insights into fracture repair using a murine model …...identified. fracture repair is delayed...

141
INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL OF OSTEOFIBROUS DYSPLASIA by Wei Xiang Xie A thesis submitted in conformity with the requirements for the degree of Master of Science University of Toronto © Copyright by Wei Xiang (William) Xie (2019)

Upload: others

Post on 24-Aug-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

i

INSIGHTS INTO FRACTURE REPAIR USING A MURINE

MODEL OF OSTEOFIBROUS DYSPLASIA

by

Wei Xiang Xie

A thesis submitted in conformity with the requirements

for the degree of Master of Science

University of Toronto

© Copyright by Wei Xiang (William) Xie (2019)

Page 2: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

ii

Abstract

Insights into Fracture Repair Using a Murine Model of

Osteofibrous Dysplasia

Wei Xiang Xie

Master of Science

Institute of Medical Science

University of Toronto

2019

Osteofibrous dysplasia (OFD) is a rare disease characterized by the development of

radiolucent lesions at the tibial periosteal surface, where it causes non-healing fractures.

We previously identified gain-of-function mutations in the MET gene as a cause for

OFD. Fracture tissue samples from OFD patients exhibit aberrant MET activity and

delays in osteoblast differentiation. We hypothesized gain-of-function MET

mutations result in delayed bone repair ability due to reduced osteoblast

differentiation. MetΔ15-HET mice exhibit aberrant and prolonged upregulation of MET

signaling and total β-catenin levels similar to OFD patients. MetΔ15-HET osteoblasts

demonstrate a differentiation defect in vitro though no gross skeletal defects were

identified. Fracture repair is delayed in MetΔ15-HET mice, with decreased bone

formation 2-weeks post fracture-inducing surgery. Our data is consistent with a novel

role for MET-mediated signaling regulating osteogenesis and open up the possibility of

modulating the MET pathway to augment fracture healing.

Page 3: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

iii

Acknowledgments

I would like to thank my everyone who has helped me throughout the completion of this

degree. Firstly, I would like to thank my supervisor Dr. Peter Kannu, for allowing me the

opportunity to pursue this degree and project and assisting in the writing of this thesis.

Secondly, I would like to thank my Program Advisory Committee Members, Drs. Morris

Manolson and Jane Mitchell for their insights, guidance and constructive criticism along

the way.

I would like to thank Dr. Kashif Ahmed for assisting and guiding me through the design

and execution of the experiments and always making himself available to assist me

when needed. Our summer student Lisa Vi in assistance with the skeletal preps and

sectioning of histological slides. I would also like to thank The Centre for

Phenogenomics for assistance in assisting with colony management.

Personally, I would like to thank my friends and family for their continuous support and

encouragement in pursuing and accomplishing my goals. Without them I would not be

the person I am today – for that I am eternally grateful.

Page 4: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

iv

Contributions

William Xie (author) solely prepared this thesis. All aspects of this thesis, including

planning, execution, analysis and writing of the original research contained within was

performed by the author. The contributions by other individuals are formally and

inclusively acknowledged:

Dr. Peter Kannu (Supervisor and Committee Member) – Mentorship, laboratory space,

guidance on the planning, execution, analysis and presentation of all experiments in

addition to thesis preparation.

Dr. Jane Mitchell (Committee Member) – Mentorship, guidance in interpretation of

results and thesis preparation.

Dr. Morris Manolson (Committee Member) – Mentorship, guidance in interpretation of

results and thesis preparation.

Drs. Irina Voronov and Ralph Zirngibl (Collaborators) – Assistance in execution of

experiments and analysis of results in Figure 4.4.

Dr. Simon Kelley (Collaborator) – Assistance in performing the tibial fracture surgeries in

Figure 4.10.

Page 5: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

v

Table of Contents

Acknowledgments ................................................................................................................ iii

Contributions ........................................................................................................................ iv

List of Figures ........................................................................................................................ ix

List of Abbreviations ............................................................................................................. xi

1 Introduction ........................................................................................................................ 1 Figure 1A. Protein expression analysis of healthy bone versus osteofibrous dysplasia bone ............................. 3 Figure 1B. Regulation of MET Signaling ............................................................................................................... 4

1.1 Osteofibrous Dysplasia ................................................................................................................. 5 Figure 1C. Osteofibrous Dysplasia lesions at the tibia ......................................................................................... 8 Figure 1D. Osteofibrous Dysplasia fractures at the tibia ..................................................................................... 9 Figure 1E. Summary of Characteristics of Osteofibrous Dysplasia .................................................................... 10 Figure 1F. Genetic Summary of Osteofibrous Dysplasia .................................................................................... 11 Figure 1G. Zonal Architecture of Osteofibrous Dysplasia Lesions ..................................................................... 12

1.2 Neurofibromatosis Type 1 (NF-1) ................................................................................................ 13

