoforms of FGF2. The results presented here unequivocally demonstrate that the HMW FGF2 isoforms are necessary and sufficient for inducing post-ischemic vascular growth, protecting skeletal muscle from damage and preserving functional capacity. The LMW isoform, conversely, was shown to have no effect on tissue survival during chronic hindlimb ischemia.

Furthermore, HMW FGF2 expression alone induced growth of capillaries (angiogenesis), small and medium-sized collateral vessels (arteriogenesis). The pathways which via which HMW FGF2 enhanced salvage of ischemic muscle were independent of FGF receptor activation and were associated with regulation of growth factor expression including FGF2, and other ischemia- triggered signals such as inflammation and myogenesis. Taken together, the findings presented here suggest a confluence of vascular remodeling and tissue repair/regeneration in HMW FGF2- induced recovery from chronic ischemia (Figure 52).

Since the discovery of FGF2 in 1984 as an angiogenic growth factor, only a handful of studies have examined the biological roles of the FGF2 isoforms as well as treated the two classes of FGF2 proteins as separate entities. The genesis of the isoform-specific knockout mouse models prompted studies that have started to show that there are differences between the FGF2 proteins

(Azhar et al., 2009). Three groups, in particular, have focused on researching the in vivo roles of the endogenous HMW isoforms distinct from LMW FGF2 in bone development and homeostasis

(Xiao et al., 2009, 2014, 2013), neuronal differentiation and degeneration (Claus, Bruns, & Grothe,

2004; Grothe, Schulze, et al., 2000; Haastert et al., 2006; Timmer et al., 2004) and cardiovascular pathophysiology (Liao et al., 2007, 2009, 2010; Manning et al., 2013). Previous studies conducted

253

by our laboratory have determined that the LMW isoform and the HMW isoforms have different

roles in acute ischemia-reperfusion injury and cardioprotection (Liao et al., 2007, 2009, 2010).

This dissertation offers, for the first time, important insights into the in vivo “angiogenic” functions

of the endogenous FGF2 isoforms, particularly the HMW isoforms which were previously

considered to act in a similar if not redundant fashion as LMW FGF2. The mitogenic effects of

HMW FGF2 in vascular biology have largely focused on the use of recombinant exogenous HMW

FGF2 in transformed and/or immortalized cells. While the later can provide valuable information

on the function of proteins, in vitro cell activity does not always depict the entire picture of the

protein's functions. This is the first study to identify the vascular effects of HMW FGF2 in vivo

and to demonstrate that the HMW isoforms are sufficient for ischemia-induced vascular growth,

both angiogenesis and arteriogenesis.

There is a large body of pre-clinical and clinical studies focusing on the therapeutic

potential of exogenous LMW FGF2 as described in the Introductory section (Tables 4-6). In summary, the multi-species animal studies provided adequate proof-of-concept data which were not translated into beneficial results in several randomized controlled trials. In this dissertation, the response to chronic ischemia included induction of angiogenesis (capillary growth) in WT,

Fgf2-/- and FGF2 HMW-only muscles as well as FGF2 LMW-only muscle. The extent of capillary growth was greater in the HMW FGF2-expressing limbs while Fgf2-/- mice were similar to WT.

FGF2 LMW-only ischemic muscles had no change in capillary number after 42 days of ischemia.

These results reveal antithetical functions of HMW and LMW FGF2 in vascular growth because ablation of both classes of FGF2 isoforms (LMW and HMW FGF2) in the Fgf2-/- mice has a milder phenotypic effect than is present in tissues expressing only LMW or only HMW FGF2.

Additionally, the absence of an angiogenic response in LMW FGF2-only mice suggests that this

254 endogenous isoform of FGF2 may exert an inhibitory effect on the angiogenic potential of HMW

FGF2 (in FGF2 LMW-only and WT muscles). This calls into question or, at the very least, warrants a re-examination of the previously identified role for endogenous LMW FGF2 as a vascular growth factor. Most important, however, is the clinical potential for the HMW isoforms of FGF2 to serve as robust inducers of revascularization in response chronic ischemia.

The FGF2 HMW-only muscles were protected from the toe/foot necrosis that was observed in WT, Fgf2-/-, and FGF2 LMW-only mice to an equal degree. Limb use impairment was also ameliorated in FGF2 HMW-only limbs and the recovery of normal function was faster than WT limbs. Recovery was present in LMW FGF2-expressing mice but occurred at a significantly slower rate (3x) than WT and FGF2 HMW-only. Fgf2-/- limbs did not recover full use of ischemic limbs.

Although both classes of FGF2 isoforms are capable of protecting skeletal muscle from ischemic injury, HMW FGF2 appears to be more efficient at this preserving limb viability. This is important because the salvage of limbs in animal models of chronic ischemia is used as a surrogate for a predictor of clinical success in humans (healing of lower extremity ulceration) (Cooke & Losordo,

2015; Dragneva et al., 2013; Waters et al., 2004).

In line with the bulk of the in vivo FGF2 studies, characterization of the downstream signaling mediated in response to FGF2 have focused on the cell surface activation of FGFRs by secreted or ECM-bound/released FGF2. Evidence from our laboratory and others has clearly shown that while LMW FGF2 is secreted during cellular stress, HMW FGF2 is largely confined to the nucleus (Chlebova et al., 2009; Jiang et al., 2009; Liao et al., 2009, 2010; Yu et al., 2007).

This localization of HMW FGF2 is critical for the cellular responses elicited by the protein isoform as evidenced when exogenously administered HMW FGF2 produces similar effects as LMW

FGF2 in cultured cells but transgenes expressing only the HMW isoforms produce differential

255

effects from LMW FGF2 expressing cells (Jiang et al., 2007; Kardami et al., 2007; Liao et al.,

2010). The absence of changes in FGFR activation in HMW FGF2-induced revascularization of

ischemic muscles suggests that the effects of this endogenous FGF2 isoform are intracrine in

nature. The nuclear localization of HMW FGF2, which is important for its protective effect in

chronic ischemia, adds a layer of complexity when considering the development of a HMW FGF2

drug. The HMW FGF2 therapeutic would have to actively get across two barriers, the plasma

membrane and nuclear envelop. Cell penetrating peptides like synthetic polyarginine or HIV-1 Tat

DNA binding domain have been successfully utilized to transport molecules into cells by an unknown mechanism (Biswas & Torchilin, 2014; Sakhrani & Padh, 2013). The inherent nuclear localization signals present on HMW FGF2 could be enhanced to ensure targeted delivery to the nucleus (Liao et al., 2009). Small peptides from viruses also that confer nuclear localization, such as the KKKRKV peptide from SV40 large T antigen may also be utilized to deliver the drug to the nucleus (Biswas & Torchilin, 2014).

Decades after the introduction of therapeutic revascularization as a potential treatment for coronary and peripheral artery disease, this intervention has yet to be successfully translated to the clinic and the need for alternative treatments persists. The research presented in this dissertation identifies a prospective new player in the field of therapeutic revascularization that has the

potential to be beneficial for patients suffering from critical limb and cardiac ischemia.

256

Figure 52. Schematic representing the hypothetical pathways activated and/or inhibited by

HMW FGF2 during adaptive revascularization and protection from ischemic muscle injury

Inhibit Activate

Unknown interaction

257

FUTURE DIRECTIONS

The studies presented herein suggest distinct roles for HMW and LMW FGF2 in regulating postnatal vessel growth/remodeling during ischemia. There remains, however, a number of important questions that still need to be addressed to better determine the roles of HMW and LMW

FGF2 in vascular growth and ischemic injury recovery. These studies may prove useful for the understanding of the molecular mechanisms by which the FGF2 isoforms regulates vascular development and promotes tissue repair in ischemic muscle as well as the potential development of HMW FGF2 as a new application for the treatment of ischemic vascular diseases. They include:

1) Role of FGF2 isoforms in muscle functional recovery: This dissertation showed that the

HMW isoforms accelerated complete functional recovery of ischemic limbs. This was

evaluated via an index of muscle function; ischemic hindlimb use scoring (Chalothorn et al.,

2007; Faber et al., 2011; Helisch et al., 2006; Peng et al., 2011). A different index of muscle

functional capacity, exercise performance measured with a treadmill endurance test, more

closely recapitulates the desired primary outcomes of PAD and CAD clinical trials (Dragneva

et al., 2013; Haas et al., 2012; Lavu, Gundewar, & Lefer, 2011). A preliminary study of this

endpoint in ischemic mice was reported in Chapter 1 and trends towards decreased functional

capacity were observed in the absence of all FGF2 expression (Fgf2-/-). In order to

conclusively determine the effect of selective FGF2 isoform expression on post-ischemic

exercise tolerance, increased sample sizes and the inclusion of additional timepoints before

and after 14 days of ischemia are necessary. A more robust examination of exercise

performance would be important for confirmation of these preliminary results and as a

secondary measure of muscle function recovery.

258

2) Role of vascular cell proliferation, and migration in the FGF2 isoform-specific response

to chronic ischemia: This dissertation showed that while HMW FGF2 augments ischemia-

induced capillary growth, LMW FGF2 inhibited angiogenesis in ischemic muscles.

Angiogenesis and arteriogenesis during embryonic development or in response to ischemia

requires the activation, proliferation, and migration of endothelial cells (ECs) and/or vascular

smooth muscle cells (vSMCs) (Carmeliet, 2000; Fischer et al., 2006). Differences in the

angiogenic capacity of HMW or LMW FGF2-expressing vascular cells may contribute to the

observed differences in FGF2 isoform stimulation of vessel growth To address this, ECs or

vSMCs isolated from FGF2 isoform knockout mice could be evaluated for proliferation and

migration using nanofiber peptide scaffold assays (an in vitro model of vascularity) (Cho et

al., 2012; Narmoneva et al., 2005). Additionally, in vivo assessment of microvascular cell

proliferation could be performed with immunohistochemical detection of proliferation markers

along with and EC or vSMC-specific markers can be evaluated.

3) Role of FGF2 isoforms in ischemia-induced myocyte death: An interesting question raised

by this dissertation research is whether the improvement in skeletal muscle functional recovery

and prevention of limb necrosis observed when only HMW FGF2 is expressed is related to the

degree of myocyte survival and/or regeneration. In ischemic muscle injury, myocyte death (via

necrosis or apoptosis) is followed by inflammatory cell accumulation and regeneration (Tidball

& Villalta, 2010; Tidball, 2011). The role of the FGF2 isoforms in the determining of the

degree myocyte survival and necrosis during chronic ischemia is still unknown.

4) Role of FGF2 isoforms in ischemia-induced myocyte regeneration: Preliminary results

presented in Chapter 2 of this dissertation suggest that the HMW isoforms have a greater

capacity for myocyte regeneration. Specifically, HMW FGF2 increased expression of satellite

259

cell activation and myoblast differentiation markers. These results were obtained from whole

muscle gene expression which is a limitation. Determination of myogenic gene expression in

satellite cells and myoblasts isolated from isoform knockout muscle tissue would confirm these

initial results and may further highlight the effects of the HMW FGF2 isoforms on ischemia-

induced muscle regeneration (Fukada et al., 2013; Sheehan & Allen, 1999; Yablonka-Reuveni

et al., 2015). In murine muscle, the large lobular nuclei of regenerating myofibers are centrally

located while healthy mature fibers have small peripheral nuclei (Paoni et al., 2002; Scharner

& Zammit, 2011). Histological analysis of this morphological marker of ongoing muscle

regeneration over time would allow for determination of the timing and extent of ischemic

muscle repair regulated by the FGF2 isoforms.

5) Role of inflammatory cell infiltration: Neutrophils and macrophages have been shown to

infiltrate ischemic muscle in hindlimb ischemia models (G. Brevetti, Giugliano, Brevetti, &

Hiatt, 2010; Silvestre et al., 2008; Tidball, 2005) and modulate the angiogenic/arteriogenic

response (Fung & Helisch, 2012; Naldini & Carraro, 2005; Silvestre et al., 2008). Preliminary

results presented here suggest that the total ablation of Fgf2 gene does have some effect on

neutrophil and macrophages infiltration after ischemia. Less clear, however, is the precise role

of the inflammatory microenvironment in FGF2 isoform-specific revascularization. The

absence of data on inflammatory cell infiltration between 0 and 3 days after ischemia is a

limitation. Neutrophil infiltration begins within a few hours of ischemic muscle injury and is

typically cleared by 3 days when monocyte infiltration starts to increase (Nguyen & Tidball,

2003; Tidball, 2011). Determination of the degree of inflammation as early as hours after the

induction of ischemia, the timing of resolution of this inflammation, and the subset of

inflammatory cells involved will provide useful information on their contribution to FGF2

260

isoform-induced vascular growth. As in Chapter 2, ischemic muscle would be evaluated for

immunohistochemical detection of myeloperoxidase (MPO) and Mac-3, markers for

neutrophils and macrophages, respectively. Additionally, the role of inflammation in ischemic-

induced revascularization could be examined by antibody depletion, genetic or

pharmacological ablation of specific inflammatory cell populations.

6) Role of FGFR signaling: Results presented in Chapter 2 of this dissertation suggest that FGFR

signaling is not associated with HMW FGF2-induced revascularization. To confirm the

relationship between the HMW isoforms and FGFR signaling in ischemia-induced

revascularization, the hindlimb ischemia studies (as described in Chapter 1) should be repeated

in the presence of FGFR inhibitors (Liao et al., 2010). Furthermore, it would be important to

determine the role of nuclear FGFR signaling as distinct from cell surface FGFR signaling in

HMW FGF2-mediated response to chronic ischemia.

7) Involvement of HSPG signaling in FGF2 isoform-induced vascular growth: In addition to

their roles as co-receptors for FGF-FGFR binding and , heparan sulfate

proteoglycans (HSPGs) are also capable of FGF ligand binding and signaling independent of

FGFR interactions (Murakami et al., 2008; Sarrazin et al., 2011). Syndecans are trans-

membrane HSPGs that regulate cell adhesion, proliferation, and migration in inflammation,

wound healing and angiogenesis (Couchman, Gopal, Lim, Nørgaard, & Multhaupt, 2015;

Murakami et al., 2008). The most widely studied syndecan, syndecan-4 is ubiquitously

expressed in many cell types including myofibroblasts, epithelial, endothelial, and smooth

muscle cells. Syndecan-4 null mice while viable, have defects in inflammatory responses to

wound healing and lipopolysaccharide injection. The role of syndecan-4 and other HSPGs in

FGF2 isoforms-mediated angiogenesis is not yet understood.

261

8) Potential function of HMW FGF2 as a transcription factor: As mentioned above,

preliminary results suggest that HMW FGF2-induced revascularization does not correlate with

activation of FGFR signaling. This fact coupled with the exclusive nuclear localization of

HMW FGF2 is suggestive of a role for the HMW isoforms as a direct modulator or co-factor

of gene transcription. Identification of DNA binding sites and other motifs present on the FGF2

isoforms may serve as protein-DNA or protein-protein interactions (interactome) and could

highlight putative targets of HMW FGF2 that mediate the vascular response to ischemia. These

interactomes can be detected by chromatin immunoprecipitation followed by sequencing

(ChIP–seq) assays. Confirmed partners of HMW FGF2 in other cell systems include the

survival of motor neurons protein (SMN), antiapoptotic protein 5 (Api5), and the ribosomal

protein L6/TAXREB 107 (Chlebova et al., 2009; Förthmann et al., 2015). In neuronal cells,

HMW FGF2 has been shown to interact with SMN and Api5 both anti-apoptotic proteins, to

potentiate cell survival and differentiation (Berghe et al., 2000; Claus, Bruns, et al., 2004;

Woodbury & Ikezu, 2013). It would be important to test if these interactions persist in and if

they promote HMW-FGF2 induced protection of myocytes and/or vascular cells.

9) Role of HMW FGF2 in other models of tissue injury: The chronic ischemia model described

in this dissertation has shown that HMW FGF2 is a promoter of functional recovery of

ischemic tissues. It would be interesting to explore if this protection extends to other chronic

ischemia models in the heart or brain. Studies are currently underway to determine the role of

HMW FGF2 isoforms in revascularization and protection from cardiac dysfunction in hearts

subjected to permanent LAD ligation to induce chronic myocardial ischemia.

10) Role of HMW FGF2 in models of ischemia with underlying compromised vascular

biology: Patients with CAD or PAD also have other risk factors including advanced age,

262

diabetes, endothelial dysfunction or stress, and hypercholesteremia that abrogate physiological

vascular growth (Dragneva et al., 2013; Haas et al., 2012). Models with forms of these human

risk factors would better mimic human ischemic disease and further characterize the

“angiogenic” efficacy of the HMW FGF2 isoforms. This can be done by crossing the FGF2

HMW-only mice with diet-induced or genetic models of diabetes or hyperlipidemia.

Additionally, the ischemia experiments could be performed in older HMW FGF2-expressing

mice.

11) Role of HMW FGF2 in tumor growth: In addition to their functions in vascular remodeling,

the FGF2 isoforms are implicated in tumor progression and . This could potentially

affect the use of the HMW isoforms as a revascularization therapeutic agent (Thomas-Mudge

et al., 2004). HMW FGF2 overexpression induces radioresistance in NIH3T3 fibroblasts and

HeLa cells (Cohen-Jonathan et al., 1997; Delrieu et al., 1998). The NIH3T3 cells had a

transformed morphology and induced tumor formation in nude mice (Quarto et al., 1991).

Conversely, bovine aortic ECs (BAECs) overexpressing HMW FGF2 had no effect on tumor

growth in nude mice but tumorigenesis did occur when all the FGF2 isoforms (LMW and

HMW) were overexpressed or when only the LMW isoform was overexpressed. Overall, these

studies indicate a potential role for the FGF2 isoforms in cancer development. These actions

of the FGF2 isoforms need to be taken into consideration if HMW FGF2 is to be used clinically

as a revascularization agent. Clinical trials in patients of PAD and CAD that received

recombinant LMW FGF2 did not report any malignancies that correlated with receipt of the

drug (Khurana & Simons, 2003; Lederman et al., 2002; Simons et al., 2002).

263

REFERENCES

Ahn, A., Frishman, W. H., Gutwein, A., Passeri, J., & Nelson, M. (2008). Therapeutic

angiogenesis: a new treatment approach for ischemic heart disease--Part II. in

Review, 16(4), 219–229.

Allen, R. E., & Boxhorn, L. K. (1989). Regulation of skeletal muscle satellite cell proliferation

and differentiation by transforming growth factor-beta, insulin-like growth factor I, and

fibroblast growth factor. Journal of Cellular Physiology, 138(2), 311–315.

Alvarez, Y., Alonso, M. T., Vendrell, V., Zelarayan, L. C., Chamero, P., Theil, T., … Schimmang,

T. (2003). Requirements for FGF3 and FGF10 during inner ear formation. Development

(Cambridge, England), 130(25), 6329–38.

Amann, K., Faulhaber, J., Campean, V., Balajew, V., Dono, R., Mall, G., & Ehmke, H. (2006).

Impaired myocardial capillarogenesis and increased adaptive capillary growth in FGF2-

deficient mice. Lab Invest, 86(1), 45–53.

Anderson, J. E., Liu, L., & Kardami, E. (1991). Distinctive patterns of basic fibroblast growth

factor (bFGF) distribution in degenerating and regenerating areas of dystrophic (mdx) striated

muscles. Developmental Biology, 147(1), 96–109.

Andrés, G., Leali, D., Mitola, S., Coltrini, D., Camozzi, M., Corsini, M., … Presta, M. (2009). A

pro-inflammatory signature mediates FGF2-induced angiogenesis. Journal of Cellular and

Molecular Medicine, 13(8B), 2083–108.

Annex, B. H. (2013). Therapeutic angiogenesis for critical limb ischaemia. Nature Reviews.

Cardiology, 10(7), 387–396. doi:10.1038/nrcardio.2013.70

264

Annex, B. H., & Simons, M. (2005). Growth factor-induced therapeutic angiogenesis in the heart:

protein therapy. Cardiovascular Research, 65(3), 649.

Antoine, M., Wirz, W., Tag, C. G., Mavituna, M., Emans, N., Korff, T., … Kiefer, P. (2005).

Expression pattern of fibroblast growth factors (FGFs), their receptors and antagonists in

primary endothelial cells and vascular smooth muscle cells. Growth Factors (Chur,

Switzerland), 23(2), 87–95.

Arese, M., Chen, Y., Florkiewicz, R. Z., Gualandris, A., Shen, B., & Rifkin, D. B. (1999). Nuclear

Activities of Basic Fibroblast Growth Factor: Potentiation of Low-Serum Growth Mediated

by Natural or Chimeric Nuclear Localization Signals. Molecular Biology of the Cell, 10(5),

1429–1444. doi:10.1091/mbc.10.5.1429

Armand, A.-S., Pariset, C., Laziz, I., Launay, T., Fiore, F., Della Gaspera, B., … Chanoine, C.

(2005). FGF6 regulates muscle differentiation through a calcineurin-dependent pathway in

regenerating soleus of adult mice. Journal of Cellular Physiology, 204(1), 297–308.

doi:10.1002/jcp.20302

Arras, M., Ito, W. D., Scholz, D., Winkler, B., Schaper, J., & Schaper, W. (1998). Monocyte

activation in angiogenesis and collateral growth in the rabbit hindlimb. The Journal of

Clinical Investigation, 101(1), 40–50.

Arroyo, A. G., & Iruela-Arispe, M. L. (2010). Extracellular matrix, inflammation, and the

angiogenic response. Cardiovascular Research, 86(2), 226–235.

Asahara, T. (1997). Isolation of Putative Progenitor Endothelial Cells for Angiogenesis. Science,

275(5302), 964–966.

265

Asahara, T., Bauters, C., Zheng, L. P., Takeshita, S., Bunting, S., Ferrara, N., … Isner, J. M.

(1995). Synergistic Effect of Vascular Endothelial Growth Factor and Basic Fibroblast

Growth Factor on Angiogenesis In Vivo. Circulation, 92(9), 365–371.

Auguste, P., Javerzat, S., & Bikfalvi, A. (2003). Regulation of vascular development by fibroblast

growth factors. Cell Tissue Res, 314(1), 157–166.