1.3 Dysregulation of MET to METΔ14 ratios causes disease: Osteofibrous Dysplasia ....................... 14

1.4 Characterization of Mesenchymal-Epithelial Transition Gene’s Structure .................................. 14

1.5 MET Signaling and Function ........................................................................................................ 16

1.6 MET’s Role in Development ........................................................................................................ 17

1.7 CBL-Dependent Regulation of MET Signaling .............................................................................. 17

1.8 Other Means of Regulating MET Signaling .................................................................................. 18

1.9 Mouse Model of Osteofibrous Dysplasia .................................................................................... 19 Figure 1H. MET Exon 14 Skipping Mutation....................................................................................................... 20

1.10 Chondrocyte Differentiation ..................................................................................................... 21

1.11 Osteoblast Differentiation ........................................................................................................ 22

1.11 Osteocyte Differentiation ......................................................................................................... 23

1.12 Osteoclast Differentiation ........................................................................................................ 24 Figure 1I. Stages of Osteoblast Differentiation .................................................................................................. 25

1.13 Selected Molecular Factors Affecting Osteoblast Differentiation ............................................. 26

1.14 Key Signaling Pathways Involved in Bone ................................................................................. 26 Figure 1J. Summary of Genes and Factors Affecting Osteoblast Differentiation ............................................... 30

1.15 Bone Development ................................................................................................................... 31 Figure 1K. Overview of Intramembranous and Endochondral Ossification ....................................................... 32

1.16 Fracture Healing ....................................................................................................................... 33 Figure 1L. Overview of Stages of Fracture Repair .............................................................................................. 35 Figure 1M. Genes Involved in Fracture Repair ................................................................................................... 36

1.17 What is known about the MET-HGF pathway in Bone .............................................................. 37

Page 6: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

vi

2 Research Aims, Hypothesis, and Summary Plan ................................................................. 38

2.1 Rationale .................................................................................................................................... 39

2.2 Hypothesis .................................................................................................................................. 39

2.3 Objectives ................................................................................................................................... 40

2.4 Clinical Significance..................................................................................................................... 42

3 Methods ........................................................................................................................... 43

3.1 Generation of a Met exon 15 splice donor mutant allele by CRISPR/Cas9-mediated genome editing .............................................................................................................................................. 43

3.2 Embryo Microinjection ............................................................................................................... 43

3.3 Confirmation of Exon 15 Skipping ............................................................................................... 43

3.4 Maintenance and Genotyping of Mice........................................................................................ 44

......................................................................................................................................................... 45

......................................................................................................................................................... 45

3.5 Murine Embryonic Fibroblast Generation................................................................................... 45

3.6 MET Pathway Analysis ................................................................................................................ 46

3.7 Activation of MET pathway ........................................................................................................ 46

3.8 Western Blot Analysis ................................................................................................................. 46

3.9 Colony Forming Unit – Fibroblast ............................................................................................... 47

3.10 Colony Forming Unit – ALP and Osteoblast............................................................................... 47

3.11 RT-PCR: Gene Expression Analysis ............................................................................................ 48

3.12 Skeletal Prep Staining ............................................................................................................... 48

3.13 Epiphyseal Growth Plate Staining ............................................................................................. 49

3.14 Visualization and Measurement of Long Bones and Epiphyseal Growth Plates ........................ 49

3.15 Tibia Fracture Generation ......................................................................................................... 50

3.16 Histological Staining of Fracture Callus ..................................................................................... 51

3.17 Statistical Analyses ................................................................................................................... 51

4 Results .............................................................................................................................. 53

4.1 Met Δ15-HET mouse embryonic fibroblast exhibit higher levels of MET protein and upregulated MET signaling ................................................................................................................................... 53

Figure 4.1. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts......................... 55

4.2 Met Δ15-HET mouse embryonic fibroblast exhibit upregulated and prolonged MET signaling after HGF stimulation................................................................................................................................ 56

Figure 4.2. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes............................................................................................ 58 Figure 4.2 Contd. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes.................................................................................... 59

Page 7: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

vii

Figure 4.2 Contd. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes.................................................................................... 61

4.3 Met Δ15-HET MEFs exhibit upregulated levels of β-catenin.......................................................... 62 Figure 4.3 Met Δ15-HET MEFs exhibit gene expression dysregulation resulting in upregulation of β-catenin ... 64 Figure 4.3 Contd. ................................................................................................................................................ 65 Figure 4.4. Met Δ15-HET bone marrow stromal cells exhibit reduced mineralization ability ........................... 68 Figure 4.4 Contd. Met Δ15-HET bone marrow stromal cells exhibit reduced mineralization ability ................... 69

4.5 Mature Met Δ15-HET osteoblasts display dysregulation of osteoblast specific gene markers ...... 70 Figure 4.5 Met Δ15-HET mature osteoblasts display dysregulation of osteoblast specific gene markers ........... 73