Aviles, R. J., Annex, B. H., & Lederman, R. J. (2003). Testing clinical therapeutic angiogenesis

using basic fibroblast growth factor (FGF-2). Br J Pharmacol, 140(4), 637–646.

Azhar, M., Yin, M., Zhou, M., Li, H., Mustafa, M., Nusayr, E., … Doetschman, T. (2009). Gene

targeted ablation of high molecular weight fibroblast growth factor-2. Developmental

Dynamics : An Official Publication of the American Association of Anatomists, 238(2), 351–

7.

Baffour, R., Berman, J., Garb, J. L., Rhee, S. W., Kaufman, J., & Friedmann, P. (1992). Enhanced

angiogenesis and growth of collaterals by in vivo administration of recombinant basic

fibroblast growth factor in a rabbit model of acute lower limb ischemia: Dose-response effect

of basic fibroblast growth factor. Journal of Vascular Surgery, 16(2), 181–191.

Baker, M., Robinson, S. D., Lechertier, T., Barber, P. R., Tavora, B., D’Amico, G., … Hodivala-

Dilke, K. (2012). Use of the mouse aortic ring assay to study angiogenesis. Nature Protocols,

7(1), 89–104.

Baron, O., Förthmann, B., Lee, Y.-W., Terranova, C., Ratzka, A., Stachowiak, E. K., …

Stachowiak, M. K. (2012). Cooperation of nuclear fibroblast growth factor receptor 1 and

Nurr1 offers new interactive mechanism in postmitotic development of mesencephalic

dopaminergic neurons. The Journal of Biological Chemistry, 287(24), 19827–40.

266

Baxter, R. C. (2013). Insulin-like growth factor binding protein-3 (IGFBP-3): Novel ligands

mediate unexpected functions. Journal of Cell Communication and Signaling, 7(3), 179–89.

Baxter, R. C. (2015). Nuclear actions of insulin-like growth factor binding protein-3. Gene, 569(1),

7–13.

Becker, L. C., Schuster, E. H., Jugdutt, B. I., Hutchins, G. M., & Bulkley, B. H. (1983).

Relationship between myocardial infarct size and occluded bed size in the dog: difference

between left anterior descending and circumflex coronary artery occlusions. Circulation,

67(3), 549–557.

Beenken, A., & Mohammadi, M. (2009). The FGF family: biology, pathophysiology and therapy.

Nature Reviews Drug Discovery, 8(3), 235–253. doi:10.1038/nrd2792

Belch, J., Hiatt, W. R., Baumgartner, I., Driver, I. V., Nikol, S., Norgren, L., & Van Belle, E.

(2011). Effect of fibroblast growth factor NV1FGF on amputation and death: a randomised

placebo-controlled trial of gene therapy in critical limb ischaemia. Lancet, 377(9781), 1929–

37. doi:10.1016/S0140-6736(11)60394-2

Belov, A. a, & Mohammadi, M. (2013). Molecular mechanisms of fibroblast growth factor

signaling in physiology and pathology. Cold Spring Harbor Perspectives in Biology, 5(6).

Berghe, L. Van den, Laurell, H., Huez, I., Zanibellato, C., Prats, H., & Bugler, B. (2000). FIF

[Fibroblast Growth Factor-2 (FGF-2)-Interacting-Factor], a Nuclear Putatively Antiapoptotic

Factor, Interacts Specifically with FGF-2. Molecular Endocrinology, 14(11), 1709–1724.

Bhora, F. Y., Dunkin, B. J., Batzri, S., Aly, H. M., Bass, B. L., Sidawy, A. N., & Harmon, J. W.

(1995). Effect of growth factors on cell proliferation and epithelialization in human skin. The

267

Journal of Surgical Research, 59(2), 236–44. doi:10.1006/jsre.1995.1160

Bikfalvi, A., Klein, S., Pintucci, G., Quarto, N., Mignatti, P., & Rifkin, D. B. (1995). Differential

modulation of cell phenotype by different molecular weight forms of basic fibroblast growth

factor: possible intracellular signaling by the high molecular weight forms. J Cell Biol,

129(1), 233–243.

Bishop, J. R., Schuksz, M., & Esko, J. D. (2007). Heparan sulphate proteoglycans fine-tune

mammalian physiology. Nature, 446(7139), 1030–7. doi:10.1038/nature05817

Biswas, S., & Torchilin, V. P. (2014). Nanopreparations for organelle-specific delivery in cancer.

Advanced Drug Delivery Reviews, 66, 26–41. doi:10.1016/j.addr.2013.11.004

Boerckel, J. D., Chandrasekharan, U. M., Waitkus, M. S., Tillmaand, E. G., Bartlett, R., Dicorleto,

P. E., … Fallis, A. . (2014). Mitogen-activated protein kinase phosphatase-1 promotes

neovascularization and angiogenic gene expression. Arteriosclerosis, Thrombosis, and

Vascular Biology, 34(5), 1020–31.

Böhlen, P., Esch, F., Baird, A., & Gospodarowicz, D. (1985). Acidic fibroblast growth factor

(FGF) from bovine brain: amino-terminal sequence and comparison with basic FGF. The

EMBO Journal, 4(8), 1951–6.

Bookout, A. L., Cummins, C. L., Kramer, M. F., Pesola, J. M., Mangelsdorf, D. J., Pesola, J. M.,

… Kramer, M. F. (2006). High-Throughput Real-Time Quantitative Reverse Transcription

PCR. Molecular Biology, 1–28.

Bosch-Marce, M., Okuyama, H., Wesley, J. B., Sarkar, K., Kimura, H., Liu, Y. V, … Semenza,

G. L. (2007). Effects of aging and hypoxia-inducible factor-1 activity on angiogenic cell

268

mobilization and recovery of perfusion after limb ischemia. Circulation Research, 101(12),

1310–8. doi:10.1161/CIRCRESAHA.107.153346

Bouxsein, M. L., Boyd, S. K., Christiansen, B. A., Guldberg, R. E., Jepsen, K. J., & Müller, R.

(2010a). Guidelines for assessment of bone microstructure in rodents using micro–computed

tomography. Journal of Bone and Mineral Research, 25(7), 1468–1486.

Bouxsein, M. L., Boyd, S. K., Christiansen, B. A., Guldberg, R. E., Jepsen, K. J., & Müller, R.

(2010b). Guidelines for assessment of bone microstructure in rodents using micro-computed

tomography. Journal of Bone and Mineral Research : The Official Journal of the American

Society for Bone and Mineral Research, 25(7), 1468–86.

Bovi, P. D., Curatola, A. M., Kern, F. G., Greco, A., Ittmann, M., & Basilico, C. (1987). An

oncogene isolated by transfection of Kaposi’s sarcoma DNA encodes a growth factor that is

a member of the FGF family. Cell, 50(5), 729–737. doi:10.1016/0092-8674(87)90331-X

Bradley, P. P., Priebat, D. A., Christensen, R. D., & Rothstein, G. (1982). Measurement of

Cutaneous Inflammation: Estimation of Neutrophil Content with an Enzyme Marker. Journal

of Investigative Dermatology, 78(3), 206–209. doi:10.1111/1523-1747.ep12506462

Brämer, G. R. (1988). International statistical classification of diseases and related health

problems. Tenth revision. World Health Statistics Quarterly. Rapport Trimestriel de

Statistiques Sanitaires Mondiales, 41(1), 32–36.

Brass, E. P. (2013). Intermittent claudication: new targets for drug development. Drugs, 73(10),

999–1014. doi:10.1007/s40265-013-0078-3

Brevetti, G., Giugliano, G., Brevetti, L., & Hiatt, W. R. (2010). Inflammation in peripheral artery

269

disease. Circulation, 122, 1862–1875. doi:10.1161/CIRCULATIONAHA.109.918417

Brevetti, L. S., Paek, R., Brady, S. E., Hoffman, J. I., Sarkar, R., & Messina, L. M. (2001).

Exercise-induced hyperemia unmasks regional blood flow deficit in experimental hindlimb

ischemia. The Journal of Surgical Research, 98(1), 21–6. doi:10.1006/jsre.2001.6161

Brogi, E., Winkles, J. a, Underwood, R., Clinton, S. K., Alberts, G. F., & Libby, P. (1993). Distinct

patterns of expression of fibroblast growth factors and their receptors in human atheroma and

nonatherosclerotic arteries. Association of acidic FGF with plaque microvessels and

macrophages. The Journal of Clinical Investigation, 92(5), 2408–18. doi:10.1172/JCI116847

Brown, M. D., Walter, H., Hansen-Smith, F. M., Hudlicka, O., Egginton, S., Hudlická, O., &

Egginton, S. (1998). Lack of involvement of basic fibroblast growth factor (FGF-2) in

capillary growth in skeletal muscles exposed to long-term contractile activity. Angiogenesis,

2(1), 81–91.

Buckingham, M., & Relaix, F. (2007). The role of Pax genes in the development of tissues and

organs: Pax3 and Pax7 regulate muscle progenitor cell functions. Annual Review of Cell and

Developmental Biology, 23, 645–73. doi:10.1146/annurev.cellbio.23.090506.123438

Bush, R. L., Pevec, W. C., Ndoye, A., Cheung, A. T. W., Sasse, J., & Pearson, D. N. (1998).

Regulation of new blood vessel growth into ischemic skeletal muscle. Journal of Vascular

Surgery, 28(5), 919–928. doi:10.1016/S0741-5214(98)70070-9

Butt, A. J., Martin, J. L., Dickson, K. A., McDougall, F., Firth, S. M., & Baxter, R. C. (2004).

Insulin-Like Growth Factor Binding Protein-3 Expression Is Associated with Growth

Stimulation of T47D Human Breast Cancer Cells: The Role of Altered Epidermal Growth

Factor Signaling. The Journal of Clinical Endocrinology & Metabolism, 89(4), 1950–1956.

270

Campuzano, R., Barrios, V., Cuevas, B., Asín-Cardiel, E., Muela, A., Castro, J. M., … Cuevas, P.

(2002). Serum basic fibroblast growth factor levels in exercise-induced myocardial ischemia

- more likely a marker of endothelial dysfunction than a marker of ischemia? European

Journal of Medical Research, 7(December 1999), 93–97.

Cao, R., Bråkenhielm, E., Pawliuk, R., Wariaro, D., Post, M. J., Wahlberg, E., … Cao, Y. (2003).

Angiogenic synergism, vascular stability and improvement of hind-limb ischemia by a

combination of PDGF-BB and FGF-2. Nature Medicine, 9(5), 604–13. doi:10.1038/nm

Cao, Y. (2009, November). Monotherapy versus combination therapy of angiogenic and

arteriogenic factors for the treatment of ischemic disorders. Current Molecular Medicine.

Cao, Y., Arbiser, J., D’Amato, R. J., D’Amore, P. A., Ingber, D. E., Kerbel, R., … Langer, R.

(2011). Forty-year journey of angiogenesis translational research. Science Translational

Medicine, 3(114), 114rv3.

Cao, Y., Hong, A., Schulten, H., & Post, M. J. (2005). Update on therapeutic neovascularization.

Cardiovasc Res., 65(3), 639–48.

Capoccia, B. J., Gregory, A. D., & Link, D. C. (2008). Recruitment of the inflammatory subset of

monocytes to sites of ischemia induces angiogenesis in a monocyte chemoattractant protein-

1-dependent fashion. Journal of Leukocyte Biology, 84(3), 760–768. doi:10.1189/jlb.1107756

Cardinal, T. R., & Hoying, J. B. (2007). A modified fluorescent microsphere-based approach for

determining resting and hyperemic blood flows in individual murine skeletal muscles. Vascul

Pharmacol, 47(1), 48–56.

Cardinal, T. R., Struthers, K. R., Kesler, T. J., Yocum, M. D., Kurjiaka, D. T., & Hoying, J. B.

271

(2011). Chronic hindlimb ischemia impairs functional vasodilation and vascular reactivity in

mouse feed arteries. Frontiers in Physiology, 2, 91. doi:10.3389/fphys.2011.00091

Carmeliet, P. (2000). Mechanisms of angiogenesis and arteriogenesis. Nat Med., 6(4), 389–95.

Carmeliet, P. (2003). Angiogenesis in health and disease. Nat Med, 9(6), 653–660.

Carmeliet, P., & Jain, R. K. (2011). Molecular mechanisms and clinical applications of

angiogenesis. Nature, 473(7347), 298–307.

Caron, A., Michelet, S., Caron, A., Sordello, S., Ivanov, M.-A., Delaère, P., … Emmanuel, F.

(2004). Human FGF-1 gene transfer promotes the formation of collateral vessels and

arterioles in ischemic muscles of hypercholesterolemic hamsters. The Journal of Gene

Medicine, 6(9), 1033–45. doi:10.1002/jgm.594

Ceafalan, L. C., Popescu, B. O., & Hinescu, M. E. (2014). Cellular players in skeletal muscle

regeneration. BioMed Research International, 2014, 957014. doi:10.1155/2014/957014

Chadi, G., Cao, Y., Pettersson, R. F., & Fuxe, K. (1994). Temporal and spatial increase of astroglial

basic fibroblast growth factor synthesis after 6-hydroxydopamine-induced degeneration of

the nigrostriatal dopamine neurons. Neuroscience, 61(4), 891–910.

Chadi, G., & Fuxe, K. (1998). Analysis of trophic responses in lesioned brain: focus on basic

fibroblast growth factor mechanisms. Brazilian Journal of Medical and Biological Research,

31(2), 231–241.

Chadi, G., & Gomide, V. C. (2004). FGF-2 and S100beta immunoreactivities increase in reactive

astrocytes, but not in microglia, in ascending dopamine pathways following a striatal 6-

OHDA-induced partial lesion of the nigrostriatal system. Cell Biology International, 28(12),

272

849–61. doi:10.1016/j.cellbi.2004.08.005

Chalothorn, D., Clayton, J. A., Zhang, H., Pomp, D., & Faber, J. E. (2007). Collateral density,

remodeling, and VEGF-A expression differ widely between mouse strains. Physiological

Genomics, 30(2), 179–91. doi:10.1152/physiolgenomics.00047.2007

Chalothorn, D., & Faber, J. E. (2010). Strain-dependent variation in collateral circulatory function

in mouse hindlimb. Physiological Genomics, 42(3), 469–79.

Chalothorn, D., Zhang, H., Clayton, J. A., Thomas, S. A., & Faber, J. E. (2005). Catecholamines

augment collateral vessel growth and angiogenesis in hindlimb ischemia. American Journal

of Physiology - Heart and Circulatory Physiology, 289(2), H947–59.

Chang, N. C., & Rudnicki, M. A. (2014). Satellite cells: the architects of skeletal muscle. Current

Topics in Developmental Biology, 107, 161–81. doi:10.1016/B978-0-12-416022-4.00006-8

Chen, C. H., Poucher, S. M., Lu, J., & Henry, P. D. (2004). Fibroblast growth factor 2: from

laboratory evidence to clinical application. Curr Vasc Pharmacol, 2(1), 33–43.

Chen, C.-C., & Lau, L. F. (2009). Functions and mechanisms of action of CCN matricellular

proteins. The International Journal of Biochemistry & Cell Biology, 41(4), 771–783.

Chen, Q., Jin, M., Yang, F., Zhu, J., Xiao, Q., & Zhang, L. (2013). Matrix metalloproteinases:

inflammatory regulators of cell behaviors in vascular formation and remodeling. Mediators

of Inflammation, 2013, 928315.

Chleboun, J. O., & Martins, R. N. (1994). The development and enhancement of the collateral

circulation in an animal model of lower limb ischaemia. Australian and New Zealand Journal

of Surgery, 64(3), 202–207.

273

Chlebova, K., Bryja, V., Dvorak, P., Kozubik, A., Wilcox, W. R., & Krejci, P. (2009). High

molecular weight FGF2: the biology of a nuclear growth factor. Cellular and Molecular Life

Sciences : CMLS, 66(2), 225–35.

Cho, H., Balaji, S., Sheikh, A. Q., Hurley, J. R., Tian, Y. F., Collier, J. H., … Narmoneva, D. A.

(2012). Regulation of endothelial cell activation and angiogenesis by injectable peptide

nanofibers. Acta Biomaterialia, 8(1), 154–64.

Claffey, K. P., Abrams, K., Shih, S.-C., Brown, L. F., Mullen, A., & Keough, M. (2001). Fibroblast

Growth Factor 2 Activation of Stromal Cell Vascular Endothelial Growth Factor Expression

and Angiogenesis. Laboratory Investigation, 81(1), 61–75. doi:10.1038/labinvest.3780212

Claus, P., Bruns, A.-F., & Grothe, C. (2004). Fibroblast growth factor-2(23) binds directly to the

survival of motoneuron protein and is associated with small nuclear RNAs. The Biochemical

Journal, 384(Pt 3), 559–65.

Claus, P., Doring, F., Gringel, S., Muller-Ostermeyer, F., Fuhlrott, J., Kraft, T., & Grothe, C.

(2003). Differential intranuclear localization of fibroblast growth factor-2 isoforms and

specific interaction with the survival of motoneuron protein. The Journal of Biological

Chemistry, 278(1), 479–485.

Claus, P., Werner, S., Timmer, M., & Grothe, C. (2004). Expression of the fibroblast growth

factor-2 isoforms and the FGF receptor 1-4 transcripts in the rat model system of Parkinson’s

disease. Neuroscience Letters, 360(3), 117–20. doi:10.1016/j.neulet.2004.01.046

Cleutjens, J. (1999). The infarcted myocardium Simply dead tissue, or a lively target for

therapeutic interventions. Cardiovascular Research, 44(2), 232–241.

274

Cochain, C., Channon, K. M., & Silvestre, J.-S. (2012). Angiogenesis in the Infarcted

Myocardium. Antioxidants & Redox Signaling, 18(9), 1100–1113.

Coffin, J. D., Florkiewicz, R. Z., Neumann, J., Mort-Hopkins, T., Dorn 2nd, G. W., Lightfoot, P.,

… Doetschman, T. (1995). Abnormal bone growth and selective translational regulation in

basic fibroblast growth factor (FGF-2) transgenic mice. Mol Biol Cell, 6(12), 1861–1873.

Cohen, M. V, Vernon, J., Yaghdjian, V., & Hatcher, V. B. (1994). Longitudinal changes in

myocardial basic fibroblast growth factor (FGF-2) activity following coronary artery ligation

in the dog. Journal of Molecular and Cellular Cardiology, 26(5), 683–690.

Cohen-Jonathan, E., Toulas, C., Monteil, S., Couderc, B., Maret, A., Bard, J. J., … Favre, G.

(1997). Radioresistance induced by the high molecular forms of the basic fibroblast growth

factor is associated with an increased G2 delay and a hyperphosphorylation of p34CDC2 in

HeLa cells. Cancer Research, 57(7), 1364–70.

Colleran, P. N., Li, Z., Yang, H.-T., Laughlin, M. H., & Terjung, R. L. (2010). Vasoresponsiveness

of collateral vessels in the rat hindlimb: influence of training. The Journal of Physiology,

588(Pt 8), 1293–307.

Comerota, A. J., Throm, R. C., Miller, K. A., Henry, T. D., Chronos, N. A., Laird, J., … Pilsudski,

R. (2002). Naked plasmid DNA encoding fibroblast growth factor type 1 for the treatment of

end-stage unreconstructible lower extremity ischemia: Preliminary results of a phase I trial.

Journal of Vascular Surgery, 35(5), 930–936.

Contreras-Shannon, V., Ochoa, O., Reyes-Reyna, S. M., Sun, D., Michalek, J. E., Kuziel, W. A.,

… Shireman, P. K. (2007). Fat accumulation with altered inflammation and regeneration in

skeletal muscle of CCR2-/- mice following ischemic injury. American Journal of Physiology.

275

Cell Physiology, 292(2), C953–67.

Cooke, J. P., & Losordo, D. W. (2015). Modulating the vascular response to limb ischemia:

angiogenic and cell therapies. Circulation Research, 116(9), 1561–78.

Cooper, R., Tajbakhsh, S., Mouly, V., Cossu, G., Buckingham, M., & Butler-Browne, G. (1999).

In vivo satellite cell activation via Myf5 and MyoD in regenerating mouse skeletal muscle. J.

Cell Sci., 112(17), 2895–2901.

Cordon-Cardo, C., Vlodavsky, I., Haimovitz-Friedman, A., Hicklin, D., & Fuks, Z. (1990).

Expression of basic fibroblast growth factor in normal human tissues. Laboratory

Investigation; a Journal of Technical Methods and Pathology, 63(6), 832–40.

Cornelison, D. D., Filla, M. S., Stanley, H. M., Rapraeger, A. C., & Olwin, B. B. (2001). Syndecan-

3 and syndecan-4 specifically mark skeletal muscle satellite cells and are implicated in

satellite cell maintenance and muscle regeneration. Developmental Biology, 239(1), 79–94.

Cornelison, D. D., Olwin, B. B., Rudnicki, M. A., & Wold, B. J. (2000). MyoD(-/-) satellite cells

in single-fiber culture are differentiation defective and MRF4 deficient. Developmental

Biology, 224(2), 122–37.

Couchman, J. R., Gopal, S., Lim, H. C., Nørgaard, S., & Multhaupt, H. A. B. (2015). Syndecans:

from peripheral coreceptors to mainstream regulators of cell behaviour. International Journal

of Experimental Pathology, 96(1), 1–10.

Couffinhal, T., Dufourcq, P., Barandon, L., Leroux, L., & Duplàa, C. (2009). Mouse models to

study angiogenesis in the context of cardiovascular diseases. Frontiers in Bioscience: A

Journal and Virtual Library, 14, 3310–3325.

276

Couffinhal, T., Silver, M., Zheng, L. P., Kearney, M., Witzenbichler, B., & Isner, J. M. (1998).

Mouse model of angiogenesis. Am J Pathol, 152(6), 1667–1679.

Cuevas, P., Barrios, V., Giménez-Gallego, G., Martinez-Coso, V., Cuevas, B., Benavides, J., …

Asin-Cardiel, E. (1997). Serum levels of basic fibroblast growth factor in acute myocardial

infarction. European Journal of Medical Research, 2(7), 282–4.