4.6 Body weight and preliminary comparisons of WT and Met Δ15-HET mice ................................... 74 Figure 4.6 Body weight and preliminary comparisons of WT and Met Δ15-HET mice ........................................ 75 Figure 4.6 Contd. ................................................................................................................................................ 76 Figure 4.6 Contd. ................................................................................................................................................ 77 Figure 4.6 Contd. ................................................................................................................................................ 78

4.7 Skeletal Staining of P5 WT and Met Δ15-HET mice ....................................................................... 79 Figure 4.7. 5-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates.......................................................................................................................................................... 82 Figure 4.7. 5-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates.......................................................................................................................................................... 83

4.8 Skeletal Staining of P21 WT and Met Δ15-HET mice ..................................................................... 84 Figure 4.8. 21-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates.......................................................................................................................................................... 87 Figure 4.8. 21-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates.......................................................................................................................................................... 88

4.9 Histological analysis of P21 WT and Met Δ15-HET mice epiphyseal growth plate......................... 89 Figure 4.9 Met Δ15-HET mice exhibit no differences in epiphyseal growth plate lengths in comparison to WT littermates.......................................................................................................................................................... 90 Figure 4.9 Contd. Met Δ15-HET mice exhibit no differences in epiphyseal growth plate lengths in comparison to WT littermates ............................................................................................................................................... 91

4.10 Overactivation of MET signaling impairs fracture healing ........................................................ 92 Figure 4.10 Aberrant MET signaling impairs fracture healing in 12-week old mice .......................................... 93 Figure 4.10 Contd. .............................................................................................................................................. 94 Figure 4.10 Contd. .............................................................................................................................................. 95 Figure 4.10 Contd. .............................................................................................................................................. 96 Figure 4.10 Contd. .............................................................................................................................................. 97

5 Discussion ......................................................................................................................... 98

5.1 MetΔ15-HET mice exhibit higher MET protein levels and upregulated MET signaling .................. 98

5.2 Overactivation of MET signaling causes osteoblast differentiation defects .............................. 100

5.3 MetΔ15-HET MEFs exhibit dysregulation of β-catenin activity ................................................... 100

5.4 Overactivation of MET signaling does not affect postnatal skeletal development ................... 101

5.5 MetΔ15-HET mice experience delayed fracture healing ability ................................................... 102

5.6 MetΔ15-HET osteoblasts display dysregulation of β-catenin ...................................................... 104

5.7 Limitations to the Study ........................................................................................................... 107

Page 8: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

viii

5.8 Proposed Mechanism for the Dysregulation of β-catenin Signaling in Osteoblast Differentiation ....................................................................................................................................................... 110

Figure 5.1. Proposed Mechanism for the Dysregulation of β-catenin Signaling in Osteoblast Differentiation ......................................................................................................................................................................... 111

6 Conclusions ..................................................................................................................... 112

7 Future Directions ............................................................................................................. 115

9 References ...................................................................................................................... 118

Page 9: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

ix

List of Figures

Figure 1A Protein expression analysis of healthy bone versus osteofibrous dysplasia bone …………………………………………………………………………………………….13

Figure 1B Regulation of MET Signaling……………………………………………………..19 Figure 1C Osteofibrous Dysplasia lesions at the tibia …………………………………….23

Figure 1D Osteofibrous Dysplasia lesions at the tibia……………………………………..24

Figure 1E Summary of Characteristics of Osteofibrous Dysplasia ………………………25 Figure 1F Genetic Summary of Osteofibrous Dysplasia…………………………………..26

Figure 1G Zonal Architecture of Osteofibrous Dysplasia………………………………….27

Figure 1H MET Exon Skipping Mutation…………………………………………………….31 Figure 1I Overview of Stages of Osteoblast Differentiation……………………………….36

Figure 1J Summary of Genes and Factors Affecting Osteoblast Differentiation………,.41

Figure 1K Overview of Intramembranous and Endochondral Ossification………………43 Figure 1L Overview of Stages of Fracture Repair………………………………………….47

Figure 1M Genes Involved in Fracture Repair……………………………………………...48 Figure 4.1 Protein expression analysis of WT and Met Δ15-HET………………………….66

Figure 4.2 Protein expression analysis of WT and Met Δ15-HET mouse embryonic

fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes ……………..69 Figure 4.13 Met Δ15-HET BMSCs exhibit differential gene expression at day 21 of

osteoblast differentiation………………………………………………………………………75

Figure 4.3 Met Δ15-HET MEFs exhibit gene expression dysregulation resulting in

upregulation of β-catenin……………………………………………………………………...82

Figure 4.4 Met Δ15-HET bone marrow stromal cell exhibit reduced mineralization

ability…………………………………………………………………………………………….87

Figure 4.5 Mature Met Δ15-HET osteoblasts display dysregulation of osteoblast specific

gene markers…………………………………………………………………………………..91

Page 10: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

x

Figure 4.6 Body weight and preliminary comparisons of WT and Met Δ15-HET mice…..93