Danoviz, M. E., & Yablonka-Reuveni, Z. (2012). Skeletal muscle satellite cells: background and

methods for isolation and analysis in a primary culture system. Methods in Molecular Biology

(Clifton, N.J.), 798, 21–52.

Davis, M. G., Zhou, M., Ali, S., Coffin, J. D., Doetschman, T., & Dorn 2nd, G. W. (1997).

Intracrine and autocrine effects of basic fibroblast growth factor in vascular smooth muscle

cells. J Mol Cell Cardiol, 29(4), 1061–1072. de Lussanet, Q. G., van Golde, J. C. G., Beets-Tan, R. G. H., Post, M. J., Huijberts, M. S., Schaper,

N. C., … Backes, W. H. (2007). Dynamic contrast-enhanced MRI of muscle perfusion

combined with MR angiography of collateral artery growth in a femoral artery ligation model.

NMR in Biomedicine, 20(8), 717–725.

Deindl, E., Hoefer, I. E., Fernandez, B., Barancik, M., Heil, M., Strniskova, M., & Schaper, W.

(2003). Involvement of the Fibroblast Growth Factor System in Adaptive and Chemokine-

Induced Arteriogenesis. Circulation Research, 92(5), 561–568.

Dell’Era, P., Belleri, M., Stabile, H., Massardi, M. L., Ribatti, D., Presta, M., & Era, P. D. (2001).

Paracrine and autocrine effects of fibroblast growth factor-4 in endothelial cells. Oncogene,

20(21), 2655–63.

277

Delrieu, I., Arnaud, E., Ferjoux, G., Bayard, F., & Faye, J. C. (1998). Overexpression of the FGF-

2 24-kDa isoform up-regulates IL-6 transcription in NIH-3T3 cells. FEBS Letters, 436(1),

17–22.

Deng, C. X., Wynshaw-Boris, A., Shen, M. M., Daugherty, C., Ornitz, D. M., & Leder, P. (1994).

Murine FGFR-1 is required for early postimplantation growth and axial organization. Genes

& Development, 8(24), 3045–57.

Deroanne, C. F., Hajitou, A., Nusgens, B. V., Lapiã, C. M., Calberg-Bacq, C.-M., & Lapiere, C.

M. (1997). Angiogenesis by Fibroblast Growth Factor 4 Is Mediated through an Autocrine

Up-Regulation of Vascular Endothelial Growth Factor Expression Angiogenesis by

Fibroblast Growth Factor 4 Is Mediated through an Autocrine Up-Regulation of Vascular

Endothelial Gr. Cancer Res., 57(24), 5590–5597.

Deveci, D., & Egginton, S. (1999). Development of the Fluorescent Microsphere Technique for

Quantifying Regional Blood Flow in Small Mammals. Experimental Physiology, 84(4), 615–

630.

DiMario, J., Buffinger, N., Yamada, S., & Strohman, R. C. (1989). Fibroblast growth factor in the

extracellular matrix of dystrophic (mdx) mouse muscle. Science (New York, N.Y.), 244(4905),

688–90.

Ding, L., Doñate, F., Parry, G. C. N., Guan, X., Maher, P. A., Levin, E. G., & Donate, F. (2002).

Inhibition of cell migration and angiogenesis by the amino-terminal fragment of 24kD basic

fibroblast growth factor. The Journal of Biological Chemistry, 277(34), 31056–61.

Dokun, A. O., Keum, S., Hazarika, S., Li, Y., Lamonte, G. M., Wheeler, F., … Annex, B. H.

(2008). A quantitative trait (LSq-1) on mouse chromosome 7 is linked to the absence

278

of tissue loss after surgical hindlimb ischemia. Circulation, 117(9), 1207–1215.

Dono, R., Texido, G., Dussel, R., Ehmke, H., & Zeller, R. (1998). Impaired cerebral cortex

development and blood pressure regulation in FGF-2-deficient mice. The EMBO Journal,

17(15), 4213–25. doi:10.1093/emboj/17.15.4213

Doukas, J., Blease, K., Craig, D., Ma, C., Chandler, L. A., Sosnowski, B. A., & Pierce, G. F.

(2002). Delivery of FGF genes to wound repair cells enhances arteriogenesis and myogenesis

in skeletal muscle. Molecular Therapy : The Journal of the American Society of Gene

Therapy, 5(5 Pt 1), 517–27.

Downey, C. M., Singla, A. K., Villemaire, M. L., Buie, H. R., Boyd, S. K., & Jirik, F. R. (2012).

Quantitative Ex-Vivo Micro-Computed Tomographic Imaging of Blood Vessels and Necrotic

Regions within Tumors. PLoS ONE, 7(7).

Dragneva, G., Korpisalo, P., & Ylä-Herttuala, S. (2013). Promoting blood vessel growth in

ischemic diseases: challenges in translating preclinical potential into clinical success. Disease

Models & Mechanisms, 6(2), 312–322. doi:10.1242/dmm.010413

Du, X., Xie, Y., Xian, C. J., & Chen, L. (2012). Role of FGFs/FGFRs in skeletal development and

bone regeneration. Journal of Cellular Physiology, 227(12), 3731–3743.

Dumont, N. a, Wang, Y. X., & Rudnicki, M. a. (2015). Intrinsic and extrinsic mechanisms

regulating satellite cell function. Development (Cambridge, England), 142(9), 1572–1581.

Dunham-Ems, S. M., Lee, Y.-W., Stachowiak, E. K., Pudavar, H., Claus, P., Prasad, P. N., &

Stachowiak, M. K. (2009). Fibroblast growth factor receptor-1 (FGFR1) nuclear dynamics

reveal a novel mechanism in transcription control. Molecular Biology of the Cell, 20(9),

279

2401–12.

Duprey-Díaz, M. V, Blagburn, J. M., & Blanco, R. E. (2012). Changes in fibroblast growth factor-

2 and FGF receptors in the frog visual system during optic nerve regeneration. Journal of

Chemical Neuroanatomy, 46(1-2), 35–44. doi:10.1016/j.jchemneu.2012.08.003

Düsterhöft, S., Putman, C. T., & Pette, D. (1999). Changes in FGF and FGF receptor expression

in low-frequency-stimulated rat muscles and rat satellite cell cultures. Differentiation;

Research in Biological Diversity, 65(4), 203–8. doi:10.1046/j.1432-0436.1999.6540203.x

Duvall, C. L., Taylor, W. R., Weiss, D., & Guldberg, R. E. (2004). Quantitative microcomputed

tomography analysis of collateral vessel development after ischemic injury. American

Journal of Physiology - Heart and Circulatory Physiology, 287(1), H302–10.

Duvall, C. L., Weiss, D., Robinson, S. T., Alameddine, F. M. F., Guldberg, R. E., & Taylor, W.

R. (2008). The role of osteopontin in recovery from hind limb ischemia. Arteriosclerosis,

Thrombosis, and Vascular Biology, 28(2), 290–5. doi:10.1161/ATVBAHA.107.158485

Eash, J., Olsen, A., Breur, G., Gerrard, D., & Hannon, K. (2007). FGFR1 inhibits skeletal muscle

atrophy associated with hindlimb suspension. BMC Musculoskeletal Disorders, 8, 32.

Eklund, L., & Saharinen, P. (2013). Angiopoietin signaling in the vasculature. Experimental Cell

Research, 319(9), 1271–80.

Elfenbein, A., & Simons, M. (2013). Syndecan-4 signaling at a glance. J Cell Sci, 126(17), 3799–

3804. doi:10.1242/jcs.124636

Elmadhun, N. Y., Sabe, A. A., Robich, M. P., Chu, L. M., Lassaletta, A. D., & Sellke, F. W. (2013).

The pig as a valuable model for testing the effect of resveratrol to prevent cardiovascular

280

disease. Annals of the New York Academy of Sciences, 1290(1), 130–135.

Endoh, M., Pulsinelli, W. A., & Wagner, J. A. (1994). Transient global ischemia induces dynamic

changes in the expression of bFGF and the FGF receptor. Molecular Brain Research, 22(1-

4), 76–88.

Esch, F., Baird, A., Ling, N., Ueno, N., Hill, F., Denoroy, L., … Guillemin, R. (1985). Primary

structure of bovine pituitary basic fibroblast growth factor (FGF) and comparison with the

amino-terminal sequence of bovine brain acidic FGF. Proceedings of the National Academy

of Sciences, 82(19), 6507–6511. doi:10.1073/pnas.82.19.6507

Estival, A., Monzat, V., Miquel, K., Gaubert, F., Hollande, E., Korc, M., … Clemente, F. (1996).

Differential Regulation of Fibroblast Growth Factor (FGF) Receptor-1 mRNA and Protein

by Two Molecular Forms of Basic FGF. Journal of Biological Chemistry, 271(10), 5663–

5670.

Eswarakumar, V. P., Lax, I., & Schlessinger, J. (2005). Cellular signaling by fibroblast growth

factor receptors. Cytokine & Growth Factor Reviews, 16(2), 139–149.

Faber, J. E., Chilian, W. M., Deindl, E., Royen, N. van, & Simons, M. (2014). A Brief Etymology

of the Collateral Circulation. Arterioscler Thromb Vasc Biol, ATVBAHA.114.303929.

Faber, J. E., Zhang, H., Lassance-Soares, R. M., Prabhakar, P., Najafi, A. H., Burnett, M. S., &

Epstein, S. E. (2011). Aging causes collateral rarefaction and increased severity of ischemic

injury in multiple tissues. Arteriosclerosis, Thrombosis, and Vascular Biology, 31(8), 1748–

1756.

Fagiani, E., & Christofori, G. (2013). Angiopoietins in angiogenesis. Cancer Letters, 328(1), 18–

281

26. doi:10.1016/j.canlet.2012.08.018

Falk, E., Nakano, M., Bentzon, J. F., Finn, A. V, & Virmani, R. (2013). Update on acute coronary

syndromes: the pathologists’ view. European Heart Journal, 34(10), 719–28.

Fang, X., Stachowiak, E. K., Dunham-Ems, S. M., Klejbor, I., & Stachowiak, M. K. (2005).

Control of CREB-binding protein signaling by nuclear fibroblast growth factor receptor-1: a

novel mechanism of gene regulation. The Journal of Biological Chemistry, 280(31), 28451–

62. doi:10.1074/jbc.M504400200

Fataccioli, V., Abergel, V., Wingertsmann, L., Neuville, P., Spitz, E., Adnot, S., … Teiger, E.

(2002). Stimulation of Angiogenesis by Cyr61 Gene: A New Therapeutic Candidate. Human

Gene Therapy, 13(12), 1461–1470.

Fei, Y., Gronowicz, G., & Hurley, M. M. (2013). Fibroblast Growth Factor-2, Bone Homeostasis

and Fracture Repair. Current Pharmaceutical Design, 19(19), 3354–3363.

Feldman, B., Poueymirou, W., Papaioannou, V. E., DeChiara, T. M., & Goldfarb, M. (1995).

Requirement of FGF-4 for postimplantation mouse development. Science (New York, N.Y.),

267(5195), 246–9.

Fewell, J. G., Osinska, H., Klevitsky, R., Ng, W., Sfyris, G., Bahrehmand, F., & Robbins, J. (1997).

A treadmill exercise regimen for identifying cardiovascular phenotypes in transgenic mice.

Am J Physiol, 273(3 Pt 2), H1595–605.

Fischer, C., Schneider, M., & Carmeliet, P. (2006). Principles and Therapeutic Implications of

Angiogenesis, Vasculogenesis and Arteriogenesis. In S. Moncada & A. Higgs (Eds.), The

Vascular Endothelium II (pp. 157–212).

282

Florkiewicz, R. Z., Anchin, J., & Baird, A. (1998). The Inhibition of Fibroblast Growth Factor-2

Export by Cardenolides Implies a Novel Function for the Catalytic Subunit of Na+,K+-

ATPase. Journal of Biological Chemistry, 273(1), 544–551. doi:10.1074/jbc.273.1.544

Florkiewicz, R. Z., Baird, A., & Gonzalez, A. M. (1991). Multiple forms of bFGF: differential

nuclear and cell surface localization. Growth Factors (Chur, Switzerland), 4(4), 265–75.

Florkiewicz, R. Z., & Sommer, A. (1989). Human basic fibroblast growth factor gene encodes four

polypeptides: three initiate translation from non-AUG codons. Proc Natl Acad Sci U S A,

86(11), 3978–3981.

Floss, T., Arnold, H. H., & Braun, T. (1997). A role for FGF-6 in skeletal muscle regeneration.

Genes and Development, 11(16), 2040–2051. doi:10.1101/gad.11.16.2040

Folkman, J., Klagsbrun, M., Sasse, J., Wadzinski, M., Ingber, D. E., & Vlodavsky, I. (1988). A

heparin-binding angiogenic protein--basic fibroblast growth factor--is stored within basement

membrane. The American Journal of Pathology, 130(2), 393–400.

Folkman, J., Merler, E., Abernathy, C., & Williams, G. (1971). Isolation of a tumor factor

responsible for angiogenesis. The Journal of Experimental Medicine, 133(2), 275–88.

Fon Tacer, K., Bookout, A. L., Ding, X., Kurosu, H., John, G. B., Wang, L., … Kliewer, S. A.

(2010). Research resource: Comprehensive expression atlas of the fibroblast growth factor

system in adult mouse. Molecular Endocrinology (Baltimore, Md.), 24(10), 2050–64.

Fontaine, V., Filipe, C., Werner, N., Gourdy, P., Billon, A., Garmy-Susini, B., … Arnal, J.-F.

(2006). Essential role of bone marrow fibroblast growth factor-2 in the effect of estradiol on

reendothelialization and endothelial progenitor cell mobilization. The American Journal of

283

Pathology, 169(5), 1855–1862.

Förthmann, B., Grothe, C., & Claus, P. (2015). A nuclear odyssey: fibroblast growth factor-2

(FGF-2) as a regulator of nuclear homeostasis in the nervous system. Cellular and Molecular

Life Sciences, 2.

Förthmann, B., van Bergeijk, J., Lee, Y.-W., Lübben, V., Schill, Y., Brinkmann, H., … Claus, P.

(2013). Regulation of neuronal differentiation by proteins associated with nuclear bodies.

PloS One, 8(12), e82871.

Frangogiannis, N. G., & Entman, M. L. (2005). Chemokines in myocardial ischemia. Trends in

Cardiovascular Medicine, 15(5), 163–169. doi:10.1016/j.tcm.2005.06.005

Frangogiannis, N. G., Smith, C. W., & Entman, M. L. (2002). The inflammatory response in

myocardial infarction. Cardiovascular Research, 53(1), 31–47.

Frantz, S., Vincent, K. A., Feron, O., & Kelly, R. A. (2005). Innate immunity and angiogenesis.

Circulation Research, 96(1), 15–26.

Fujii, T., Yonemitsu, Y., Onimaru, M., Inoue, M., Hasegawa, M., Kuwano, H., & Sueishi, K.

(2008). VEGF function for upregulation of endogenous PlGF expression during FGF-2-

mediated therapeutic angiogenesis. Atherosclerosis, 200(1), 51–7.

Fujii, T., Yonemitsu, Y., Onimaru, M., Tanii, M., Nakano, T., Egashira, K., … Sueishi, K. (2006).

Nonendothelial mesenchymal cell-derived MCP-1 is required for FGF-2-mediated

therapeutic neovascularization: critical role of the inflammatory/arteriogenic pathway.

Arteriosclerosis, Thrombosis, and Vascular Biology, 26(11), 2483–2489.

Fujiu, K., & Nagai, R. (2014). Fibroblast-mediated pathways in cardiac hypertrophy. Journal of

284

Molecular and Cellular Cardiology. doi:10.1016/j.yjmcc.2014.01.013

Fukada, S., Ma, Y., Ohtani, T., Watanabe, Y., Murakami, S., & Yamaguchi, M. (2013). Isolation,

characterization, and molecular regulation of muscle stem cells. Frontiers in Striated Muscle

Physiology, 4, 317.

Fung, E., & Helisch, A. (2012). Macrophages in collateral arteriogenesis. Frontiers in Physiology,

3, 353. doi:10.3389/fphys.2012.00353

Gao, M. H., Lai, N. C., McKirnan, M. D., Roth, D. A., Rubanyi, G. M., Dalton, N., … Hammond,

H. K. (2004). Increased regional function and perfusion after intracoronary delivery of

adenovirus encoding fibroblast growth factor 4: report of preclinical data. Human Gene

Therapy, 15(6), 574–87.

Garmy-Susini, B., Delmas, E., Gourdy, P., Zhou, M., Bossard, C., Bugler, B., … Arnal, J. F.

(2004). Role of fibroblast growth factor-2 isoforms in the effect of estradiol on endothelial

cell migration and proliferation. Circulation Research, 94(10), 1301–9.

Gaubert, F., Escaffit, F., Bertrand, C., Korc, M., Pradayrol, L., Clemente, F., & Estival, A. (2001).

Expression of the high molecular weight fibroblast growth factor-2 isoform of 210 amino

acids is associated with modulation of protein kinases C delta and epsilon and ERK activation.

The Journal of Biological Chemistry, 276(2), 1545–54. doi:10.1074/jbc.M001184200

Gayraud-Morel, B., Chrétien, F., Flamant, P., Gomès, D., Zammit, P. S., & Tajbakhsh, S. (2007).

A role for the myogenic determination gene Myf5 in adult regenerative myogenesis.

Developmental Biology, 312(1), 13–28. doi:10.1016/j.ydbio.2007.08.059

Giordano, F. J., Ping, P., McKirnan, M. D., Nozaki, S., Demaria, A. N., Dillmann, W. H., …

285

Hammond, H. K. (1996). Intracoronary gene transfer of fibroblast growth factor–5 increases

blood flow and contractile function in an ischemic region of the heart. Nature Medicine, 2(5),

534–539.

Giordano, S., Sherman, L., Lyman, W., & Morrison, R. (1992). Multiple molecular weight forms

of basic fibroblast growth factor are developmentally regulated in the central nervous system.

Developmental Biology, 152(2), 293–303.

Goldfarb, M. (2005). Fibroblast growth factor homologous factors: evolution, structure, and

function. Cytokine Growth Factor Rev, 16(2), 215–220.

Gonzalez, A. M. (1990). Distribution of basic fibroblast growth factor in the 18-day rat fetus:

localization in the basement membranes of diverse tissues. The Journal of Cell Biology,

110(3), 753–765. doi:10.1083/jcb.110.3.753

Gospodarowicz, D. (1975). Purification of a fibroblast growth factor from bovine pituitary.

Journal of Biological Chemistry, 250(7), 2515–2520.

Gospodarowicz, D., Bialecki, H., & Greenburg, G. (1978). Purification of the fibroblast growth

factor activity from bovine brain. Journal of Biological Chemistry, 253(10), 3736–3743.

Greenbaum, D., Colangelo, C., Williams, K., & Gerstein, M. (2003). Comparing protein

abundance and mRNA expression levels on a genomic scale. Genome Biology, 4(9), 117.

Grines, C. L., Watkins, M. W., Helmer, G., Penny, W., Brinker, J., Marmur, J. D., … Engler, R.

L. (2002). Angiogenic Gene Therapy (AGENT) trial in patients with stable angina pectoris.

Circulation, 105(11), 1291–1297.

Grines, C. L., Watkins, M. W., Mahmarian, J. J., Iskandrian, A. E., Rade, J. J., Marrott, P., …

286

Kleiman, N. (2003). A randomized, double-blind, placebo-controlled trial of Ad5FGF-4 gene

therapy and its effect on myocardial perfusion in patients with stable angina. Journal of the

American College of Cardiology, 42(8), 1339–1347.

Grochot-Przeczek, A., Dulak, J., & Jozkowicz, A. (2013). Therapeutic angiogenesis for

revascularization in peripheral artery disease. Gene, 525(2), 220–228.

Grothe, C., Haastert, K., & Jungnickel, J. (2006). Physiological function and putative therapeutic

impact of the FGF-2 system in peripheral nerve regeneration--lessons from in vivo studies in

mice and rats. Brain Research Reviews, 51(2), 293–9. doi:10.1016/j.brainresrev.2005.12.001

Grothe, C., Heese, K., Meisinger, C., Wewetzer, K., Kunz, D., Cattini, P. A., & Otten, U. (2000).

Expression of interleukin-6 and its receptor in the sciatic nerve and cultured Schwann cells:

relation to 18-kD fibroblast growth factor-2. Brain Research, 885(2), 172–181.

Grothe, C., Meisinger, C., & Claus, P. (2001). In vivo expression and localization of the fibroblast

growth factor system in the intact and lesioned rat peripheral nerve and spinal ganglia. The

Journal of Comparative Neurology, 434(3), 342–357.

Grothe, C., Meisinger, C., Hertenstein, A., Kurz, H., & Wewetzer, K. (1996). Expression of

fibroblast growth factor-2 and fibroblast growth factor receptor 1 messenger RNAs in spinal

ganglia and sciatic nerve: regulation after peripheral nerve lesion. Neuroscience, 76(1), 123–

135. doi:10.1016/S0306-4522(96)00355-7

Grothe, C., Meisinger, C., Holzschuh, J., Wewetzer, K., & Cattini, P. A. (1998). Over-expression

of the 18 kD and 21/23 kD fibroblast growth factor-2 isoforms in PC12 cells and Schwann

cells results in altered cell morphology and growth. Brain Research. Molecular Brain

Research, 57(1), 97–105.

287

Grothe, C., Schulze, A., Semkova, I., Müller-Ostermeyer, F., Rege, A., & Wewetzer, K. (2000).

The high molecular weight fibroblast growth factor-2 isoforms (21,000mol. wt and

23,000mol. wt) mediate neurotrophic activity on rat embryonic mesencephalic dopaminergic

neurons in vitro. Neuroscience, 100(1), 73–86.

Gu, J. W., Santiago, D., Olowe, Y., & Weinberger, J. (1997). Basic fibroblast growth factor as a

biochemical marker of exercise- induced ischemia. Circulation, 95(5), 1165–1168.

Gualandris, A., Rusnati, M., Belleri, M., Nelli, E. E., Bastaki, M., Molinari-Tosatti, M. P., …

Presta, M. (1996). Basic fibroblast growth factor overexpression in endothelial cells: an

autocrine mechanism for angiogenesis and angioproliferative diseases. Cell Growth &

Differentiation : The Molecular Biology Journal of the American Association for Cancer

Research, 7(2), 147–60.

Gualandris, A., Urbinati, C., Rusnati, M., Ziche, M., & Presta, M. (1994). Interaction of high-

molecular-weight basic fibroblast growth factor with endothelium: biological activity and

intracellular fate of human recombinant M(r) 24,000 bFGF. J Cell Physiol, 161(1), 149–159.

Haas, T. L., Lloyd, P. G., Yang, H.-T., & Terjung, R. L. (2012). Exercise training and peripheral

arterial disease. Comprehensive Physiology, 2(4), 2933–3017. doi:10.1002/cphy.c110065

Haase, I., Evans, R., Pofahl, R., & Watt, F. M. (2003). Regulation of keratinocyte shape, migration

and wound epithelialization by IGF-1- and EGF-dependent signalling pathways. Journal of

Cell Science, 116(Pt 15), 3227–38.

Haastert, K., Lipokatic, E., Fischer, M., Timmer, M., & Grothe, C. (2006). Differentially promoted

peripheral nerve regeneration by grafted Schwann cells over-expressing different FGF-2

isoforms. Neurobiology of Disease, 21(1), 138–53. doi:10.1016/j.nbd.2005.06.020

288

Hampton, T. G., Amende, I., Fong, J., Laubach, V. E., Li, J., Metais, C., & Simons, M. (2000).

Basic FGF reduces stunning via a NOS2-dependent pathway in coronary-perfused mouse

hearts. American Journal of Physiology. Heart and Circulatory Physiology, 279(1), H260–8.

Han, X., Xiao, Z., & Quarles, L. D. (2015). Membrane and Integrative Nuclear Fibroblastic

Growth Factor Receptor (FGFR) Regulation of FGF-23. The Journal of Biological

Chemistry, 290(16), 10447–59. doi:10.1074/jbc.M114.609230

Hao, X., Månsson-Broberg, A., Gustafsson, T., Grinnemo, K. H., Blomberg, P., Siddiqui, A. J.,

… Sylvén, C. (2004). Angiogenic effects of dual gene transfer of bFGF and PDGF-BB after

myocardial infarction. Biochemical and Biophysical Research Communications, 315(4),

1058–63. doi:10.1016/j.bbrc.2004.01.165

Hasdai, D., Barak, V., Leibovitz, E., Herz, I., Sclarovsky, S., Eldar, M., & Scheinowitz, M. (1997).

Serum basic fibroblast growth factor levels in patients with ischemic heart disease.

International Journal of Cardiology, 59(2), 133–8.

Haub, O., Drucker, B., & Goldfarb, M. (1990). Expression of the murine fibroblast growth factor

5 gene in the adult central nervous system. Proceedings of the National Academy of Sciences

of the United States of America, 87(20), 8022–6.

Haub, O., & Goldfarb, M. (1991). Expression of the fibroblast growth factor-5 gene in the mouse

embryo. Development (Cambridge, England), 112(2), 397–406.

Hearse, D. J. (2000). The elusive coypu: the importance of collateral flow and the search for an

alternative to the dog. Cardiovascular Research, 45(1), 215–219.

Hébert, J. M., Rosenquist, T., Götz, J., & Martin, G. R. (1994). FGF5 as a regulator of the hair

289

growth cycle: Evidence from targeted and spontaneous mutations. Cell, 78(6), 1017–1025.

Hedrich, H. J., & Bullock, G. R. (Eds.). (2004). The . Amsterdam: Elsevier

Academic Press.

Heidbreder, M., Fröhlich, F., Jöhren, O., Dendorfer, A., Qadri, F., & Dominiak, P. (2003). Hypoxia

rapidly activates HIF-3alpha mRNA expression. FASEB Journal : Official Publication of the

Federation of American Societies for Experimental Biology, 17(11), 1541–3.

Heil, M., Eitenmüller, I., Schmitz-Rixen, T., & Schaper, W. (2006). Arteriogenesis versus

angiogenesis: similarities and differences. Journal of Cellular and Molecular Medicine,

10(1), 45–55.

Heil, M., & Schaper, W. (2004). Influence of mechanical, cellular, and molecular factors on

collateral artery growth (arteriogenesis). Circulation Research, 95(5), 449–58.

Heil, M., & Schaper, W. (2007). Insights into pathways of arteriogenesis. Current Pharmaceutical

Biotechnology, 8(1), 35–42.

Heil, M., Ziegelhoeffer, T., Pipp, F., Kostin, S., Martin, S., Clauss, M., & Schaper, W. (2002).

Blood monocyte concentration is critical for enhancement of collateral artery growth.

American Journal of Physiology. Heart and Circulatory Physiology, 283(6), H2411–9.

Helisch, A., Wagner, S., Khan, N., Drinane, M., Wolfram, S., Heil, M., … Schaper, W. (2006).

Impact of Mouse Strain Differences in Innate Hindlimb Collateral Vasculature. Arterioscler

Thromb Vasc Biol, 26(3), 520–526.

Hennessey, J. A., Wei, E. Q., & Pitt, G. S. (2013). Fibroblast growth factor homologous factors

modulate cardiac calcium channels. Circulation Research, 113(4), 381–8.

290

Henry, T. D., Grines, C. L., Watkins, M. W., Dib, N., Barbeau, G., Moreadith, R., … Engler, R.

L. (2007). Effects of Ad5FGF-4 in patients with angina: an analysis of pooled data from the

AGENT-3 and AGENT-4 trials. Journal of the American College of Cardiology, 50(11),

1038–46. doi:10.1016/j.jacc.2007.06.010

Hensel, N., Ratzka, A., Brinkmann, H., Klimaschewski, L., Grothe, C., & Claus, P. (2012).

Analysis of the fibroblast growth factor system reveals alterations in a mouse model of spinal

muscular atrophy. PloS One, 7(2), e31202. doi:10.1371/journal.pone.0031202

Herbert, S. P., & Stainier, D. Y. R. (2011). Molecular control of endothelial cell behaviour during

blood vessel morphogenesis. Nature Reviews. Molecular Cell Biology, 12(9), 551–64.

Hickey, M. M., & Simon, M. C. (2006). Regulation of angiogenesis by hypoxia and hypoxia-

inducible factors. Current Topics in Developmental Biology, 76, 217–257.

Hoefer, I. E., Grundmann, S., van Royen, N., Voskuil, M., Schirmer, S. H., Ulusans, S., … Piek,

J. J. (2005). Leukocyte subpopulations and arteriogenesis: specific role of monocytes,

lymphocytes and granulocytes. Atherosclerosis, 181(2), 285–93.

Hoefer, I. E., van Royen, N., Buschmann, I. R., Piek, J. J., & Schaper, W. (2001). Time course of

arteriogenesis following femoral artery occlusion in the rabbit. Cardiovascular Research,

49(3), 609–617.

Hoefer, I. E., van Royen, N., & Jost, M. M. (2006). Experimental models of arteriogenesis:

differences and implications. Lab Animal, 35(2), 36–44. doi:10.1038/laban0206-36

Homer-Bouthiette, C., Doetschman, T., Xiao, L., & Hurley, M. M. (2014). Knockout of Nuclear

High Molecular Weight FGF2 Isoforms in Mice Modulates Bone and Phosphate

291

Homeostasis. J Biol Chem, 289(52), 36303–36314. doi:10.1074/jbc.M114.619569

Hortala, M., Estival, A., Pradayrol, L., Susini, C., & Clemente, F. (2005). Identification of c-Jun

as a critical mediator for the intracrine 24 kDa FGF-2 isoform-induced cell proliferation.

International Journal of Cancer. Journal International Du Cancer, 114(6), 863–9.

House, S. L., Bolte, C., Zhou, M., Doetschman, T., Klevitsky, R., Newman, G., & Schultz, J. E. J.

(2003). Cardiac-specific overexpression of fibroblast growth factor-2 protects against

myocardial dysfunction and infarction in a murine model of low-flow ischemia. Circulation,

108(25), 3140–3148.

House, S. L., Branch, K., Newman, G., Doetschman, T., & Schultz, J. E. J. (2005).

Cardioprotection induced by cardiac-specific overexpression of fibroblast growth factor-2 is

mediated by the MAPK cascade. American Journal of Physiology. Heart and Circulatory

Physiology, 289(5), H2167–75. doi:10.1152/ajpheart.00392.2005

House, S. L., House, B. E., Glascock, B., Kimball, T., Nusayr, E., Schultz, J. E. J., & Doetschman,

T. (2010, May 10). Fibroblast Growth Factor 2 Mediates Isoproterenol-induced Cardiac

Hypertrophy through Activation of the Extracellular Regulated Kinase. Molecular and

Cellular Pharmacology. doi:10.4255/mcpharmacol.10.20

House, S. L., Melhorn, S. J., Newman, G., Doetschman, T., & Schultz, J. E. J. (2007). The protein

kinase C pathway mediates cardioprotection induced by cardiac-specific overexpression of

fibroblast growth factor-2. American Journal of Physiology. Heart and Circulatory

Physiology, 293(1), H354–65.

House, S. L., Wang, J., Castro, A. M., Weinheimer, C., Kovacs, A., & Ornitz, D. M. (2015).

Fibroblast growth factor 2 is an essential cardioprotective factor in a closed-chest model of

292

cardiac ischemia-reperfusion injury. Physiological Reports, 3(1).

Huang, P.-H., Chen, Y.-H., Wang, C.-H., Chen, J.-S., Tsai, H.-Y., Lin, F.-Y., … Lin, S.-J. (2009).

Matrix metalloproteinase-9 is essential for ischemia-induced neovascularization by

modulating bone marrow-derived endothelial progenitor cells. Arteriosclerosis, Thrombosis,

and Vascular Biology, 29(8), 1179–84.

Huang, Z., Chen, K., Huang, P. L., Finklestein, S. P., & Moskowitz, M. A. (1997). bFGF

ameliorates focal ischemic injury by blood flow-independent mechanisms in eNOS mutant

mice. The American Journal of Physiology, 272(3 Pt 2), H1401–5.

Hudlicka, O., & Brown, M. D. (2009). Adaptation of skeletal muscle microvasculature to increased

or decreased blood flow: role of shear stress, nitric oxide and vascular endothelial growth

factor. J Vasc Res, 46(5), 504–512.

Hudlicka, O., Brown, M. D., Egginton, S., & Dawson, J. M. (1994). Effect of long-term electrical

stimulation on vascular supply and fatigue in chronically ischemic muscles. J Appl Physiol,

77(3), 1317–1324.

Hughes, G. C., Post, M. J., Simons, M., & Annex, B. H. (2003). Translational physiology: porcine

models of human coronary artery disease: implications for preclinical trials of therapeutic

angiogenesis. Journal of Applied Physiology (Bethesda, Md. : 1985), 94(5), 1689–701.

Hughes, S. E. (1997). Differential Expression of the Fibroblast Growth Factor Receptor (FGFR)

Multigene Family in Normal Human Adult Tissues. Journal of Histochemistry &

Cytochemistry, 45(7), 1005–1019. doi:10.1177/002215549704500710

Imai, T., Hieshima, K., Haskell, C., Baba, M., Nagira, M., Nishimura, M., … Yoshie, O. (1997).

293

Identification and molecular characterization of fractalkine receptor CX3CR1, which

mediates both leukocyte migration and adhesion. Cell, 91(4), 521–30.

Itoh, N., & Ohta, H. (2013). Pathophysiological roles of FGF signaling in the heart. Frontiers in

Physiology, 4. doi:10.3389/fphys.2013.00247

Itoh, N., & Ornitz, D. M. (2011). Fibroblast growth factors: From molecular evolution to roles in

development, metabolism and disease. Journal of Biochemistry, 149(2), 121–130.

Javerzat, S., Auguste, P., & Bikfalvi, A. (2002). The role of fibroblast growth factors in vascular

development. Trends in Molecular Medicine, 8(10), 483–489.

Jetten, N., Verbruggen, S., Gijbels, M. J., Post, M. J., Winther, M. P. J. De, Donners, M. M. P. C.,

& De Winther, M. P. J. (2014). Anti-inflammatory M2, but not pro-inflammatory M1

macrophages promote angiogenesis in vivo. Angiogenesis, 17(1), 109–118.

Jha, P., Ramasundarahettige, C., Landsman, V., Rostron, B., Thun, M., Anderson, R. N., … Peto,

R. (2013). 21st-Century Hazards of Smoking and Benefits of Cessation in the United States.

New England Journal of Medicine, 368(4), 341–350.

Jiang, Z.-S. S., Jeyaraman, M., Wen, G. B., Fandrich, R. R., Dixon, I. M., Cattini, P. A., &

Kardami, E. (2007). High- but not low-molecular weight FGF-2 causes cardiac hypertrophy

in vivo; possible involvement of cardiotrophin-1. J Mol Cell Cardiol, 42(1), 222–233.

Jiang, Z.-S. S., Padua, R. R., Ju, H., Doble, B. W., Jin, Y., Hao, J., … Kardami, E. (2002). Acute

protection of ischemic heart by FGF-2: involvement of FGF-2 receptors and protein kinase

C. American Journal of Physiology - Heart and Circulatory Physiology, 282(3), H1071–80.

Jiang, Z.-S. S., Srisakuldee, W., Soulet, F., Bouche, G., & Kardami, E. (2004). Non-angiogenic

294

FGF-2 protects the ischemic heart from injury, in the presence or absence of reperfusion.

Cardiovasc Res, 62(1), 154–166.

Jiang, Z.-S. S., Wen, G.-B., Tang, Z.-H., Srisakuldee, W., Fandrich, R. R., & Kardami, E. (2009).

High molecular weight FGF-2 promotes postconditioning-like cardioprotection linked to

activation of protein kinase C isoforms, as well as Akt and p70 S6 kinases. [corrected].

Canadian Journal of Physiology and Pharmacology, 87(10), 798–804. doi:10.1139/Y09-049

Jih, Y. J., Lien, W. H., Tsai, W. C., Yang, G. W., Li, C., & Wu, L. W. (2001). Distinct regulation

of genes by bFGF and VEGF-A in endothelial cells. Angiogenesis, 4(4), 313–21.

Jimenez, S. K., Sheikh, F., Jin, Y., Detillieux, K. A., Dhaliwal, J., Kardami, E., & Cattini, P. A.

(2004). Transcriptional regulation of FGF-2 gene expression in cardiac myocytes.

Cardiovascular Research, 62(3), 548–57.

Jun, J.-I., & Lau, L. F. (2011). Taking aim at the extracellular matrix: CCN proteins as emerging

therapeutic targets. Nature Reviews. Drug Discovery, 10(12), 945–63. doi:10.1038/nrd3599

Jungnickel, J., Claus, P., Gransalke, K., Timmer, M., & Grothe, C. (2004). Targeted disruption of

the FGF-2 gene affects the response to peripheral nerve injury. Molecular and Cellular

Neurosciences, 25(3), 444–52.

Jungnickel, J., Haastert, K., Grzybek, M., Thau, N., Lipokatic-Takacs, E., Ratzka, A., … Grothe,

C. (2010). Mice lacking basic fibroblast growth factor showed faster sensory recovery.

Experimental Neurology, 223(1), 166–72.

Kajita, Y., Takayasu, M., Yoshida, J., Dietrich, H. H., & Dacey, R. G. (2001). Vasodilatory effect

of basic fibroblast growth factor in isolated rat cerebral arterioles: mechanisms involving

295

nitric oxide and membrane hyperpolarization. Neurologia Medico-Chirurgica, 41(4), 177–

85; discussion 185–6.

Kanda, S., Miyata, Y., & Kanetake, H. (2004). Fibroblast growth factor-2-mediated capillary

morphogenesis of endothelial cells requires signals via Flt-1/vascular endothelial growth

factor receptor-1: possible involvement of c-Akt. The Journal of Biological Chemistry,

279(6), 4007–16. doi:10.1074/jbc.M307569200

Kang, J.-S., & Krauss, R. S. (2010). Muscle stem cells in developmental and regenerative

myogenesis. Current Opinion in Clinical Nutrition and Metabolic Care, 13(3), 243–8.

Kano, M. R., Morishita, Y., Iwata, C., Iwasaka, S., Watabe, T., Ouchi, Y., … Miyazawa, K. (2005).

VEGF-A and FGF-2 synergistically promote neoangiogenesis through enhancement of

endogenous PDGF-B-PDGFRbeta signaling. Journal of Cell Science, 118(Pt 16), 3759–68.

Karalaki, M., Fili, S., Philippou, A., & Koutsilieris, M. (2009). Muscle regeneration: cellular and

molecular events. In Vivo (Athens, Greece), 23(5), 779–796.

Kardami, E., Detillieux, K. A., Ma, X., Jiang, Z.-S. S., Santiago, J.-J., Jimenez, S. K., & Cattini,

P. A. (2007). Fibroblast growth factor-2 and cardioprotection. Heart Failure Reviews, 12(3-

4), 267–77.

Kardami, E., Jiang, Z.-S. S., Jimenez, S. K., Hirst, C. J., Sheikh, F., Zahradka, P., & Cattini, P. A.

(2004). Fibroblast growth factor 2 isoforms and cardiac hypertrophy. Cardiovasc Res, 63(3),

458–466.

Kästner, S., Elias, M. C., Rivera, A. J., & Yablonka-Reuveni, Z. (2000). Gene expression patterns

of the fibroblast growth factors and their receptors during myogenesis of rat satellite cells.

296

The Journal of Histochemistry and Cytochemistry: Official Journal of the Histochemistry

Society, 48(8), 1079–1096.

Katoh, M. (2013). Therapeutics targeting angiogenesis: genetics and epigenetics, extracellular

miRNAs and signaling networks (Review). International Journal of Molecular Medicine,

32(4), 763–767. doi:10.3892/ijmm.2013.1444

Keeley, E. C., Mehrad, B., & Strieter, R. M. (2008). Chemokines as mediators of

neovascularization. Arteriosclerosis, Thrombosis, and Vascular Biology, 28(11), 1928–36.

doi:10.1161/ATVBAHA.108.162925

Kelsall, C. J., Brown, M. D., & Hudlicka, O. (2001). Alterations in reactivity of small arterioles in

rat skeletal muscle as a result of chronic ischaemia. Journal of Vascular Research, 38(3),

212–8. doi:51049

Kelsall, C. J., Brown, M. D., Kent, J., Kloehn, M., & Hudlicka, O. (2004). Arteriolar endothelial

dysfunction is restored in ischaemic muscles by chronic electrical stimulation. Journal of

Vascular Research, 41(3), 241–51.

Keum, E., Kim, Y., Kim, J., Kwon, S., Lim, Y., Han, I., & Oh, E.-S. (2004). Syndecan-4 regulates

localization, activity and stability of protein kinase C-alpha. The Biochemical Journal, 378(Pt

3), 1007–14. doi:10.1042/BJ20031734

Khurana, R., & Simons, M. (2003). Insights from angiogenesis trials using fibroblast growth factor

for advanced arteriosclerotic disease. Trends in Cardiovascular Medicine, 13(3), 116–22.

Kiyota, T., Ingraham, K. L., Jacobsen, M. T., Xiong, H., & Ikezu, T. (2011). FGF2 gene transfer

restores hippocampal functions in mouse models of Alzheimer’s disease and has therapeutic

297

implications for neurocognitive disorders. Proceedings of the National Academy of Sciences

of the United States of America, 108(49), E1339–48. doi:10.1073/pnas.1102349108

Kline, T. L., & Ritman, E. L. (2012). Studying Microcirculation with Micro-CT. In rtin J. Leahy

(Ed.), Microcirculation Imaging (pp. 313–347). Wiley-VCH Verlag GmbH & Co. KGaA.

Koestler, T. P., Rieman, D., Muirhead, K., Greig, R. G., & Poste, G. (1984). Identification and

characterization of a monoclonal antibody to an antigen expressed on activated macrophages.

Proceedings of the National Academy of Sciences of the United States of America, 81(14),

4505–9.

Korada, S., Zheng, W., Basilico, C., Schwartz, M. L., & Vaccarino, F. M. (2002). Fibroblast

Growth Factor 2 Is Necessary for the Growth of Glutamate Projection Neurons in the Anterior

Neocortex. J. Neurosci., 22(3), 863–875.

Krejci, P., Pejchalova, K., Rosenbloom, B. E., Rosenfelt, F. P., Tran, E. L., Laurell, H., & Wilcox,

W. R. (2007). The antiapoptotic protein Api5 and its partner, high molecular weight FGF2,

are up-regulated in B cell chronic lymphoid leukemia. Journal of Leukocyte Biology, 82(6),

1363–1364. doi:10.1189/jlb.0607425

Krejci, P., Prochazkova, J., Bryja, V., Kozubik, A., & Wilcox, W. R. (2009). Molecular pathology

of the fibroblast growth factor family. Human Mutation, 30(9), 1245–1255.

Krock, B. L., Skuli, N., & Simon, M. C. (2011). Hypoxia-Induced Angiogenesis. Genes & Cancer,

2(12), 1117–1133.

Kubota, S., & Takigawa, M. (2007). CCN family proteins and angiogenesis: from embryo to

adulthood. Angiogenesis, 10(1), 1–11. doi:10.1007/s10456-006-9058-5

298

la Sala, A., Pontecorvo, L., Agresta, A., Rosano, G., & Stabile, E. (2012). Regulation of collateral

blood vessel development by the innate and adaptive immune system. Trends in Molecular

Medicine, 18(8), 494–501. doi:10.1016/j.molmed.2012.06.007

Laestander, C., & Engström, W. (2014). Role of fibroblast growth factors in elicitation of cell

responses. Cell Proliferation, 47(1), 3–11. doi:10.1111/cpr.12084

Laezza, F., Gerber, B. R., Lou, J. Y., Kozel, M. A., Hartman, H., Craig, A. M., … Nerbonne, J.

M. (2007). The FGF14(F145S) mutation disrupts the interaction of FGF14 with voltage-gated

Na+ channels and impairs neuronal excitability. J Neurosci, 27(44), 12033–12044.