Figure 4.7 5-day old Met Δ15-HET mice exhibit no differences in long bone lengths in

comparison WT littermates………………………………………………………………….101

Figure 4.8 21-day old Met Δ15-HET mice exhibit no differences in long bone lengths in

comparison WT littermates………………………………………………………………….106

Figure 4.9 Met Δ15-HET mice exhibit no differences in epiphyseal growth plate lengths in

comparison to WT littermates…………………………………………………………….108

Figure 4.10 Aberrant MET signaling impairs fracture healing in 12-week old mice………………………………………………………………………………………...….111

Figure 5.1 Proposed Mechanism for the Dysregulation of β-catenin Signaling in Osteoblast Differentiation……………………………………………………………………111

Page 11: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

xi

List of Abbreviations

ALP Alkaline Phosphatase

BMP Bone Morphogenetic Protein

BMSC Bone Marrow Stromal Cell

BSP Bone Sialoprotein

CFU-F Colony Forming Unit–Fibroblast

CFU-O Colony Forming Unit-Osteoblast

ColI Collagen Type I

ERK1/2 Extracellular Signal-Regulated Protein Kinases 1 and 2

HGF Hepatocyte Growth Factor

IL Interleukin

LRP6 Low Density Lipoprotein Receptor-Related Protein 6 MAPK Mitogen-

Activated Protein Kinases

MAPK Mitogen-Activated Protein Kinase

M-CSF Macrophage Colony Stimulating Factor

MET Mesenchymal-Epithelial Transition Gene

MMP Matrix Metalloproteinase

MSC Mesenchymal Stem Cell

OCN Osteocalcin

OPG Osteoprotegerin

OPN Osteopontin

OSX Osterix

PI3K Phosphatidylinositol-3-Kinase

RANK Receptor Activator of Nuclear Factor-κβ

RANKL Receptor Activator of Nuclear Factor-κβ Ligand

RUNX2 Runt-related transcription factor 2

SSC Skeletal Stem Cell

SOX9 SRY (Sex Determining Region Y)-Box-9

STAT1 Signal Transducers and Activators of Transcription 1 TD Thanatophoric

Dysplasia

Page 12: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

xii

TGF-β Transforming Growth Factor Beta

TNF-α Tumor Necrosis Factor-α

Page 13: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

1

1 Introduction

Osteofibrous dysplasia (OFD) is a rare (<0.2% of bone tumors) non-neoplastic condition

which can cause tibial bowing prior to weight bearing in affected individuals. OFD can

also progress to spontaneous non-healing fractures resulting in impaired mobility and

constant pain, significantly affecting quality of life. Through exome sequencing, our lab

and others have previously identified gain-of-function mutations in the tyrosine kinase

MET as the cause of the OFD in four unrelated familial cases. The gain of function MET

mutations results in upregulated MET levels, changes in downstream MET effectors and

changes in β-catenin levels (Fig 1A). We have also shown impaired osteoblast

mineralization and decreased osteoclast function in affected patient bone tissue (Gray

et al., 2015). However, the mechanistic explanation connecting these observations

remains to be clarified.

Surprisingly little is known about MET in fracture repair despite our identification of a

human single gene disorder manifesting abnormal bone repair secondary to gain-of-

function MET mutations. Interestingly, we have also found increased MET expression

in a sub-set of human fracture non-union tissue. Fracture non-union is defined as an

inability to form a bridging callus at the fracture site. Delays in fracture repair causing

non-unions are a significant health problem with approximately 10% of all long bone

fractures healing with delay and 7.5% resulting in non-unions, despite improved surgical

techniques (Mills, Aitken, & Simpson, 2017).

There is currently no cure for OFD, and treatment is limited to pain management of

affected individuals and management of symptoms. Understanding the underlying

mechanisms resulting in reduced bone healing ability may potentially uncover novel

therapeutic targets to treat OFD and improve bone repair in general. The knowledge

Page 14: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

2

gained from such an endeavor would be transferable to managing fracture non-unions

from a variety of other causes.

Here we look to determine the role of MET signalling in the osteoblast differentiation and

fracture healing. In order to study these processes, we have created a murine mouse

model which reproduces one of the gain-of-function MET mutations identified in OFD.

Our genetic mouse model appears phenotypically normal but has yet to be

characterized and studied in detail. The purpose of this project is to determine the

underlying signaling changes secondary to the Met mutation in our murine model,

characterize the skeletal phenotype and utilize the model to study fracture repair in vivo.