Laham, R. J., Chronos, N. A., Pike, M., Leimbach, M. E., Udelson, J. E., Pearlman, J. D., …

Simons, M. (2000). Intracoronary basic fibroblast growth factor (FGF-2) in patients with

severe ischemic heart disease: results of a Phase I open-label dose escalation study. Journal

of the American College of Cardiology, 36(7), 2132–2139.

Laham, R. J., Rezaee, M., Post, M., Novicki, D., Sellke, F. W., Pearlman, J. D., … Hung, D.

(2000). Intrapericardial delivery of fibroblast growth factor-2 induces neovascularization in

a porcine model of chronic myocardial ischemia. The Journal of Pharmacology and

Experimental Therapeutics, 292(2), 795–802.

Laham, R. J., Sellke, F. W., Edelman, E. R., Pearlman, J. D., Ware, J. A., Brown, D. L., … Simons,

M. (1999). Local perivascular delivery of basic fibroblast growth factor in patients

undergoing coronary bypass surgery: results of a phase I randomized, double-blind, placebo-

controlled trial. Circulation, 100(18), 1865–71.

Laitinen, L. (1987). Griffonia simplicifolia lectins bind specifically to endothelial cells and some

epithelial cells in mouse tissues. The Histochemical Journal, 19(4), 225–34.

299

Landázuri, N., Joseph, G., Guldberg, R. E., & Taylor, W. R. (2012). Growth and regression of

vasculature in healthy and diabetic mice after hindlimb ischemia. American Journal of

Physiology. Regulatory, Integrative and Comparative Physiology, 303(1), R48–56.

Lau, L. F. (2011). CCN1/CYR61: the very model of a modern matricellular protein. Cellular and

Molecular Life Sciences : CMLS, 68(19), 3149–63. doi:10.1007/s00018-011-0778-3

Lavu, M., Gundewar, S., & Lefer, D. J. (2011). Gene therapy for ischemic heart disease. Journal

of Molecular and Cellular Cardiology, 50(5), 742–50. doi:10.1016/j.yjmcc.2010.06.007

Lazarous, D. F., Scheinowitz, M., Shou, M., Hodge, E., Rajanayagam, M. A. S., Hunsberger, S.,

… Unger, E. F. (1995). Effects of Chronic Systemic Administration of Basic Fibroblast

Growth Factor on Collateral Development in the Canine Heart. Circulation, 91(1), 145–153.

Lazarous, D. F., Unger, E. F., Epstein, S. E., Stine, A., Arevalo, J. L., Chew, E. Y., & Quyyumi,

A. A. (2000). Basic fibroblast growth factor in patients with intermittent claudication: results

of a phase I trial. Journal of the American College of Cardiology, 36(4), 1239–1244.

Le Grand, F., & Rudnicki, M. A. (2007). Skeletal muscle satellite cells and adult myogenesis.

Current Opinion in Cell Biology, 19(6), 628–33. doi:10.1016/j.ceb.2007.09.012

Lea, R., Papalopulu, N., Amaya, E., & Dorey, K. (2009). Temporal and spatial expression of FGF

ligands and receptors during Xenopus development. Developmental Dynamics : An Official

Publication of the American Association of Anatomists, 238(6), 1467–79.

Leahy, M. J., de Mul, F. F., Nilsson, G. E., & Maniewski, R. (1999). Principles and practice of the

laser-Doppler perfusion technique. Technol Health Care, 7(2-3), 143–162.

Leali, D., Dell’Era, P., Stabile, H., Sennino, B., Chambers, A. F., Naldini, A., … Presta, M. (2003).

300

Osteopontin (Eta-1) and Fibroblast Growth Factor-2 Cross-Talk in Angiogenesis. The

Journal of Immunology, 171(2), 1085–1093. doi:10.4049/jimmunol.171.2.1085

LeBlanc, A. J., Krishnan, L., Sullivan, C. J., Williams, S. K., & Hoying, J. B. (2012).

Microvascular Repair: Post-Angiogenesis Vascular Dynamics. Microcirculation, 19(8), 676–

695.

Lederman, R. J., Mendelsohn, F. O., Anderson, R. D., Saucedo, J. F., Tenaglia, A. N., Hermiller,

J. B., … Annex, B. H. (2002). Therapeutic angiogenesis with recombinant fibroblast growth

factor-2 for intermittent claudication (the TRAFFIC study): a randomised trial. Lancet,

359(9323), 2053–2058.

Lee, H. J., Cho, C.-H., Hwang, S.-J., Choi, H.-H., Kim, K.-T., Ahn, S. Y., … Koh, G. Y. (2004).

Biological characterization of angiopoietin-3 and angiopoietin-4. FASEB Journal : Official

Publication of the Federation of American Societies for Experimental Biology, 18(11), 1200–

8.

Lee, J. E., Mirza, S. P., Didier, D. N., Scalf, M., Olivier, M., Greene, A. S., & Smith, L. M. (2008).

Identification of cell surface markers to differentiate rat endothelial and fibroblast cells using

lectin arrays and LC-ESI-MS/MS. Analytical Chemistry, 80(21), 8269–75.

Lee, S.-J., Namkoong, S., Kim, Y.-M., Kim, C.-K., Lee, H., Ha, K.-S., … Kwon, Y.-G. (2006).

Fractalkine stimulates angiogenesis by activating the Raf-1/MEK/ERK- and

PI3K/Akt/eNOS-dependent signal pathways. AJP: Heart and Circulatory Physiology,

291(6), H2836–H2846. doi:10.1152/ajpheart.00113.2006

Leonard, S., Luthman, D., Logel, J., Luthman, J., Antle, C., Freedman, R., & Hoffer, B. (1993).

Acidic and basic fibroblast growth factor mRNAs are increased in striatum following MPTP-

301

induced dopamine neurofiber lesion: assay by quantitative PCR. Molecular Brain Research,

18(4), 275–284. doi:10.1016/0169-328X(93)90090-C

Lerman, I., Harrison, B. C., Freeman, K., Hewett, T. E., Allen, D. L., Robbins, J., & Leinwand, L.

A. (2002). Genetic variability in forced and voluntary endurance exercise performance in

seven inbred mouse strains. Journal of Applied Physiology, 92(6), 2245–55.

Levin, E. G., Sikora, L., Ding, L., Rao, S. ., & Sriramarao, P. (2004). Suppression of tumor growth

and angiogenesis in vivo by a truncated form of 24-kd fibroblast growth factor (FGF)-2. The

American Journal of Pathology, 164(4), 1183–90. doi:10.1016/S0002-9440(10)63206-3

Li, A., Guo, H., Luo, X., Sheng, J., Yang, S., Yin, Y., & Zhou, J. (2006). Apomorphine-induced

activation of dopamine receptors modulates FGF-2 expression in astrocytic cultures and

promotes survival of dopaminergic neurons. FASEB Journal : Official Publication of the

Federation of American Societies for Experimental Biology, 20(8), 1263–5.

Li, J., Wei, Y., Liu, K., Yuan, C., Tang, Y., Quan, Q., … Yang, L. (2010). Synergistic effects of

FGF-2 and PDGF-BB on angiogenesis and muscle regeneration in rabbit hindlimb ischemia

model. Microvascular Research, 80(1), 10–7. doi:10.1016/j.mvr.2009.12.002

Li, Q. F., Wang, X. R., Yang, Y. W., & Lin, H. (2006). Hypoxia upregulates hypoxia inducible

factor (HIF)-3alpha expression in lung epithelial cells: characterization and comparison with

HIF-1alpha. Cell Research, 16(6), 548–58. doi:10.1038/sj.cr.7310072

Li, W., Shen, W., Gill, R., Corbly, A., Jones, B., Belagaje, R., … Singh, J. (2006). High-resolution

quantitative computed tomography demonstrating selective enhancement of medium-size

collaterals by placental growth factor-1 in the mouse ischemic hindlimb. Circulation,

113(20), 2445–2453.

302

Liao, S. (2008). The Role of Fibroblast Grown Factor-2 Isoforms in Ischemia-Reperfusion Injury

and Cardioprotection. (L. E. T. Ohio & C. Dissertations, Eds.). University of Cincinnati,

Cincinnati, Ohio :

Liao, S., Bodmer, J., Pietras, D., Azhar, M., Doetschman, T., & Schultz, J. E. J. (2009). Biological

functions of the low and high molecular weight protein isoforms of fibroblast growth factor-

2 in cardiovascular development and disease. Developmental Dynamics, 238(2), 249–264.

Liao, S., Bodmer, J. R., Azhar, M., Newman, G., Coffin, J. D., Doetschman, T., & Schultz, J. E.

J. (2010). The influence of FGF2 high molecular weight (HMW) isoforms in the development

of cardiac ischemia-reperfusion injury. Journal of Molecular and Cellular Cardiology, 48(6),

1245–1254.

Liao, S., Porter, D., Scott, A., Newman, G., Doetschman, T., & Schultz, J. E. J. (2007). The

cardioprotective effect of the low molecular weight isoform of fibroblast growth factor-2: the

role of JNK signaling. J Mol Cell Cardiol, 42(1), 106–120.

Libby, P., & Theroux, P. (2005). Pathophysiology of Coronary Artery Disease. Circulation,

111(25), 3481–3488. doi:10.1161/CIRCULATIONAHA.105.537878

Limbourg, A., Korff, T., Napp, L. C., Schaper, W., Drexler, H., & Limbourg, F. P. (2009).

Evaluation of postnatal arteriogenesis and angiogenesis in a mouse model of hind-limb

ischemia. Nature Protocols, 4(12), 1737–46. doi:10.1038/nprot.2009.185

Lin, A. S., Palmer, A. W., Duvall, C. L., Robertson, G. C., Oest, M. E., Rai, B., … Guldberg, R.

E. (2007). Contrast-Enhanced Micro-CT Imaging of Soft Tissues. In L. Qin, H. Genant, J.

Griffith, & K. Leung (Eds.), Advanced Bioimaging Technologies in Assessment of the Quality

of Bone and Scaffold Materials SE - 15 (pp. 239–256). Springer Berlin Heidelberg.

303

doi:10.1007/978-3-540-45456-4_15

Lin, C. G., Leu, S.-J., Chen, N., Tebeau, C. M., Lin, S.-X., Yeung, C.-Y., & Lau, L. F. (2003).

CCN3 (NOV) is a novel angiogenic regulator of the CCN protein family. The Journal of

Biological Chemistry, 278(26), 24200–8.

Lin, T. N., Te, J., Lee, M., Sun, G. Y., & Hsu, C. Y. (1997). Induction of basic fibroblast growth

factor (bFGF) expression following focal cerebral ischemia. Brain Research. Molecular

Brain Research, 49(1-2), 255–65.

Liu, J., Wang, Y., Akamatsu, Y., Lee, C. C., Stetler, R. A., Lawton, M. T., & Yang, G.-Y. (2014).

Vascular remodeling after ischemic stroke: Mechanisms and therapeutic potentials. Progress

in Neurobiology, 115, 138–156.

Liu, L., Doble, B. W., & Kardami, E. (1993). Perinatal phenotype and hypothyroidism are

associated with elevated levels of 21.5- to 22-kDa basic fibroblast growth factor in cardiac

ventricles. Developmental Biology, 157(2), 507–16. doi:10.1006/dbio.1993.1153

Liu, X., & Zhu, X.-Z. (1999). Increased expression and nuclear accumulation of basic fibroblast

growth factor in primary cultured astrocytes following ischemic-like insults. Molecular Brain

Research, 71(2), 171–177.

Livak, K. J., & Schmittgen, T. D. (2001). Analysis of relative gene expression data using real-time

quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods (San Diego, Calif.), 25(4),

402–8. doi:10.1006/meth.2001.1262

Lizotte, F., Paré, M., Denhez, B., Leitges, M., Guay, A., & Geraldes, P. (2013). PKCδ impaired

vessel formation and angiogenic factor expression in diabetic ischemic limbs. Diabetes,

304

62(8), 2948–57.

Lloyd, P. G., Prior, B. M., Yang, H.-T., & Terjung, R. L. (2003). Angiogenic growth factor

expression in rat skeletal muscle in response to exercise training. American Journal of

Physiology. Heart and Circulatory Physiology, 284(5), H1668–78.

Lobb, R. R., Alderman, E. M., & Fett, J. W. (1985). Induction of angiogenesis by bovine brain

derived class 1 heparin-binding growth factor. Biochemistry, 24(19), 4969–73.

Lobb, R. R., & Fett, J. W. (1984). Purification of two distinct growth factors from bovine neural

tissue by heparin affinity chromatography. Biochemistry, 23(26), 6295–6299.

Lotfi, S., Patel, A. S., Mattock, K., Egginton, S., Smith, A., & Modarai, B. (2013). Towards a more

relevant hind limb model of muscle ischaemia. Atherosclerosis, 227(1), 1–8.

Lu, H., Xu, X., Zhang, M., Cao, R., Bråkenhielm, E., Li, C., … Cao, Y. (2007). Combinatorial

protein therapy of angiogenic and arteriogenic factors remarkably improves collaterogenesis

and cardiac function in pigs. Proceedings of the National Academy of Sciences of the United

States of America, 104(29), 12140–5. doi:10.1073/pnas.0704966104

Lu, J., Pompili, V. J., & Das, H. (2013). Neovascularization and hematopoietic stem cells. Cell

Biochemistry and Biophysics, 67(2), 235–45. doi:10.1007/s12013-011-9298-x

Ma, X., Dang, X., Claus, P., Hirst, C., Fandrich, R. R., Jin, Y., … Kardami, E. (2007). Chromatin

compaction and cell death by high molecular weight FGF-2 depend on its nuclear

localization, intracrine ERK activation, and engagement of mitochondria. Journal of Cellular

Physiology, 213(3), 690–698. doi:10.1002/jcp.21139

MacGabhann, F., & Peirce, S. M. (2010). Collateral Capillary Arterialization following Arteriolar

305

Ligation in Murine Skeletal Muscle. Microcirculation, 17(5), 333–347. doi:10.1111/j.1549-

8719.2010.00034.x

Madeddu, P., Emanueli, C., Spillmann, F., Meloni, M., Bouby, N., Richer, C., … Levy, B. I.

(2006). Murine models of myocardial and limb ischemia: diagnostic end-points and relevance

to clinical problems. Vascul Pharmacol, 45(5), 281–301.

Malik, A. R., Liszewska, E., & Jaworski, J. (2015). Matricellular proteins of the

Cyr61/CTGF/NOV (CCN) family and the nervous system. Frontiers in Cellular

Neuroscience, 9, 237.

Manning, J. R., Perkins, S. O., Sinclair, E. A., Gao, X., Zhang, Y., Newman, G., … Schultz, J. E.

J. (2013). Low molecular weight fibroblast growth factor-2 signals via protein kinase C and

myofibrillar proteins to protect against postischemic cardiac dysfunction. American Journal

of Physiology. Heart and Circulatory Physiology, 304(10), H1382–96.

Maretzky, T., Evers, A., Zhou, W., Swendeman, S. L., Wong, P.-M., Rafii, S., … Blobel, C. P.

(2011). Migration of growth factor-stimulated epithelial and endothelial cells depends on

EGFR transactivation by ADAM17. Nature Communications, 2, 229.

doi:10.1038/ncomms1232

Marie, P. J., Miraoui, H., & Sévère, N. (2012). FGF/FGFR signaling in bone formation: progress

and perspectives. Growth Factors (Chur, Switzerland), 30(2), 117–123.

doi:10.3109/08977194.2012.656761

Martinez, C. O., McHale, M. J., Wells, J. T., Ochoa, O., Michalek, J. E., McManus, L. M., &

Shireman, P. K. (2010). Regulation of skeletal muscle regeneration by CCR2-activating

chemokines is directly related to macrophage recruitment. American Journal of Physiology.

306

Regulatory, Integrative and Comparative Physiology, 299(3), R832–842.

Marui, A., Kanematsu, A., Yamahara, K., Doi, K., Kushibiki, T., Yamamoto, M., … Komeda, M.

(2005). Simultaneous application of basic fibroblast growth factor and hepatocyte growth

factor to enhance the blood vessels formation. Journal of Vascular Surgery, 41(1), 82–90.

doi:10.1016/j.jvs.2004.10.029

Masaki, I., Yonemitsu, Y., Yamashita, A., Sata, S., Tanii, M., Komori, K., … Sueishi, K. (2002).

Angiogenic Gene Therapy for Experimental Critical Limb Ischemia: Acceleration of Limb

Loss by Overexpression of Vascular Endothelial Growth Factor 165 but not of Fibroblast

Growth Factor-2. Circulation Research, 90(9), 966–973.

Mascarenhas, J. V, Albayati, M. A., Shearman, C. P., & Jude, E. B. (2014). Peripheral Arterial

Disease. Endocrinology and Metabolism Clinics of North America, 43(1), 149–166.

Masumura, M., Murayama, N., Inoue, T., & Ohno, T. (1996). Selective induction of fibroblast

growth factor receptor-1 mRNA after transient focal ischemia in the cerebral cortex of rats.

Neuroscience Letters, 213(2), 119–122. doi:10.1016/0304-3940(96)12841-X

Matsakas, A., Yadav, V., Lorca, S., Evans, R. M., & Narkar, V. a. (2012). Revascularization of

Ischemic Skeletal Muscle by Estrogen-Related Receptor-γ. Circulation Research, 110(8),

1087–1096. doi:10.1161/CIRCRESAHA.112.266478

Matsui, Y., Morimoto, J., & Uede, T. (2010). Role of matricellular proteins in cardiac tissue

remodeling after myocardial infarction. World Journal of Biological Chemistry, 1(5), 69–80.

Matsumoto, T., Turesson, I., Book, M., Gerwins, P., & Claesson-Welsh, L. (2002). p38 MAP

kinase negatively regulates endothelial cell survival, proliferation, and differentiation in FGF-

307

2-stimulated angiogenesis. J Cell Biol, 156(1), 149–160. doi:10.1083/jcb.200103096

Maxwell, M. P., Hearse, D. J., & Yellon, D. M. (1987). Species variation in the coronary collateral

circulation during regional myocardial ischaemia: a critical determinant of the rate of

evolution and extent of myocardial infarction. Cardiovascular Research, 21(10), 737–746.

McClung, J. M., McCord, T. J., Keum, S., Johnson, S., Annex, B. H., Marchuk, D. A., & Kontos,

C. D. (2012). Skeletal muscle-specific genetic determinants contribute to the differential

strain-dependent effects of hindlimb ischemia in mice. The American Journal of Pathology,

180(5), 2156–69.

McClung, J. M., Reinardy, J. L., Mueller, S. B., McCord, T. J., Kontos, C. D., Brown, D. A., …

Green, T. D. (2015). Muscle cell derived angiopoietin-1 contributes to both myogenesis and

angiogenesis in the ischemic environment. Frontiers in Physiology, 6, 161.

McIntosh, J., Dennison, G., Holly, J. M. P., Jarrett, C., Frankow, A., Foulstone, E. J., … Perks, C.

M. (2010). IGFBP-3 Can Either Inhibit or Enhance EGF-mediated Growth of Breast

Epithelial Cells Dependent upon the Presence of Fibronectin. Journal of Biological

Chemistry, 285(50), 38788–38800.

Medcalf, R. L. (2007). Fibrinolysis, inflammation, and regulation of the plasminogen activating

system. Journal of Thrombosis and Haemostasis, 5, 132–142.

Meisinger, C., & Grothe, C. (1997a). Differential expression of FGF-2 isoforms in the rat adrenal

medulla during postnatal development in vivo. Brain Research, 757(2), 291–294.

Meisinger, C., & Grothe, C. (1997b). Differential regulation of fibroblast growth factor (FGF)-2

and FGF receptor 1 mRNAs and FGF-2 isoforms in spinal ganglia and sciatic nerve after

308

peripheral nerve lesion. Journal of Neurochemistry, 68(3), 1150–8.

Meisner, J. K., Annex, B. H., & Price, R. J. (2014). Despite normal arteriogenic and angiogenic

responses, hind limb perfusion recovery and necrotic and fibroadipose tissue clearance are

impaired in matrix metalloproteinase 9-deficient mice. Journal of Vascular Surgery.

Meisner, J. K., & Price, R. J. (2010). Spatial and temporal coordination of bone marrow-derived

cell activity during arteriogenesis: regulation of the endogenous response and therapeutic

implications. Microcirculation (New York, N.Y.: 1994), 17(8), 583–599.

Meisner, J. K., Song, J., Annex, B. H., & Price, R. J. (2013). Myoglobin Overexpression Inhibits

Reperfusion in the Ischemic Mouse Hindlimb through Impaired Angiogenesis but Not

Arteriogenesis. The American Journal of Pathology, 183(6), 1710–1718.

Mellin, T. N., Cashen, D. E., Ronan, J. J., Murphy, B. S., DiSalvo, J., & Thomas, K. A. (1995).

Acidic Fibroblast Growth Factor Accelerates Dermal Wound Healing in Diabetic Mice.

Journal of Investigative Dermatology, 104(5), 850–855. doi:10.1111/1523-1747.ep12607026

Mignatti, P., Morimoto, T., & Rifkin, D. B. (1992). Basic fibroblast growth factor, a protein devoid

of secretory signal sequence, is released by cells via a pathway independent of the

endoplasmic reticulum-Golgi complex. Journal of Cellular Physiology, 151(1), 81–93.

Milkiewicz, M., Hudlicka, O., Shiner, R., Egginton, S., & Brown, M. D. (2006). Vascular

endothelial growth factor mRNA and protein do not change in parallel during non-

inflammatory skeletal muscle ischaemia in rat. The Journal of Physiology, 577(Pt 2), 671–8.

Milkiewicz, M., Hudlicka, O., Verhaeg, J., Egginton, S., & Brown, M. D. (2003). Differential

expression of Flk-1 and Flt-1 in rat skeletal muscle in response to chronic ischaemia:

309

favourable effect of muscle activity. Clinical Science (London, England : 1979), 105(4), 473–

82. doi:10.1042/CS20030035

Miller, D. L., Ortega, S., Bashayan, O., Basch, R., & Basilico, C. (2000). Compensation by

fibroblast growth factor 1 (FGF1) does not account for the mild phenotypic defects observed

in FGF2 null mice. Mol Cell Biol, 20(6), 2260–2268. doi:10.1128/MCB.20.6.2260-

2268.2000

Miyagawa, K., Sakamoto, H., Yoshida, T., Yamashita, Y., Mitsui, Y., Furusawa, M., … Terada,

M. (1988). hst-1 transforming protein: expression in silkworm cells and characterization as a

novel heparin-binding growth factor. Oncogene, 3(4), 383–9.