Page 15: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

3

Actin

β-catenin

Actin

AKT

p-AKT

Actin

MET

Control

Bone OFD

Bone

Figure 1A. Protein expression analysis of healthy bone versus osteofibrous dysplasia bone

Representative western blots of healthy and osteofibrous dysplasia patient bone protein expression analysis. Healthy bone samples were obtained from

the iliac crest of healthy patients during surgical intervention, while osteofibrous dysplasia bone samples were obtained during excision of affected lesional tissue. Osteofibrous dysplasia patients exhibit upregulated

MET protein along with upregulation of its downstream effectors AKT and β-catenin in comparison to healthy bone samples. In addition, osteofibrous dysplasia patients exhibit upregulated phosphorylation at AKT indicative of

upregulated MET signaling.

A

Page 16: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

4

Regulation of MET Signaling

Regulation of MET Signaling

Figure 1B. Regulation of MET Signaling MET receptor without HGF binding (A). HGF binding to wild-type MET receptor leads to receptor dimerization and auto-phosphorylation. CBL is a E3 ubiquitin ligase involved in downregulation of the receptor (B). Depicts the structure of the exon 14 exclusion in MET-HET

mice resulting in loss of the regulatory juxtamembrane domain and CBL binding site (Y1003) (C).

A B

C

Page 17: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

5

1.1 Osteofibrous Dysplasia Osteofibrous dysplasia (OFD) (OMIM #607278), also known as ossifying fibroma of the

long bone, is a developmental skeletal disorder characterized by benign fibro-osseous

lesions of the bone (Hitachi et al., 2018). First described as congenital osteitis fibrosa in

1921 by Frangenheim, Kempson later reported affected individuals displayed similar

physical characteristics to fibrous dysplasia. In 1976 Campanacci characterized the

condition as “osteofibrous dysplasia of the tibia and fibula” to describe its anatomical

position, origin of development, and histologic similarities to fibrous dysplasia (Park,

Unni, McLeod, & Pritchard, 1993).

OFD makes up approximately 0.2% of all primary bone tumors, commonly affecting

children under the age of 10, with no sex preference (Most, Sim, & Inwards, 2010) (Fig

1E). OFD is frequently sporadic (not inherited) affecting one limb but has more rarely

been described in familial cases where the disease is bilateral (Fig 1F). OFD is

characterized by the development of radiolucent lesions at the periosteal surface of the

diaphyseal cortex (Fig 1C) (Gray et al., 2015) with a strong predilection for the mid-

diaphysis of the tibia. While almost exclusively occurring in this bone, the fibula, radius

and ulna may also be affected as well (Gray et al., 2015; Hitachi et al., 2018; Kahn,

2003; Park et al., 1993; Taylor et al., 2012). OFD lesions exhibit an elongated shape,

involving 10 – 40% of the bone length (Park et al., 1993).

OFD typically presents as asymptomatic lower leg swelling (Hitachi et al., 2018; Most et

al., 2010). Affected individuals suffer anterior tibial bowing prior to weight bearing and

face an increased risk of fracture. Approximately 31% of affected individuals present

with pain, 19% with pathologic fracture and 13% with tibial bowing. In addition, bone

lesions may be discovered incidentally during examination and imaging for unrelated

issues (Fig 1E) (Gleason et al., 2008; Park et al., 1993). Lesions may resolve

spontaneously during skeletal maturation (Gray et al., 2015; Most et al., 2010), but if a

fracture develops, fracture non-union often complicates the healing process. Fracture

non-union is defined as the inability to form a bridging callus at the fracture site. The

fracture non-union phenotype in OFD appear very similar to the radiological appearance

of a tibial pseudoarthrosis which complicates another single gene disorder,

Page 18: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

6

Neurofibromatosis Type 1 (NF-1) (OMIM #162200) (Ghadakzadeh, Kannu, Whetstone,

Howard, & Alman, 2016).

Histologically, OFD is characterized by a well-defined zonal architecture (Fig 1G) (Gray

et al., 2015) with a central osteolytic region, abundant in fibrous tissue and spindle-

shaped fibroblast (Gray et al., 2015; Park et al., 1993). More peripherally, cells of

osteoblastic lineage predominate (Gray et al., 2015). Additionally, OFD exhibits

immature woven bone trabeculae at the center of the lesion, which is more abundant

moving outwards (Most et al., 2010). Immature woven bone trabeculae eventually

undergoes anastomosis with one another and merges with the bone of the outer and

inner cortices (Park et al., 1993). Moving towards the outer edge of the lesions, the

trabeculae are larger and greater in number, eventually transitioning to organized

parallel layers of bone, or lamellae.