Miyashiro, M., Ogata, N., Takahashi, K., Matsushima, M., Yamamoto, C., Yamada, H., & Uyama,

M. (1998). Expression of basic fibroblast growth factor and its receptor mRNA in retinal

tissue following ischemic injury in the rat. Graefe’s Archive for Clinical and Experimental

Ophthalmology = Albrecht von Graefes Archiv Für Klinische Und Experimentelle

Ophthalmologie, 236(4), 295–300.

Mohammadi, M., Dikic, I., Sorokin, A., Burgess, W. H., Jaye, M., & Schlessinger, J. (1996).

Identification of six novel autophosphorylation sites on fibroblast growth factor receptor 1

and elucidation of their importance in receptor activation and signal transduction .

Identification of Six Novel Autophosphorylation Sites on Fibroblast Growth Fa. Mol. Cell.

Biol., 16(3), 977–989.

Mohammadi, M., Olsen, S. K., & Ibrahimi, O. A. (2005). Structural basis for fibroblast growth

factor receptor activation. Cytokine & Growth Factor Reviews, 16(2), 107–37.

Molin, D., & Post, M. J. (2007). Therapeutic angiogenesis in the heart: protect and serve. Current

310

Opinion in Pharmacology, 7(2), 158–163.

Montarras, D., L’honoré, A., & Buckingham, M. (2013). Lying low but ready for action: the

quiescent muscle satellite cell. The FEBS Journal, 280(17), 4036–50. doi:10.1111/febs.12372

Montero, A., Okada, Y., Tomita, M., Ito, M., Tsurukami, H., Nakamura, T., … Hurley, M. M.

(2000). Disruption of the fibroblast growth factor-2 gene results in decreased bone mass and

bone formation. The Journal of Clinical Investigation, 105(8), 1085–93.

Montesano, R., Vassalli, J. D., Baird, A., Guillemin, R., & Orci, L. (1986). Basic fibroblast growth

factor induces angiogenesis in vitro. Proceedings of the National Academy of Sciences of the

United States of America, 83(19), 7297–301.

Morisada, T., Kubota, Y., Urano, T., Suda, T., & Oike, Y. (2006). Angiopoietins and angiopoietin-

like proteins in angiogenesis. Endothelium : Journal of Endothelial Cell Research, 13(2), 71–

9.

Moriya, J., Wu, X., Zavala-Solorio, J., Ross, J., Liang, X. H., & Ferrara, N. (2013). PDGF-C

promotes revascularization in ischemic limbs of diabetic mice. Journal of Vascular Surgery.

doi:10.1016/j.jvs.2013.04.053

Morrison, A. R., Yarovinsky, T. O., Young, B. D., Moraes, F., Ross, T. D., Ceneri, N., … Bender,

J. R. (2014). Chemokine-coupled 2 -induced macrophage Rac2-Myosin IIA

interaction regulates VEGF-A mRNA stability and arteriogenesis. Journal of Experimental

Medicine, 211(10), 1957–1968. doi:10.1084/jem.20132130

Moscatelli, D., Joseph-Silverstein, J., Manejias, R., & Rifkin, D. B. (1987). Mr 25,000 heparin-

binding protein from guinea pig brain is a high molecular weight form of basic fibroblast

311

growth factor. Proceedings of the National Academy of Sciences, 84(16), 5778–5782.

Moscatelli, D., Presta, M., & Rifkin, D. B. (1986). Purification of a factor from human placenta

that stimulates capillary endothelial cell protease production, DNA synthesis, and migration.

Proceedings of the National Academy of Sciences, 83(7), 2091–2095.

Mozaffarian, D., Benjamin, E. J., Go, A. S., Arnett, D. K., Blaha, M. J., Cushman, M., … Turner,

M. B. (2015). Heart Disease and Stroke Statistics—2016 Update: A Report From the

American Heart Association. Circulation.

Mueller, M. M. (2008). Inflammation and Angiogenesis: Innate Immune Cells as Modulators of

Tumor Vascularization. In Tumor Angiogenesis Basic Mechanisms and Cancer Therapy (pp.

351–362).

Mühlhauser, J., Pili, R., Merrill, M. J., Maeda, H., Passaniti, A., Crystal, R. G., & Capogrossi, M.

C. (1995). In vivo angiogenesis induced by recombinant adenovirus vectors coding either for

secreted or nonsecreted forms of acidic fibroblast growth factor. Human Gene Therapy, 6(11),

1457–65.

Muir, R. L. (2009). Peripheral arterial disease: Pathophysiology, risk factors, diagnosis, treatment,

and prevention. Journal of Vascular Nursing, 27(2), 26–30. doi:10.1016/j.jvn.2009.03.001

Muller, M. D., Reed, A. B., Leuenberger, U. a, & Sinoway, L. I. (2013). Physiology in medicine:

peripheral arterial disease. Journal of Applied Physiology (Bethesda, Md. : 1985),

115(August 2013), 1219–26.

Müller-Ostermeyer, F., Claus, P., & Grothe, C. (2001). Distinctive effects of rat fibroblast growth

factor-2 isoforms on PC12 and Schwann cells. Growth Factors (Chur, Switzerland), 19(3),

312

175–91.

Murakami, M., Elfenbein, A., & Simons, M. (2008). Non-canonical fibroblast growth factor

signaling in angiogenesis. Cardiovasc Res, 78(2), 223–231.

Murakami, M., & Simons, M. (2008). Fibroblast growth factor regulation of neovascularization.

Curr Opin Hematol, 15(3), 215–220.

Murohara, T., Asahara, T., Silver, M., Bauters, C., Masuda, H., Kalka, C., … Isner, J. M. (1998).

Nitric oxide synthase modulates angiogenesis in response to tissue ischemia. J Clin Invest,

101(11), 2567–2578.

Naganawa, T., Xiao, L., Coffin, J. D., Doetschman, T., Sabbieti, M. G., Agas, D., & Hurley, M.

M. (2008). Reduced expression and function of bone morphogenetic protein-2 in bones of

Fgf2 null mice. Journal of Cellular Biochemistry, 103(6), 1975–88. doi:10.1002/jcb.21589

Nakagawa, T., & Schwartz, J. P. (2006). Gene Expression Profiles of Reactive Astrocytes in

Dopamine-depleted Striatum. Brain Pathology, 14(3), 275–280.

Nakajima, A., Nakajima, F., Shimizu, S., Ogasawara, A., Wanaka, A., Moriya, H., … Yamazaki,

M. (2001). Spatial and temporal gene expression for fibroblast growth factor type I receptor

(FGFR1) during fracture healing in the rat. Bone, 29(5), 458–466. doi:10.1016/S8756-

3282(01)00604-4

Naldini, A., & Carraro, F. (2005). Role of inflammatory mediators in angiogenesis. Current Drug

Targets. Inflammation and Allergy, 4(1), 3–8.

Narmoneva, D. A., Oni, O., Sieminski, A. L., Zhang, S., Gertler, J. P., Kamm, R. D., & Lee, R. T.

(2005). Self-assembling short oligopeptides and the promotion of angiogenesis. Biomaterials,

313

26(23), 4837–46. doi:10.1016/j.biomaterials.2005.01.005

Neuhaus, P., Oustanina, S., Loch, T., Krüger, M., Bober, E., Dono, R., … Braun, T. (2003).

Reduced mobility of fibroblast growth factor (FGF)-deficient myoblasts might contribute to

dystrophic changes in the musculature of FGF2/FGF6/mdx triple-mutant mice. Molecular

and Cellular Biology, 23(17), 6037–6048.

Nguyen, H. X., & Tidball, J. G. (2003). Interactions between neutrophils and macrophages

promote macrophage killing of rat muscle cells in vitro. The Journal of Physiology, 547(Pt

1), 125–32.

Nikol, S., Baumgartner, I., Van Belle, E., Diehm, C., Visoná, A., Capogrossi, M. C., … Meyer, F.

(2008). Therapeutic angiogenesis with intramuscular NV1FGF improves amputation-free

survival in patients with critical limb ischemia. Molecular Therapy : The Journal of the

American Society of Gene Therapy, 16(5), 972–8. doi:10.1038/mt.2008.33

Nindl, W., Kavakebi, P., Claus, P., Grothe, C., Pfaller, K., & Klimaschewski, L. (2004).

Expression of basic fibroblast growth factor isoforms in postmitotic sympathetic neurons:

synthesis, intracellular localization and involvement in karyokinesis. Neuroscience, 124(3),

561–572.

Nissen, L. J., Cao, R., Hedlund, E.-M., Wang, Z., Zhao, X., Wetterskog, D., … Cao, Y. (2007).

Angiogenic factors FGF2 and PDGF-BB synergistically promote murine tumor

neovascularization and metastasis. The Journal of Clinical Investigation, 117(10), 2766–77.

Niu, J., Chang, Z., Peng, B., Xia, Q., Lu, W., Huang, P., … Chiao, P. J. (2007). Keratinocyte

growth factor/fibroblast growth factor-7-regulated cell migration and invasion through

activation of NF-kappaB transcription factors. The Journal of Biological Chemistry, 282(9),

314

6001–11.

Noda, M., Takii, K., Parajuli, B., Kawanokuchi, J., Sonobe, Y., Takeuchi, H., … Suzumura, A.

(2014). FGF-2 released from degenerating neurons exerts microglial-induced neuroprotection

via FGFR3-ERK signaling pathway. Journal of Neuroinflammation, 11(1), 76.

Norgren, L., Hiatt, W. R., Dormandy, J. A., Nehler, M. R., Harris, K. A., & Fowkes, F. G. R.

(2007). Inter-Society Consensus for the Management of Peripheral Arterial Disease (TASC

II). Journal of Vascular Surgery, 45 Suppl S(1), S5–67. doi:10.1016/j.jvs.2006.12.037

Nusayr, E., & Doetschman, T. (2013). Cardiac development and physiology are modulated by

FGF2 in an isoform- and sex-specific manner. Physiological Reports, 1(4).

doi:10.1002/phy2.87

Nusayr, E., Sadideen, D. T., & Doetschman, T. (2013). FGF2 modulates cardiac remodeling in an

isoform- and sex-specific manner. Physiological Reports, 1(4). doi:10.1002/phy2.88

Oh, Y., Gucev, Z., Ng, L., Müller, H. L., & Rosenfeld, R. G. (1995). Antiproliferative actions of

insulin-like growth factor binding protein (IGFBP)-3 in human breast cancer cells. Progress

in Growth Factor Research, 6(2-4), 503–12.

Okada-Ban, M., Thiery, J. P., & Jouanneau, J. (2000). Fibroblast growth factor-2. Int J Biochem

Cell Biol, 32(3), 263–267. doi:S1357-2725(99)00133-8 [pii]

Oklu, R., Albadawi, H., Jones, J. E., Yoo, H.-J., & Watkins, M. T. (2013). Reduced hind limb

ischemia-reperfusion injury in Toll-like receptor-4 mutant mice is associated with decreased

neutrophil extracellular traps. Journal of Vascular Surgery, 58(6), 1627–36.

Olfert, I. M. (2015). Physiological capillary regression is not dependent on reducing VEGF

315

expression. Microcirculation (New York, N.Y. : 1994), 146–156. doi:10.1111/micc.12263

Onimaru, M., Yonemitsu, Y., Tanii, M., Nakagawa, K., Masaki, I., Okano, S., … Sueishi, K.

(2002). Fibroblast Growth Factor-2 Gene Transfer Can Stimulate Hepatocyte Growth Factor

Expression Irrespective of Hypoxia-Mediated Downregulation in Ischemic Limbs.

Circulation Research, 91(10), 923–930.

Ornitz, D. M., & Itoh, N. (2001). Fibroblast growth factors. Genome Biology, 2(3).

Ornitz, D. M., & Itoh, N. (2015). The Fibroblast Growth Factor signaling pathway. Wiley

Interdisciplinary Reviews. Developmental Biology, 4(3), 215–66. doi:10.1002/wdev.176

Ornitz, D. M., Xu, J., Colvin, J. S., McEwen, D. G., MacArthur, C. A., Coulier, F., … Goldfarb,

M. (1996). Receptor specificity of the fibroblast growth factor family. The Journal of

Biological Chemistry, 271(25), 15292–7.

Ortega, S., Ittmann, M., Tsang, S. H., Ehrlich, M., & Basilico, C. (1998). Neuronal defects and

delayed wound healing in mice lacking fibroblast growth factor 2. Proceedings of the

National Academy of Sciences, 95(10), 5672–5677.

Oses, P., Renault, M.-A., Chauvel, R., Leroux, L., Allières, C., Séguy, B., … Duplàa, C. (2009).

Mapping 3-dimensional neovessel organization steps using micro-computed tomography in

a murine model of hindlimb ischemia-brief report. Arteriosclerosis, Thrombosis, and

Vascular Biology, 29(12), 2090–2092.

Owen, J. L., & Mohamadzadeh, M. (2013). Macrophages and chemokines as mediators of

angiogenesis. Frontiers in Physiology, 4, 159. doi:10.3389/fphys.2013.00159

Ozaki, H., Okamoto, N., Ortega, S., Chang, M., Ozaki, K., Sadda, S., … Campochiaro, P. A.

316

(1998). Basic fibroblast growth factor is neither necessary nor sufficient for the development

of retinal neovascularization. The American Journal of Pathology, 153(3), 757–65.

Padua, R. R., & Kardami, E. (1993). Increased basic fibroblast growth factor (bFGF) accumulation

and distinct patterns of localization in isoproterenol-induced cardiomyocyte injury. Growth

Factors (Chur, Switzerland), 8(4), 291–306.

Paoni, N. F., Peale, F., Wang, F., Errett-Baroncini, C., Steinmetz, H., Toy, K., … Powell-Braxton,

L. (2002). Time course of skeletal muscle repair and gene expression following acute hind

limb ischemia in mice. Physiological Genomics, 11(3), 263–72.

Papetti, M., & Herman, I. M. (2002). Mechanisms of normal and tumor-derived angiogenesis. Am

J Physiol Cell Physiol, 282(5), C947–70. doi:10.1152/ajpcell.00389.2001 [doi]

Partanen, J. M., Schwartz, L., & Rossant, J. (1998). Opposite phenotypes of hypomorphic and

Y766 phosphorylation site mutations reveal a function for Fgfr1 in anteroposterior patterning

of mouse embryos. Genes & Development, 12(15), 2332–44.

Pasumarthi, K. B., Kardami, E., & Cattini, P. A. (1996). High and low molecular weight fibroblast

growth factor-2 increase proliferation of neonatal rat cardiac myocytes but have differential

effects on binucleation and nuclear morphology. Evidence for both paracrine and intracrine

actions of fibroblast grow. Circulation Research, 78(1), 126–136.

Pellieux, C., Foletti, A., Peduto, G., Aubert, J. F., Nussberger, J., Beermann, F., … Pedrazzini, T.

(2001). Dilated cardiomyopathy and impaired cardiac hypertrophic response to angiotensin

II in mice lacking FGF-2. The Journal of Clinical Investigation, 108(12), 1843–51.

Peng, X., Wang, J., Lassance-Soares, R. M., Najafi, A. H., Sood, S., Aghili, N., … Burnett, M. S.

317

(2011). Gender differences affect blood flow recovery in a mouse model of hindlimb

ischemia. American Journal of Physiology. Heart and Circulatory Physiology, 300(6),

H2027–34.

Peplow, P. V, & Chatterjee, M. P. (2013). A review of the influence of growth factors and

cytokines in in vitro human keratinocyte migration. Cytokine, 62(1), 1–21.

Persson, A. B. (2011). Vascular growth in health and disease. Frontiers in Molecular

Neuroscience, 4, 14.

Philippou, A., Maridaki, M., Theos, A., & Koutsilieris, M. (2012). Cytokines in Muscle Damage.

In G. S. Makowski (Ed.), Advances in Clinical Chemistry (Vol. Volume 58, pp. 49–87).

Elsevier.

Pintucci, G., Moscatelli, D., Saponara, F., Biernacki, P. R., Baumann, F. G., Bizekis, C., …

Mignatti, P. (2002). Lack of ERK activation and cell migration in FGF-2-deficient endothelial

cells. FASEB Journal : Official Publication of the Federation of American Societies for

Experimental Biology, 16(6), 598–600.

Pintucci, G., Yu, P.-J. J., Saponara, F., Kadian-Dodov, D. L., Galloway, A. C., & Mignatti, P.

(2005). PDGF-BB induces vascular smooth muscle cell expression of high molecular weight

FGF-2, which accumulates in the nucleus. Journal of Cellular Biochemistry, 95(6), 1292–

1300.

Piotrowicz, R. S., Ding, L., Maher, P. A., & Levin, E. G. (2001). Inhibition of Cell Migration by

24-kDa Fibroblast Growth Factor-2 Is Dependent upon the Estrogen Receptor. Journal of

Biological Chemistry, 276(6), 3963–3970.

318

Piotrowicz, R. S., Maher, P. A., & Levin, E. G. (1999). Dual activities of 22-24 kDA basic

fibroblast growth factor: inhibition of migration and stimulation of proliferation. Journal of

Cellular Physiology, 178(2), 144–53.

Piotrowicz, R. S., Martin, J. L., Dillman, W. H., & Levin, E. G. (1997). The 27-kDa Heat Shock

Protein Facilitates Basic Fibroblast Growth Factor Release from Endothelial Cells. Journal

of Biological Chemistry, 272(11), 7042–7047. Potente, M., Gerhardt, H., & Carmeliet, P.

(2011). Basic and therapeutic aspects of angiogenesis. Cell, 146(6), 873–87.

doi:10.1016/j.cell.2011.08.039

Powers, C. J., McLeskey, S. W., & Wellstein, A. (2000). Fibroblast growth factors, their receptors

and signaling. Endocr Relat Cancer, 7(3), 165–197.

Prats, H., Kaghad, M., Prats, Klagsbrun, M., Lélias, J. M., Liauzun, P., … Smith, J. A. (1989).

High molecular mass forms of basic fibroblast growth factor are initiated by alternative CUG

codons. Proceedings of the National Academy of Sciences of the United States of America,

86(6), 1836–40.

Presta, M., Andrés, G., Leali, D., Dell’Era, P., & Ronca, R. (2009). Inflammatory cells and

chemokines sustain FGF2-induced angiogenesis. European Cytokine Network, 20(2), 39–50.

doi:10.1684/ecn.2009.0155

Presta, M., Dell’Era, P., Mitola, S., Moroni, E., Ronca, R., & Rusnati, M. (2005). Fibroblast growth

factor/fibroblast growth factor receptor system in angiogenesis. Cytokine Growth Factor Rev,

16(2), 159–178.

Presta, M., Maier, J. A., Rusnati, M., & Ragnotti, G. (1989). Basic fibroblast growth factor is

released from endothelial extracellular matrix in a biologically active form. Journal of

319

Cellular Physiology, 140(1), 68–74.

Presta, M., Mitola, S., Leali, D., Nicoli, S., Moroni, E., Rusnati, M., … Rusnati, M. (2007).

Fibroblast Growth Factor-2 in Angiogenesis. Angiogenesis: An Integrative Approach From

Science to Medicine, 77–88.

Presta, M., Moscatelli, D., Joseph-Silverstein, J., & Rifkin, D. B. (1986). Purification from a

human hepatoma cell line of a basic fibroblast growth factor-like molecule that stimulates

capillary endothelial cell plasminogen activator production, DNA synthesis, and migration.

Molecular and Cellular Biology, 6(11), 4060–4066.

Prinzen, F. W., & Bassingthwaighte, J. B. (2000). Blood flow distributions by microsphere

deposition methods. Cardiovasc Res, 45(1), 13–21.

Prior, B. M., Lloyd, P. G., Ren, J., Li, H., Yang, H.-T., Laughlin, M. H., & Terjung, R. L. (2004).

Time course of changes in collateral blood flow and isolated vessel size and gene expression

after femoral artery occlusion in rats. American Journal of Physiology - Heart and

Circulatory Physiology, 287(6), H2434–47.

Pu, L. Q., Jackson, S., Lachapelle, K. J., Arekat, Z., Graham, A. M., Lisbona, R., … Symes, J. F.

(1994). A persistent hindlimb ischemia model in the rabbit. Journal of Investigative Surgery :

The Official Journal of the Academy of Surgical Research, 7(1), 49–60.

Pu, L. Q., Sniderman, A. D., Brassard, R., Lachapelle, K. J., Graham, A. M., Lisbona, R., & Symes,

J. F. (1993). Enhanced revascularization of the ischemic limb by angiogenic therapy.

Circulation, 88(1), 208–215.

Qi, J. H., Matsumoto, T., Huang, K., Olausson, K., Christofferson, R., & Claesson-Welsh, L.

320

(1999). Phosphoinositide 3 kinase is critical for survival, mitogenesis and migration but not

for differentiation of endothelial cells. Angiogenesis, 3(4), 371–80.

Quarto, N., Fong, K. D., & Longaker, M. T. (2005). Gene profiling of cells expressing different

FGF-2 forms. Gene, 356, 49–68.

Quarto, N., Talarico, D., Florkiewicz, R., & Rifkin, D. B. (1991). Selective expression of high

molecular weight basic fibroblast growth factor confers a unique phenotype to NIH 3T3 cells.

Cell Regulation, 2(9), 699–708.

Raballo, R., Rhee, J., Lyn-Cook, R., Leckman, J. F., Schwartz, M. L., & Vaccarino, F. M. (2000).

Basic Fibroblast Growth Factor (Fgf2) Is Necessary for Cell Proliferation and Neurogenesis

in the Developing Cerebral Cortex. J. Neurosci., 20(13), 5012–5023.

Radke, P. W., Heinl-Green, A., Frass, O. M., Post, M. J., Sato, K., Geddes, D. M., & Alton, E. W.