OFD presents with two fundamental histological patterns: fibrous tissue surrounded by

boney trabeculae rimmed by active osteoblasts and a zonal architecture (Park et al.,

1993). Cells found at the center of the lesions exhibit undifferentiated mesenchymal cell

markers with a subset of these cells positive for both osteoblast and epithelial cell

markers (Gray et al., 2015; Park et al., 1993). Conversely, bridging zones of the osteoid

are rich in surface osteoblasts and embedded osteocytes interspersed between lesions,

pointing to an osteoblast differentiation defect at the lesions (Gray et al., 2015).

The histological appearance of OFD is somewhat similar to that of fibrous dysplasia,

with both entities displaying variably shaped spicules of woven bone separated by

fibrous tissue (Kahn, 2003). OFD’s distinctive osteoblast rimming of the boney

trabeculae and zonal architecture distinguishes it from fibrous dysplasia, thus allowing

histological delineation between the two conditions (Fig 1G) (Gray et al., 2015; Most et

al., 2010).

Radiologically, adamantinomas (ADM) are typically considered in the differential

diagnosis of OFD (Karol, Brown, Wise, & Waldron, 2005; Park et al., 1993).

Adamantinomas are low-grade malignant tumours occurring predominantly in mature

Page 19: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

7

skeletons with a predilection for the tibial midshaft. Similar to OFD, affected individuals

report pain, anterior bowing of the tibia, fractures and fracture non-unions. Histologically,

OFD and ADM lesions exhibit similar cytokeratin immunoprofiles and cytogenetic

changes (Kahn, 2003). The commonalities in clinical presentation along with similar

histological features previously led many to hypothesize that OFD progresses into ADM,

though this has been proven to be untrue (Park et al., 1993). Unlike ADM, OFD lesions

do not undergo neoplastic transformation and are self-limiting in nature. Therefore, the

delineation between these two conditions is imperative since they vary greatly in

treatment (Most et al., 2010). Untreated or undertreated ADM results in tumour

metastasis and fatality (Kahn, 2003).

Page 20: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

8

OFD Lesions at the Tibia

Tibia Tibia

Femur Femur

Figure 1C. Osteofibrous Dysplasia lesions at the tibia

Anteroposterior (A) and lateral (B) radiographs of diaphyseal tibial osteofibrous dysplasia. Osteofibrous dysplasia is characterized by the development of fibro-osseous lesions demonstrated by the differences in radiopacity (arrows). These lesions

commonly develop at the mid-diaphysis of the tibia and in some cases affecting the fibula as well.

A B

Page 21: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

9

OFD Fractures at the Tibia

Figure 1D. Osteofibrous Dysplasia fractures at the tibia

Anteroposterior radiographs of the left and right tibias of a patient suffering from familial bilateral case of osteofibrous dysplasia. The patient presents with spontaneous non-healing tibial fractures (arrows) prior to weight bearing. Diaphyseal tibial osteofibrous

dysplasia is indicated by the radiolucent gaps.

Page 22: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

10

Adapted with Permission from Bethapudi et al., 2014

Figure 1E. Summary of Characteristics of Osteofibrous Dysplasia

Brief summary of the common characteristics and presentations of osteofibrous

dysplasia. Table describes the nature, age, sexual preference, commonly affected sites, clinical symptoms, histopathological symptoms, metastatic status and inheritance patterns of the disease.

Page 23: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

11

DNA Mutation Protein Mutation Bones Affected Bilateral or Unilateral

Family 1, Familial

c.3010_3028+8del

p.Leu964_Asp1010del

Tibia, Fibula, Ulna

Bilateral

Family 2, Familial

c.3028+1G>T

p.Leu964_Asp1010del

Tibia, Fibula Bilateral

Family 3, Familial

c.3028+1G>T

p.Leu964_Asp1010del

Tibia, Fibula Bilateral

Family 4, Simplex

c.3028+1G>T

p.Leu964_Asp1010del

Tibia, Fibula Bilateral

Sporadic c.3008A>C

p.Tyr1003Ser

Tibia, Fibula Unilateral

Figure 1F. Genetic Summary of Osteofibrous Dysplasia

Selected cases of genetic mutations resulting in osteofibrous dysplasia. All cases described here result in the loss of negat ive feedback regulatory control through either the loss of the juxtamembrane regulatory domain or the key Tyrosine residue 1003, within the juxtamembrane domain required for its function. Loss of this key tyrosine residue results in loss of CBL-mediated

degradation of the receptor post-activation and therefore elevated downstream MET signaling.

Adapted with Permission from Gray et al.,2015

Page 24: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

12

Osteofibrous Dysplasia Architecture

Figure 1G. Zonal Architecture of Osteofibrous Dysplasia Lesions Haemotoxylin & Eosin stain of osteofibrous dysplasia lesional tissue excised at surgery from a 9-

year old patient exhibiting familial osteofibrous dysplasia. The stain demonstrates spindle-shaped cells at the center of lesions (arrow) surrounded by osteoid (o). Immunohistochemical staining of lesional tissue for alkaline phosphatase (ALP) show it to be maximally expressed at the periphery

of lesions while absent at the center (arrows). Osteofibrous dysplasia lesions are characterized by rimming of osteoblasts staining positive for the osteoblastic marker, osterix (OSX), and osteoclastic markers tartrate resistant acid phosphatase (TRAP).