F. W. (2006). Evaluation of the porcine ameroid constrictor model of myocardial ischemia

for therapeutic angiogenesis studies. Endothelium: Journal Of Endothelial Cell Research,

13(1), 25–33.

Rajanayagam, M. A., Shou, M., Thirumurti, V., Lazarous, D. F., Quyyumi, A. A., Goncalves, L.,

… Unger, E. F. (2000). Intracoronary basic fibroblast growth factor enhances myocardial

collateral perfusion in dogs. J Am Coll Cardiol, 35(2), 519–526.

Rakue, H., Nakajima, H., Katoh, T., Usui, M., Amemiya, T., Miyagi, M., … Ibukiyama, C. (1998).

Low-Dose Basic Fibroblast Growth Factor and Vascular Endothelial Growth Factor for

Angiogenesis in Canine. Japanese Circulation Journal, 62(December), 933–939.

Raman, D., Sobolik-Delmaire, T., & Richmond, A. (2011). Chemokines in health and disease.

321

Experimental Cell Research, 317(5), 575–589.

Ratzka, A., Baron, O., & Grothe, C. (2011). FGF-2 deficiency does not influence FGF ligand and

receptor expression during development of the nigrostriatal system. PloS One, 6(8), e23564.

Ratzka, A., Baron, O., Stachowiak, M. K., & Grothe, C. (2012). Fibroblast growth factor 2

regulates dopaminergic neuron development in vivo. Journal of Neurochemistry, 122(1), 94–

105.

Reimer, K. A., Ideker, R. E., & Jennings, R. B. (1981). Effect of coronary occlusion site on

ischaemic bed size and collateral blood flow in dogs. Cardiovascular Research, 15(11), 668–

674.

Rey, S., Lee, K., Wang, C. J., Gupta, K., Chen, S., McMillan, A., … Semenza, G. L. (2009).

Synergistic effect of HIF-1alpha gene therapy and HIF-1-activated bone marrow-derived

angiogenic cells in a mouse model of limb ischemia. Proceedings of the National Academy

of Sciences of the United States of America, 106(48), 20399–404.

Ribatti, D. (2014). History of research on angiogenesis. Chemical Immunology and Allergy, 99,

1–14.

Rissanen, T. T., Markkanen, J. E., Arve, K., Rutanen, J., Kettunen, M. I., Vajanto, I., … Ylä-

Herttuala, S. (2003). Fibroblast growth factor 4 induces vascular permeability, angiogenesis

and arteriogenesis in a rabbit hindlimb ischemia model. FASEB Journal : Official Publication

of the Federation of American Societies for Experimental Biology, 17(1), 100–2.

Rissanen, T. T., Vajanto, I., Hiltunen, M. O., Rutanen, J., Kettunen, M. I., Niemi, M., … Ylä-

Herttuala, S. (2002). Expression of Vascular Endothelial Growth Factor and Vascular

322

Endothelial Growth Factor Receptor-2 (KDR/Flk-1) in Ischemic Skeletal Muscle and Its

Regeneration. The American Journal of Pathology, 160(4), 1393–1403.

Rivard, A., Fabre, J.-E., Silver, M., Chen, D., Murohara, T., Kearney, M., … Isner, J. M. (1999).

Age-Dependent Impairment of Angiogenesis. Circulation, 99(1), 111–120.

Rohovsky, S., Kearney, M., Pieczek, A., Rosenfield, K., Schainfeld, R., D’Amore, P. A., & Isner,

J. M. (1996). Elevated levels of basic fibroblast growth factor in patients with limb ischemia.

American Heart Journal, 132(5), 1015–9.

Romagnani, P., Lasagni, L., Annunziato, F., Serio, M., & Romagnani, S. (2004). CXC

chemokines: the regulatory link between inflammation and angiogenesis. Trends in

Immunology, 25(4), 201–9.

Rosenblatt, S., Irikura, K., Caday, C. G., Finklestein, S. P., & Moskowitz, M. A. (1994). Basic

Fibroblast Growth Factor Dilates Rat Pial Arterioles. Journal of Cerebral Blood Flow &

Metabolism, 14(1), 70–74.

Rosengart, T. K., Budenbender, K. T., Duenas, M., Mack, C. A., Zhang, Q.-X., & Isom, O. W.

(1997). Therapeutic angiogenesis: A comparative study of the angiogenic potential of acidic

fibroblast growth factor and heparin. Journal of Vascular Surgery, 26(2), 302–312.

Roth, D. A., McKirnan, M. D., Canestrelli, I., Gao, M. H., Dalton, N., Lai, N. C., … Hammond,

H. K. (2006). Intracoronary delivery of an adenovirus encoding fibroblast growth factor-4 in

myocardial ischemia: effect of serum antibodies and previous exposure to adenovirus. Human

Gene Therapy, 17(2), 230–8. doi:10.1089/hum.2006.17.230

Rubanyi, G. M. (2013). Mechanistic, technical, and clinical perspectives in therapeutic stimulation

323

of coronary collateral development by angiogenic growth factors. Molecular Therapy : The

Journal of the American Society of Gene Therapy, 21(4), 725–38.

Rudnicki, M. A., Le Grand, F., McKinnell, I., & Kuang, S. (2008). The molecular regulation of

muscle stem cell function. Cold Spring Harbor Symposia on Quantitative Biology, 73(0),

323–31. doi:10.1101/sqb.2008.73.064

Ruifrok, W.-P. T., de Boer, R. A., Iwakura, A., Silver, M., Kusano, K., Tio, R. A., & Losordo, D.

W. (2009). Estradiol-induced, endothelial progenitor cell-mediated neovascularization in

male mice with hind-limb ischemia. Vascular Medicine (London, England), 14(1), 29–36.

Rundle, C. H., Miyakoshi, N., Ramirez, E., Wergedal, J. E., Lau, K.-H. W., & Baylink, D. J.

(2002). Expression of the fibroblast growth factor receptor genes in fracture repair. Clinical

Orthopaedics and Related Research, (403), 253–63.

Rutherford, R. B., Baker, J. D., Ernst, C., Johnston, K. W., Porter, J. M., Ahn, S., & Jones, D. N.

(1997). Recommended standards for reports dealing with lower extremity ischemia: revised

version. J Vasc Surg, 26(3), 517–538.

Sabbieti, M. G., Agas, D., Marchetti, L., Coffin, J. D., Xiao, L., & Hurley, M. M. (2013). BMP-2

differentially modulates FGF-2 isoform effects in osteoblasts from newborn transgenic mice.

Endocrinology, 154(8), 2723–2733.

Sakhrani, N. M., & Padh, H. (2013). Organelle targeting: third level of drug targeting. Drug

Design, Development and Therapy, 7, 585–99.

Sambasivan, R., Yao, R., Kissenpfennig, A., Van Wittenberghe, L., Paldi, A., Gayraud-Morel, B.,

… Galy, A. (2011). Pax7-expressing satellite cells are indispensable for adult skeletal muscle

324

regeneration. Development (Cambridge, England), 138(17), 3647–56.

Santiago, J.-J., Ma, X., McNaughton, L. J., Nickel, B. E., Bestvater, B. P., Yu, L., … Kardami, E.

(2011). Preferential accumulation and export of high molecular weight FGF-2 by rat cardiac

non-myocytes. Cardiovascular Research, 89(1), 139–147.

Sarrazin, S., Lamanna, W. C., & Esko, J. D. (2011). Heparan sulfate proteoglycans. Cold Spring

Harbor Perspectives in Biology, 3(7).

Schaper, W. (2000). Quo vadis collateral blood flow? A commentary on a highly cited paper.

Cardiovascular Research, 45(1), 220–3. doi:10.1016/S0008-6363(99)00332-6

Schaper, W. (2009). Collateral circulation: past and present. Basic Research in Cardiology, 104(1),

5–21.

Scharner, J., & Zammit, P. S. (2011). The muscle satellite cell at 50: the formative years. Skeletal

Muscle, 1(1).

Schirmer, S. H., van Nooijen, F. C., Piek, J. J., & van Royen, N. (2009). Stimulation of collateral

artery growth: travelling further down the road to clinical application. Heart, 95(3), 191–197.

Schlessinger, J. (2000). Cell Signaling by Receptor Tyrosine Kinases. Cell, 103(2), 211–225.

Schlessinger, J., Plotnikov, A. N., Ibrahimi, O. A., Eliseenkova, A. V., Yeh, B. K., Yayon, A., …

Mohammadi, M. (2000). Crystal Structure of a Ternary FGF-FGFR-Heparin Complex

Reveals a Dual Role for Heparin in FGFR Binding and Dimerization. Molecular Cell, 6(3),

743–750.

Schmid, G. J., Kobayashi, C., Sandell, L. J., & Ornitz, D. M. (2009). Fibroblast growth factor

expression during skeletal fracture healing in mice. Developmental Dynamics : An Official

325

Publication of the American Association of Anatomists, 238(3), 766–74.

Schmittgen, T. D., & Livak, K. J. (2008). Analyzing real-time PCR data by the comparative C(T)

method. Nature Protocols, 3(6), 1101–8.

Scholz, D., Ziegelhoeffer, T., Helisch, A., Wagner, S., Friedrich, C., Podzuweit, T., & Schaper,

W. (2002). Contribution of arteriogenesis and angiogenesis to postocclusive hindlimb

perfusion in mice. J Mol Cell Cardiol, 34(7), 775–787.

Schultz, J. E. J., Witt, S. A., Nieman, M. L., Reiser, P. J., Engle, S. J., Zhou, M., … Doetschman,

T. (1999). Fibroblast growth factor-2 mediates pressure-induced hypertrophic response. The

Journal of Clinical Investigation, 104(6), 709–19.

Schumacher, B., Pecher, P., von Specht, B. U., & Stegmann, T. (1998). Induction of

neoangiogenesis in ischemic myocardium by human growth factors: first clinical results of a

new treatment of coronary heart disease. Circulation, 97(7), 645–50.

Seeger, M. A., & Paller, A. S. (2015). The Roles of Growth Factors in Keratinocyte Migration.

Advances in Wound Care, 4(4), 213–224. doi:10.1089/wound.2014.0540

Seghezzi, G., Patel, S., Ren, C. J., Gualandris, A., Pintucci, G., Robbins, E. S., … Mignatti, P.

(1998). Fibroblast growth factor-2 (FGF-2) induces vascular endothelial growth factor

(VEGF) expression in the endothelial cells of forming capillaries: an autocrine mechanism

contributing to angiogenesis. J Cell Biol, 141(7), 1659–1673. doi:10.1083/jcb.141.7.1659

Semenza, G. L. (2007). Vasculogenesis, angiogenesis, and arteriogenesis: mechanisms of blood

vessel formation and remodeling. Journal of Cellular Biochemistry, 102(4), 840 – 847.

Serhan, C. N., & Savill, J. (2005). Resolution of inflammation: the beginning programs the end.

326

Nature Immunology, 6(12), 1191–1197.

Sheehan, S. M., & Allen, R. E. (1999). Skeletal muscle satellite cell proliferation in response to

members of the fibroblast growth factor family and hepatocyte growth factor. Journal of

Cellular Physiology, 181(3), 499–506.

Sheikh, F., Sontag, D. P., Fandrich, R. R., Kardami, E., & Cattini, P. A. (2001). Overexpression

of FGF-2 increases cardiac myocyte viability after injury in isolated mouse hearts. American

Journal of Physiology. Heart and Circulatory Physiology, 280(3), H1039–1050.

Shen, M., Gao, J., Li, J., Su, J., Shen, Mei1, 2, & N. (2009). Effect of stimulation frequency on

angiogenesis and gene expression in ischemic skeletal muscle of rabbit. Canadian Journal of

Physiology & Pharmacology, 87(5), 396–401. doi:10.1139/y09-007

Shing, Y., Folkman, J., Haudenschild, C. C., Lund, D., Crum, R., & Klagsbrun, M. (1985).

Angiogenesis is stimulated by a tumor-derived endothelial cell growth factor. Journal of

Cellular Biochemistry, 29, 275–287. doi:10.1002/jcb.240290402

Shing, Y., Folkman, J., Sullivan, R., Butterfield, C., Murray, J., & Klagsbrun, M. (1984). Heparin

affinity: purification of a tumor-derived capillary endothelial cell growth factor. Science (New

York, N.Y.), 223(4642), 1296–9.

Shireman, P. K. (2007). The chemokine system in arteriogenesis and hind limb ischemia. Journal

of Vascular Surgery: Official Publication, the Society for Vascular Surgery [and]

International Society for Cardiovascular Surgery, North American Chapter, 45 Suppl A(6),

A48–56. doi:10.1016/j.jvs.2007.02.030

Shireman, P. K., Contreras-Shannon, V., Ochoa, O., Karia, B. P., Michalek, J. E., & McManus, L.

327

M. (2007). MCP-1 deficiency causes altered inflammation with impaired skeletal muscle

regeneration. Journal of Leukocyte Biology, 81(3), 775–85. doi:10.1189/jlb.0506356

Shireman, P. K., Contreras-Shannon, V., Reyes-Reyna, S. M., Robinson, S. C., & McManus, L.

M. (2006). MCP-1 parallels inflammatory and regenerative responses in ischemic muscle.

The Journal of Surgical Research, 134(1), 145–157. doi:10.1016/j.jss.2005.12.003

Shireman, P. K., & Quinones, M. P. (2005). Differential Necrosis Despite Similar Perfusion in

Mouse Strains after Ischemia. Journal of Surgical Research, 129(2), 242–250.

Silva, C., Fuxe, K., & Chadi, G. (2009). Involvement of astroglial fibroblast growth factor-2 and

microglia in the nigral 6-OHDA Parkinsonism and a possible role of glucocorticoid hormone

on the glial mediated local trophism and wound repair. In Journal of Neural Transmission,

Supplementa (Vol. 73, pp. 185–202). Vienna: Springer Vienna. doi:10.1007/978-3-211-

92660-4

Silvestre, J.-S., Mallat, Z., Tedgui, A., & Lévy, B. I. (2008). Post-ischaemic neovascularization

and inflammation. Cardiovascular Research, 78(2), 242–9. doi:10.1093/cvr/cvn027

Silvestre, J.-S., Smadja, D. M., & Lévy, B. I. (2013). Postischemic revascularization: from cellular

and molecular mechanisms to clinical applications. Physiological Reviews, 93(4), 1743–802.

Simons, M., Annex, B. H., Laham, R. J., Kleiman, N., Henry, T., Dauerman, H., … Chronos, N.

A. (2002). Pharmacological treatment of coronary artery disease with recombinant fibroblast

growth factor-2 double-blind, randomized, controlled clinical trial. Circulation., 105(7), 788–

93.

Skalli, O., Pelte, M. F., Peclet, M. C., Gabbiani, G., Gugliotta, P., Bussolati, G., … Orci, L. (1989).

328

Alpha-smooth muscle actin, a differentiation marker of smooth muscle cells, is present in

microfilamentous bundles of pericytes. The Journal of Histochemistry and Cytochemistry :

Official Journal of the Histochemistry Society, 37(3), 315–21.

Slatter, D. H. (Ed.). (2003). Textbook of small animal surgery, Volume 1 (3rd ed.). Philadelphia,

PA: Saunders.

Smith, C., Kruger, M. J., Smith, R. M., & Myburgh, K. H. (2008). The inflammatory response to

skeletal muscle injury: illuminating complexities. Sports Medicine (Auckland, N.Z.), 38(11),

947–969.

Sobue, T., Naganawa, T., Xiao, L., Okada, Y., Tanaka, Y., Ito, M., … Hurley, M. M. (2005). Over-

expression of fibroblast growth factor-2 causes defective bone mineralization and osteopenia

in transgenic mice. Journal of Cellular Biochemistry, 95(1), 83–94. doi:10.1002/jcb.20389

Sørensen, V., Nilsen, T., & Wiȩdłocha, A. (2006). Functional diversity of FGF-2 isoforms by

intracellular sorting. BioEssays, 28(5), 504–514. doi:10.1002/bies.20405

Speliotes, E. K., Caday, C. G., Do, T., Weise, J., Kowall, N. W., & Finkelstein, S. P. (1996).

Increased expression of basic fibroblast growth factor (bFGF) following focal cerebral

infarction in the rat. Molecular Brain Research, 39(1-2), 31–42.

Stabile, E., Burnett, M. S., Watkins, C., Kinnaird, T., Bachis, A., la Sala, A., … Fuchs, S. (2003).

Impaired arteriogenic response to acute hindlimb ischemia in CD4-knockout mice.

Circulation, 108(2), 205–210.

Stabile, E., Kinnaird, T., la Sala, A., Hanson, S. K., Watkins, C., Campia, U., … Burnett, M. S.

(2006). CD8+ T lymphocytes regulate the arteriogenic response to ischemia by infiltrating

329

the site of collateral vessel development and recruiting CD4+ mononuclear cells through the

expression of interleukin-16. Circulation, 113(1), 118–24.

Stachowiak, M. K., Birkaya, B., Aletta, J. M. M., Narla, S. T. T., Benson, C. a. A., Decker, B., &

Stachowiak, E. K. (2015). “Nuclear FGF Receptor-1 and CREB Binding Protein: An

Integrative Signaling Module.” Journal of Cellular Physiology, 230(5), 989–1002.

Stachowiak, M. K., Fang, X., Myers, J. M., Dunham, S. M., Berezney, R., Maher, P. A., &

Stachowiak, E. K. (2003). Integrative nuclear FGFR1 signaling INFS as a part of a universal

“feed-forward-and-gate” signaling module that controls cell growth and differentiation.

Journal of Cellular Biochemistry, 90(4), 662–691. doi:10.1002/jcb.10606

Stachowiak, M. K., Maher, P. A., & Stachowiak, E. K. (2007). Integrative nuclear signaling in cell

development--a role for FGF receptor-1. DNA and Cell Biology, 26(12), 811–826.

doi:10.1089/dna.2007.0664

Staton, C. A., Reed, M. W., & Brown, N. J. (2009). A critical analysis of current in vitro and in

vivo angiogenesis assays. International Journal of Experimental Pathology, 90(3), 195–221.

Steringer, J. P., Müller, H.-M., & Nickel, W. (2014). Unconventional Secretion of Fibroblast

Growth Factor 2-A Novel Type of Protein Translocation across Membranes? Journal of

Molecular Biology. doi:10.1016/j.jmb.2014.07.012

Stockmann, C., Schadendorf, D., Klose, R., & Helfrich, I. (2014). The impact of the immune

system on tumor: angiogenesis and vascular remodeling. Frontiers in Oncology, 4, 69.

Suckow, M. A., Weisbroth, S. H., & Franklin, C. L. (Eds.). (2006). The Laboratory Rat. The

Handbook of Experimental Animals. Burlington, MA: Academic Press.

330

Sullivan, C. J. (2002). The role of fibroblast growth factor-2 (FGF2) in vascular remodeling and

adaptation. The University of Arizona.

Sullivan, C. J., Doetschman, T., & Hoying, J. B. (2002). Targeted disruption of the Fgf2 gene does

not affect vascular growth in the mouse ischemic hindlimb. Journal of Applied Physiology

(Bethesda, Md. : 1985), 93(6), 2009–17.

Sullivan, C. J., & Hoying, J. B. (2002). Flow-Dependent Remodeling in the Carotid Artery of

Fibroblast Growth Factor-2 Knockout Mice. Arteriosclerosis, Thrombosis, and Vascular

Biology, 22(7), 1100–1105.

Sun, Y. (2009). Myocardial repair/remodelling following infarction: roles of local factors.

Cardiovascular Research, 81(3), 482–90. doi:10.1093/cvr/cvn333

Tabata, H. (1997). Arterial gene transfer of acidic fibroblast growth factor for therapeutic

angiogenesis in vivo: critical role of secretion signal in use of naked DNA. Cardiovascular

Research, 35(3), 470–479.

Tang, G., Charo, D. N., Wang, R., Charo, I. F., & Messina, L. M. (2004). CCR2-/- knockout mice

revascularize normally in response to severe hindlimb ischemia. Journal of Vascular Surgery,

40(4), 786–95. doi:10.1016/j.jvs.2004.07.012

Tang, G. L., Chang, D. S., Sarkar, R., Wang, R., & Messina, L. M. (2005). The effect of gradual

or acute arterial occlusion on skeletal muscle blood flow, arteriogenesis, and inflammation in

rat hindlimb ischemia. Journal of Vascular Surgery, 41(2), 312–20.

Tanghetti, E., Ria, R., Dell’Era, P., Urbinati, C., Rusnati, M., Ennas, M. G., & Presta, M. (2002).

Biological activity of substrate-bound basic fibroblast growth factor (FGF2): recruitment of

331

FGF receptor-1 in endothelial cell adhesion contacts. Oncogene, 21(24), 3889–97.

Tashiro, Y., Nishida, C., Sato-Kusubata, K., Ohki-Koizumi, M., Ishihara, M., Sato, A., … Hattori,

K. (2012). Inhibition of PAI-1 induces neutrophil-driven neoangiogenesis and promotes

tissue regeneration via production of angiocrine factors in mice. Blood, 119(26), 6382–93.

Taverna, S., Ghersi, G., Ginestra, A., Rigogliuso, S., Pecorella, S., Alaimo, G., … Vittorelli, M.

L. (2003). Shedding of Membrane Vesicles Mediates Fibroblast Growth Factor-2 Release

from Cells. Journal of Biological Chemistry, 278(51), 51911–51919.

Tedesco, F. S., Dellavalle, A., Diaz-Manera, J., Messina, G., & Cossu, G. (2010). Repairing

skeletal muscle: regenerative potential of skeletal muscle stem cells. J. Clin. Invest., 120(1),

11–19.

Tekin, M., Oztürkmen Akay, H., Fitoz, S., Birnbaum, S., Cengiz, F. B., Sennaroğlu, L., … Duman,

D. (2008). Homozygous FGF3 mutations result in congenital deafness with inner ear

agenesis, microtia, and microdontia. Clinical Genetics, 73(6), 554–65.