Reproduced with Permission from Gray et al., 2015

Page 25: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

13

1.2 Neurofibromatosis Type 1 (NF-1) NF-1 (OMIM #162200) is associated with a number of different skeletal complications

including the development of tibial bowing and non-healing tibial fractures (tibial

pseudoarthrosis). The clinical presentation of OFD thus shares many similarities with

NF-1. This autosomal dominant disorder is caused by mutations in the neurofibromin-1

(NF1) protein leading to overactivation of RAS signaling and dysregulation of β-catenin.

Similar to OFD, NF-1 is also characterized by defects in osteoblast differentiation; cells

at the NF-1 tibial fracture site do not undergo osteoblast differentiation (Ghadakzadeh et

al., 2016). The development of NF-1 related bone disease has been shown to result

from RAS pathway overactivation and upregulation of downstream ERK1/2 signaling

(Sharma et al., 2013). More recently, Ghadakzadeh et al., (2016) determined that β-

catenin levels were elevated in NF-1 pseudoarthrosis. Furthermore, localized

inactivation of the Wnt pathway with the antagonist Dickkopft-1 (Dkk1) at the fracture

site of Nf1-deficient mice resulted in improved fracture healing and osteoblast

differentiation. This study’s data is consistent with previous findings indicating β-catenin

protein levels must be tightly regulated for normal osteoblast differentiation and any

deviations result in reduced differentiation (Y. Chen et al., 2007). The similarities

between OFD and NF-1 pseudoarthrosis are uncanny and it is therefore reasonable to

hypothesize that these two single gene diseases may share similar disease

mechanisms.

Page 26: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

14

1.3 Dysregulation of MET to METΔ14 ratios causes disease: Osteofibrous Dysplasia Through exome sequencing our lab in collaboration with other groups has identified

somatic and germline mutations in the MET gene cause familial and simplex

osteofibrous dysplasia (Gray et al., 2015). Curiously, the identical germline MET exon

14 skipping mutation which causes OFD has also been reported as a somatic mutation

in lung and gastric cancer. In lung and gastric cell lines, the mutation results in an

upregulation of a MET downstream effector, mitogen-activated protein kinase (MAPK)

signaling (Heist et al., 2016). MET germline mutations different from those which cause

OFD have previously been shown to cause papillary renal cell carcinomas (OMIM

#605074). In hereditary papillary renal cell carcinoma, the MET receptor tyrosine kinase

domain undergoes autoactivating amino acid substitutions, which promotes cellular

transformation. Autosomal recessive MET mutations are also thought to be the cause of

a rare form of deafness (OMIM #616705).

All identified OFD somatic and germline mutations result in exon 14 skipping (Gray et

al., 2015). As previously mentioned MET’s exon 14 encodes for the regulatory

juxtamembrane domain of the MET receptor where the ubiquitin ligase CBL binds (Fig

1H). One of the identified OFD mutations results in alternative splicing of MET and

exclusion of the juxtamembrane domain pY1003 ubiquitination target. Loss of the CBL

pY1003 docking site leads to reduced internalization and MET receptor degradation.

The receptor’s signaling function remains competent in downstream signal transduction

resulting in aberrant MET signaling (Gray et al., 2015). These results suggest the

mutations underlying OFD stabilize the MET receptor and confer a gain-of-function to

the protein.

1.4 Characterization of Mesenchymal-Epithelial Transition Gene’s Structure The mesenchymal-epithelial transition gene (MET) is a proto-oncogene found on

chromosome 7 band 7q21–q31. Its expression is regulated by Ets (E-twenty six), Pax3

(paired box 3), AP2 (activator protein-2) and Tcf-4 (transcription factor 4) (Boon et al.,

2002). MET is synthesized as a 170 kDa precursor which undergoes post-translational

modification into a mature 190 kDa transmembrane receptor (Parikh et al., 2018). The

protein product of this gene, MET (also called c-MET) tyrosine kinase, is actively

Page 27: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

15

expressed during embryogenesis and adulthood on epithelial cells of many organs such

as the liver, blood vessels, muscle, and bone marrow. MET receptor expression can

also be found on cells involved in bone homeostasis such as osteoblasts, osteoclasts,

fibroblasts, and chondrocytes (Gray et al., 2015; Guévremont et al., 2003; Organ &

Tsao, 2011).

The MET receptor is composed of an and -subunit with the subunit composed

solely of extracellular segments while the -subunit encompasses intracellular and

extracellular domains. The extracellular domain of the MET receptor comprises an N

terminal domain, plexin-semaphorin-integrin (PSI) domain, and transmembrane domain.