Terranova, C., Narla, S. T., Lee, Y.-W., Bard, J., Parikh, A., Stachowiak, E. K., … Stachowiak,

M. K. (2015). Global Developmental Gene Programing Involves a Nuclear Form of

Fibroblast Growth Factor Receptor-1 (FGFR1). PloS One, 10(4), e0123380.

Teunissen, P. F. A., Horrevoets, A. J. G., & van Royen, N. (2012). The coronary collateral

circulation: genetic and environmental determinants in experimental models and humans.

Journal of Molecular and Cellular Cardiology, 52(4), 897–904.

Thomas, K. A., Rios-Candelore, M., & Fitzpatrick, S. (1984). Purification and characterization of

acidic fibroblast growth factor from bovine brain. Proceedings of the National Academy of

332

Sciences of the United States of America, 81(2), 357–61.

Thomas, K. A., Rios-Candelore, M., Giménez-Gallego, G., DiSalvo, J., Bennett, C., Rodkey, J.,

& Fitzpatrick, S. (1985). Pure brain-derived acidic fibroblast growth factor is a potent

angiogenic vascular endothelial cell mitogen with sequence homology to interleukin 1.

Proceedings of the National Academy of Sciences of the United States of America, 82(19),

6409–13.

Thomas-Mudge, R. J., Okada-Ban, M., Vandenbroucke, F., Vincent-Salomon, A., Girault, J.-M.,

Thiery, J.-P., & Jouanneau, J. (2004). Nuclear FGF-2 facilitates cell survival in vitro and

during establishment of metastases. Oncogene, 23(27), 4771–4779.

Tidball, J. G. (2005). Inflammatory processes in muscle injury and repair. American Journal of

Physiology. Regulatory, Integrative and Comparative Physiology, 288(2), R345–353.

Tidball, J. G. (2011). Mechanisms of muscle injury, repair, and regeneration. Comprehensive

Physiology, 1(4), 2029–62.

Tidball, J. G., & Villalta, S. A. (2010). Regulatory interactions between muscle and the immune

system during muscle regeneration. American Journal of Physiology. Regulatory, Integrative

and Comparative Physiology, 298(5), R1173–1187. doi:10.1152/ajpregu.00735.2009

Tiefenbacher, C. P., & Chilian, W. M. (1997). Basic fibroblast growth factor and heparin influence

coronary arteriolar tone by causing endothelium-dependent dilation. Cardiovascular

Research, 34(2), 411–417.

Tilan, J. U., Everhart, L. M., Abe, K., Kuo-Bonde, L., Chalothorn, D., Kitlinska, J., … Zukowska,

Z. (2013). Platelet neuropeptide Y is critical for ischemic revascularization in mice. FASEB

333

Journal : Official Publication of the Federation of American Societies for Experimental

Biology, 27(6), 2244–55. doi:10.1096/fj.12-213546

Timmer, M., Müller-Ostermeyer, F., Kloth, V., Winkler, C., Grothe, C., & Nikkhah, G. (2004).

Enhanced survival, reinnervation, and functional recovery of intrastriatal dopamine grafts co-

transplanted with Schwann cells overexpressing high molecular weight FGF-2 isoforms.

Experimental Neurology, 187(1), 118–36.

Timmer, M., Robben, S., Müller-Ostermeyer, F., Nikkhah, G., & Grothe, C. (2003). Axonal

Regeneration Across Long Gaps in Silicone Chambers Filled With Schwann Cells

Overexpressing High Molecular Weight FGF-2. Cell Transplantation, 265.

Tirziu, D., Moodie, K. L., Zhuang, Z. W., Singer, K., Helisch, A., Dunn, J. F., … Simons, M.

(2005). Delayed Arteriogenesis in Hypercholesterolemic Mice. Circulation, 112(16), 2501–

2509.

Tobe, T., Ortega, S., Luna, J. D., Ozaki, H., Okamoto, N., Derevjanik, N. L., … Campochiaro, P.

A. (1998). Targeted disruption of the FGF2 gene does not prevent choroidal

neovascularization in a murine model. The American Journal of Pathology, 153(5), 1641–6.

Touriol, C., Bornes, S., Bonnal, S., Audigier, S., Prats, H., Prats, A.-C., & Vagner, S. (2003).

Generation of protein isoform diversity by alternative initiation of translation at non-AUG

codons. Biology of the Cell / under the Auspices of the European Cell Biology Organization,

95(3-4), 169–178.

Tsuboi, R., Sato, C., Kurita, Y., Ron, D., Rubin, J. S., & Ogawa, H. (1993). Keratinocyte Growth

Factor (FGF-7) Stimulates Migration and Plasminogen Activator Activity of Normal Human

Keratinocytes. Journal of Investigative Dermatology, 101(1), 49–53.

334

Turner, N., & Grose, R. (2010). Fibroblast growth factor signalling: from development to cancer.

Nature Reviews. Cancer, 10(2), 116–129. doi:10.1038/nrc2780

Uaesoontrachoon, K., Wasgewatte Wijesinghe, D. K., Mackie, E. J., & Pagel, C. N. (2013).

Osteopontin deficiency delays inflammatory infiltration and the onset of muscle regeneration

in a mouse model of muscle injury. Disease Models & Mechanisms, 6(1), 197–205.

Unger, E. F., Banai, S., Shou, M., Lazarous, D. F., Jaklitsch, M. T., Scheinowitz, M., … Epstein,

S. E. (1994). Basic fibroblast growth factor enhances myocardial collateral flow in a canine

model. Am J Physiol, 266(4 Pt 2), H1588–95.

Unger, E. F., Goncalves, L., Epstein, S. E., Chew, E. Y., Trapnell, C. B., Cannon, R. O., … Stiber,

J. A. (2000). Effects of a single intracoronary injection of basic fibroblast growth factor in

stable angina pectoris. The American Journal of Cardiology, 85(12), 1414–1419.

Valtink, M., Knels, L., Stanke, N., Engelmann, K., Funk, R. H., & Lindemann, D. (2012).

Overexpression of human HMW FGF-2 but not LMW FGF-2 reduces the cytotoxic effect of

lentiviral gene transfer in human corneal endothelial cells. Invest Ophthalmol Vis Sci, 53(6),

3207–3214. doi:iovs.12-9423 [pii] 10.1167/iovs.12-9423 [doi]

Van De Craen, B., Declerck, P. J., & Gils, A. (2012). The Biochemistry, Physiology and

Pathological roles of PAI-1 and the requirements for PAI-1 inhibition in vivo. Thrombosis

Research, 130(4), 576–585. van Oostrom, M. C., van Oostrom, O., Quax, P. H. a, Verhaar, M. C., & Hoefer, I. E. (2008).

Insights into mechanisms behind arteriogenesis: what does the future hold? Journal of

Leukocyte Biology, 84(6), 1379–91. doi:10.1189/jlb.0508281

335

van Royen, N., Piek, J. J., Schaper, W., & Fulton, W. F. (2009). A Critical Review of Clinical

Arteriogenesis Research. Journal of the American College of Cardiology, 55(1), 17–25.

Van Weel, V., van Tongeren, R. B., van Hinsbergh, V. W. M., van Bockel, J. H., & Quax, P. H.

A. (2008). Vascular Growth in Ischemic Limbs: A Review of Mechanisms and Possible

Therapeutic Stimulation. Annals of Vascular Surgery, 22(4), 582–597.

Virag, J. I., Rolle, M. L., Reece, J., Hardouin, S., Feigl, E. O., & Murry, C. E. (2007). Fibroblast

growth factor-2 regulates myocardial infarct repair: effects on cell proliferation, scar

contraction, and ventricular function. Am J Pathol, 171(5), 1431–1440.

Vogel, C., & Marcotte, E. M. (2012). Insights into the regulation of protein abundance from

proteomic and transcriptomic analyses. Nature Reviews. Genetics, 13(4), 227–32. von Maltzahn, J., Jones, A. E., Parks, R. J., & Rudnicki, M. A. (2013). Pax7 is critical for the

normal function of satellite cells in adult skeletal muscle. Proceedings of the National

Academy of Sciences of the United States of America, 110(41), 16474–9.

Walgenbach, K. J., Gratas, C., Shestak, K. C., & Becker, D. (1995). Ischaemia-induced expression

of bFGF in normal skeletal muscle: a potential paracrine mechanism for mediating

angiogenesis in ischaemic skeletal muscle. Nat Med, 1(5), 453–459.

Wårdell, K., Jakobsson, A., & Nilsson, G. E. (1993). Laser Doppler perfusion imaging by dynamic

light scattering. IEEE Transactions on Bio-Medical Engineering, 40(4), 309–16.

Watanabe, E., Smith, D. M. M., Sun, J., Smart, F. W. W., Delcarpio, J. B. B., Roberts, T. B. B.,

… Claycomb, W. C. C. (1998). Effect of basic fibroblast growth factor on angiogenesis in

the infarcted porcine heart. Basic Research in Cardiology, 93(1), 30–37.

336

Waters, R. E., Terjung, R. L., Peters, K. G., & Annex, B. H. (2004). Preclinical models of human

peripheral arterial occlusive disease: implications for investigation of therapeutic agents.

Journal of Applied Physiology (Bethesda, Md. : 1985), 97(2), 773–780.

Weber, K. T., Sun, Y., Bhattacharya, S. K., Ahokas, R. A., & Gerling, I. C. (2013). -

mediated mechanisms of pathological remodelling of the heart. Nature Reviews Cardiology,

10(1), 15–26.

Wempe, F., Lindner, V., & Augustin, H. G. (1997). Basic fibroblast growth factor (bFGF)

regulates the expression of the CC chemokine monocyte chemoattractant protein-1 (MCP-1)

in autocrine-activated endothelial cells. Arterioscler Thromb Vasc Biol, 17(11), 2471–2478.

Werner, S., Unsicker, K., & von Bohlen und Halbach, O. (2011). Fibroblast growth factor-2

deficiency causes defects in adult hippocampal neurogenesis, which are not rescued by

exogenous fibroblast growth factor-2. Journal of Neuroscience Research, 89(10), 1605–17.

Westvik, T. S., Fitzgerald, T. N., Muto, A., Maloney, S. P., Pimiento, J. M., Fancher, T. T., …

Dardik, A. (2009). Limb ischemia after iliac ligation in aged mice stimulates angiogenesis

without arteriogenesis. Journal of Vascular Surgery, 49(2), 464–473.

White, G. E., & Greaves, D. R. (2009). Fractalkine: one chemokine, many functions. Blood,

113(4), 767–8. doi:10.1182/blood-2008-11-189860

Wieder, R., Wang, H., Shirke, S., Wang, Q., Menzel, T., Feirt, N., … Gabrilove, J. L. (1997). Low

level expression of basic FGF upregulates Bcl-2 and delays apoptosis, but high intracellular

levels are required to induce transformation in NIH 3T3 cells. Growth Factors (Chur,

Switzerland), 15(1), 41–60.

337

Wietecha, M. S., Cerny, W. L., & DiPietro, L. A. (2013). Mechanisms of vessel regression: toward

an understanding of the resolution of angiogenesis. Current Topics in Microbiology and

Immunology, 367(January), 3–32. doi:10.1007/82

Witzenbichler, B., Mahfoudi, A., Soubrier, F., Le Roux, A., Branellec, D., Schultheiss, H.-P., &

Isner, J. M. (2006). Intramuscular gene transfer of fibroblast growth factor-1 using improved

pCOR plasmid design stimulates collateral formation in a rabbit ischemic hindlimb model.

Journal of Molecular Medicine (Berlin, Germany), 84(6), 491–502. doi:10.1007/s00109-005-

0031-3

Woodbury, M. E., & Ikezu, T. (2013). Fibroblast Growth Factor-2 Signaling in Neurogenesis and

Neurodegeneration. Journal of Neuroimmune Pharmacology: The Official Journal of the

Society on NeuroImmune Pharmacology, 9(2), 92–101. doi:10.1007/s11481-013-9501-5

Wu, H. M., Yuan, Y., McCarthy, M., & Granger, H. J. (1996). Acidic and basic FGFs dilate

arterioles of skeletal muscle through a NO-dependent mechanism. The American Journal of

Physiology, 271(3 Pt 2), H1087–93.

Wu, Q.-F., Yang, L., Li, S., Wang, Q., Yuan, X.-B., Gao, X., … Zhang, X. (2012). Fibroblast

growth factor 13 is a microtubule-stabilizing protein regulating neuronal polarization and

migration. Cell, 149(7), 1549–64. doi:10.1016/j.cell.2012.04.046

Xiao, L., Esliger, A., & Hurley, M. M. (2013). Nuclear fibroblast growth factor 2 (FGF2) isoforms

inhibit bone marrow stromal cell mineralization through FGF23/FGFR/MAPK in vitro.

Journal of Bone and Mineral Research : The Official Journal of the American Society for

Bone and Mineral Research, 28(1), 35–45. doi:10.1002/jbmr.1721

Xiao, L., Liu, P., Li, X., Doetschman, T., Coffin, J. D., Drissi, H., & Hurley, M. M. (2009).

338

Exported 18-kDa isoform of fibroblast growth factor-2 is a critical determinant of bone mass

in mice. The Journal of Biological Chemistry, 284(5), 3170–82.

Xiao, L., Liu, P., Sobue, T., Lichtler, A., Coffin, J. D., & Hurley, M. M. (2003). Effect of

overexpressing fibroblast growth factor 2 protein isoforms in osteoblastic ROS 17/2.8 cells.

Journal of Cellular Biochemistry, 89(6), 1291–301.

Xiao, L., Naganawa, T., Lorenzo, J., Carpenter, T. O., Coffin, J. D., & Hurley, M. M. (2010).

Nuclear isoforms of fibroblast growth factor 2 are novel inducers of hypophosphatemia via

modulation of FGF23 and KLOTHO. The Journal of Biological Chemistry, 285(4), 2834–

2846.

Xiao, L., Ueno, D., Catros, S., Homer-Bouthiette, C., Charles, L., Kuhn, L., & Hurley, M. M.

(2014). Fibroblast Growth Factor-2 Isoform (Low Molecular Weight/18 kDa)

Overexpression in Preosteoblast Cells Promotes Bone Regeneration in Critical Size Calvarial

Defects in Male Mice. Endocrinology, 155(3), 965–974. doi:10.1210/en.2013-1919

Yablonka-Reuveni, Z. (2011). The skeletal muscle satellite cell: still young and fascinating at 50.

The Journal of Histochemistry and Cytochemistry : Official Journal of the Histochemistry

Society, 59(12), 1041–59.

Yablonka-Reuveni, Z., Danoviz, M. E., Phelps, M., & Stuelsatz, P. (2015). Myogenic-specific

ablation of Fgfr1 impairs FGF2-mediated proliferation of satellite cells at the myofiber niche

but does not abolish the capacity for muscle regeneration. Frontiers in Aging Neuroscience,

7, 85.

Yablonka-Reuveni, Z., Day, K., Vine, A., & Shefer, G. (2008). Defining the transcriptional

signature of skeletal muscle stem cells. Journal of Animal Science, 86(14 Suppl), E207–16.

339

Yamada, P. M., & Lee, K.-W. (2009). Perspectives in mammalian IGFBP-3 biology: local vs.

systemic action. American Journal of Physiology. Cell Physiology, 296(5), C954–76.

Yamaguchi, T. P., Harpal, K., Henkemeyer, M., & Rossant, J. (1994). fgfr-1 is required for

embryonic growth and mesodermal patterning during mouse . Genes &

Development, 8(24), 3032–44.

Yanagisawa-Miwa, A., Uchida, Y., Nakamura, F., Tomaru, T., Kido, H., Kamijo, T., …

Kurashima, C. (1992). Salvage of infarcted myocardium by angiogenic action of basic

fibroblast growth factor. Science, 257(5075), 1401–3.

Yang, H.-T., Deschenes, M. R., Ogilvie, R. W., & Terjung, R. L. (1996). Basic Fibroblast Growth

Factor Increases Collateral Blood Flow in Rats With Femoral Arterial Ligation. Circ Res,

79(1), 62–69.

Yang, Q., Hewett, T. E., Klevitsky, R., Sanbe, A., Wang, X., & Robbins, J. (2001). PKA-

dependent phosphorylation of cardiac myosin binding protein C in transgenic mice.

Cardiovascular Research, 51(1), 80–88.

Yang, X., Liaw, L., Prudovsky, I., Brooks, P. C., Vary, C., Oxburgh, L., & Friesel, R. (2015).

Fibroblast growth factor signaling in the vasculature. Current Atherosclerosis Reports, 17(6),

509.

Yaylaoglu, M. B., Titmus, A., Visel, A., Alvarez-Bolado, G., Thaller, C., & Eichele, G. (2005).

Comprehensive expression atlas of fibroblast growth factors and their receptors generated by

a novel robotic in situ hybridization platform. Developmental Dynamics : An Official

Publication of the American Association of Anatomists, 234(2), 371–86.

340

Yen, T. T. H., Thao, D. T. P., & Thuoc, T. L. (2014). An overview on keratinocyte growth factor:

from the molecular properties to clinical applications. Protein and Peptide Letters, 21(3),

306–17.

Yin, H., Price, F., & Rudnicki, M. A. (2013). Satellite Cells and the Muscle Stem Cell Niche.

Physiological Reviews, 93(1), 23–67.

Yoshida, T., Ishimaru, K., Sakamoto, H., Yokota, J., Hirohashi, S., Igarashi, K., … Terada, M.

(1994). Angiogenic activity of the recombinant hst-1 protein. Cancer Letters, 83(1-2), 261–

268.

Yoshida, T., Miyagawa, K., Odagiri, H., Sakamoto, H., Little, P. F., Terada, M., & Sugimura, T.

(1987). Genomic sequence of hst, a transforming gene encoding a protein homologous to

fibroblast growth factors and the int-2-encoded protein. Proceedings of the National Academy

of Sciences of the United States of America, 84(20), 7305–9.

Yu, P.-J., Ferrari, G., Galloway, A. C., Mignatti, P., & Pintucci, G. (2007). Basic fibroblast growth

factor (FGF-2): the high molecular weight forms come of age. Journal of Cellular

Biochemistry, 100(5), 1100–8.

Yurek, D. M., Fletcher, A. M., Peters, L. E., & Cass, W. A. (2010). Strain difference in the up-

regulation of FGF-2 protein following a neurotoxic lesion of the nigrostriatal pathway.

Neurochemical Research, 35(4), 531–9.

Zagorchev, L., Oses, P., Zhuang, Z. W., Moodie, K. L., Mulligan-Kehoe, M. J., Simons, M., &

Couffinhal, T. (2010). Micro computed tomography for vascular exploration. J Angiogenes

Res, 2(1), 7.

341

Zakrzewska, M., Marcinkowska, E., & Wiedłocha, A. (2008). FGF-1: from biology through

engineering to potential medical applications. Critical reviews in clinical laboratory sciences

(Vol. 45). doi:10.1080/10408360701713120

Zakrzewska, M., Wiedłocha, A., Szlachcic, A., Krowarsch, D., Otlewski, J., & Olsnes, S. (2009).

Increased protein stability of FGF1 can compensate for its reduced affinity for heparin. The

Journal of Biological Chemistry, 284(37), 25388–403.

Zhan, X., Bates, B., Hu, X. G., & Goldfarb, M. (1988). The human FGF-5 oncogene encodes a

novel protein related to fibroblast growth factors. Molecular and Cellular Biology, 8(8),

3487–95.

Zhan, X., Culpepper, A., Reddy, M., Loveless, J., & Goldfarb, M. (1987). Human oncogenes

detected by a defined medium culture assay. Oncogene, 1(4), 369–76.

Zhang, J., Cao, R., Zhang, Y., Jia, T., Cao, Y., & Wahlberg, E. (2009). Differential roles of

PDGFR-alpha and PDGFR-beta in angiogenesis and vessel stability. FASEB Journal :

Official Publication of the Federation of American Societies for Experimental Biology, 23(1),

153–63.

Zhang, X., Bao, L., Yang, L., Wu, Q., & Li, S. (2012). Roles of intracellular fibroblast growth

factors in neural development and functions. Science China. Life Sciences, 55(12), 1038–44.

Zhang, X., Ibrahimi, O. A., Olsen, S. K., Umemori, H., Mohammadi, M., & Ornitz, D. M. (2006).

Receptor specificity of the fibroblast growth factor family. The complete mammalian FGF

family. The Journal of Biological Chemistry, 281(23), 15694–700.

Zhao, P., & Hoffman, E. P. (2004). Embryonic myogenesis pathways in muscle regeneration.

342

Developmental Dynamics: An Official Publication of the American Association of

Anatomists, 229(2), 380–392. doi:10.1002/dvdy.10457

Zhao, X.-C., Zhang, L.-M., Tong, D.-Y., An, P., Jiang, C., Zhao, P., … Wang, J. (2013). Propofol

increases expression of basic fibroblast growth factor after transient cerebral ischemia in rats.

Neurochemical Research, 38(3), 530–7.

Zhou, M., Sutliff, R. L., Paul, R. J., Lorenz, J. N., Hoying, J. B., Haudenschild, C. C., …

Doetschman, T. (1998). Fibroblast growth factor 2 control of vascular tone. Nat Med, 4(2),

201–207.

Zhu, X., Komiya, H., Chirino, A., Faham, S., Fox, G. M., Arakawa, T., … Rees, D. C. (1991).

Three-dimensional structures of acidic and basic fibroblast growth factors. Science (New

York, N.Y.), 251(4989), 90–3.

Ziegler, M. a, Distasi, M. R., Bills, R. G., Miller, S. J., Alloosh, M., Murphy, M. P., … Akingba,

A. G. (2010). Marvels, mysteries, and misconceptions of vascular compensation to peripheral

artery occlusion. Microcirculation, 17(1), 3–20.

Zittermann, S. I., & Issekutz, A. C. (2006a). Basic fibroblast growth factor (bFGF, FGF-2)

potentiates leukocyte recruitment to inflammation by enhancing endothelial adhesion

molecule expression. The American Journal of Pathology, 168(3), 835–46.

Zittermann, S. I., & Issekutz, A. C. (2006b). Endothelial growth factors VEGF and bFGF

differentially enhance monocyte and neutrophil recruitment to inflammation. Journal of

Leukocyte Biology, 80(2), 247–57.

343

344

APPENDIX

345

346