The N-terminal domain is composed of a 500-residue semaphorin domain, sharing

homologies with domains found in the semaphorin and plexin families. This region

encompasses the extracellular subunit and a portion of the transmembrane -

subunits, linked together by disulfide bonds. The PSI domain follows the semaphorin

domain, spanning approximately 50 residues with four disulfide bonds (Organ et al.,

2011). The PSI domain connects the extracellular region of the receptor to the

transmembrane helix via four immunoglobulin-plexin-transcription (IPT) domains (Organ

et al., 2011).

Intracellularly, the MET receptor is composed of a tyrosine kinase catalytic domain

flanked by a distinctive regulatory juxtamembrane domain and carboxy-terminal

sequences essential for substrate docking and downstream signalling (Liu et al., 2008;

Organ et al., 2011). There are numerous key tyrosine residues along the intracellular

domain of the receptor, critical for normal cellular function and regulation of receptor

signaling. Of note, catalytic tyrosine residues, found within the catalytic domain, Tyr1234

and Tyr1235 positively enhance downstream enzyme activity (Organ & Tsao, 2011).

Negative regulation of the MET-receptor occurs at Tyr1003 of the juxtamembrane

domain, in a negative feedback fashion with the recruitment of ubiquitin ligase casitas B-

lineage lymphoma (CBL) (Gray et al., 2015; Organ et al., 2011) (Fig 1B).

MET is a prototypic member of the receptor tyrosine kinase (RTK) family but

differentiates itself from the rest of the family through its distinct structure. It is also the

Page 28: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

16

only high-affinity receptor for hepatocyte growth factor (HGF) also known as scatter

factor (SF) (Christensen et al., 2005). In a similar fashion to the MET receptor, HGF is

synthesized as an inactive precursor and is converted to a two-chain, disulfide-linked

heterodimer through proteolytic cleavage (X. Liu et al., 2008). HGF is sequestered in its

active form in the extracellular matrices of most tissues by heparin-like proteoglycans

(T. Kobayashi et al., 1994). Cells of mesenchymal origin, such as osteoblasts,

osteoclasts, and fibroblasts are the main source of HGF production, with osteoblasts

being the preeminent source of HGF in bones (Grano et al., 2002; Vallet et al., 2016).

RTKs have been shown to be intricately involved in key processes of mammalian

development, cell function and tissue homeostasis. Dysregulation of RTKs affect crucial

processes such as cell growth and survival, organ morphogenesis, neovascularization

and tissue repair and regeneration (Christensen et al., 2005).

1.5 MET Signaling and Function Stimulation of the MET receptor by HGF, results in receptor dimerization or

multimerization and phosphorylation of tyrosine residues along the juxtamembrane,

catalytic and cytoplasmic tail domains allowing for regulation of internalization, catalytic

activity, and docking of regulatory substrates respectively (X. Liu et al., 2008; Naldini et

al., 1991). Some of these phosphorylation events result in activation of cellular growth

programs, promoting mitogenesis, motility, invasion and morphogenesis (Frampton et

al., 2015). Upon activation, phosphorylation of tyrosine residues found in the catalytic

loop of the kinase domain, Tyr1234 and Tyr1235, ensues (Birchmeier & Gherardi,

1998). Phosphorylation at these sites triggers a conformational change in the receptor

resulting in phosphorylation events at Tyr1349 and Tyr1356 of the C-terminal domain

allowing them to act as docking sites for adaptor and effector proteins with SH2-

containing domains, phosphotyrosine binding (PTB) domains, and MET binding

domains (MBDs) (Sattler et al., 2009). These docking site are crucial for activating

downstream pathways such as RAS/mitogen-activated protein kinase (RAS/MAPK),

phosphatidylinositol 3-kinase/protein kinase B (PI3K/AKT), signal transducer and

activator of transcription (STATs), phospholipase C (PLC), and proto-oncogene

tyrosine kinase Src (c-Src) (Furge et al., 2000; Ponzetto et al., 1994).

Page 29: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

17

The tyrosine kinase domain of the MET receptor is also responsible for the initiation of

various downstream signalling pathways, chiefly through the following major pathway

regulators: MAPK, PI3K/AKT, and STAT (Parikh et al., 2018). Under normal

physiological conditions these pathways regulate cell proliferation, motility, and survival

among other numerous cellular processes (Parikh et al., 2018). MET activation also has

a crucial role in epithelial-mesenchymal interactions during injury repair (Lesko, 2007).

MET mediates epithelial cell dissociation, migration towards the site of injury,

proliferation and reconstruction of the epidermal layer (Lesko, 2007). Additional post-

translational and receptor domain modifications further contribute to the regulation of

biological functions downstream of the activated MET receptor (Sattler et al., 2009).

1.6 MET’s Role in Development Interactions between HGF and its receptor have been previously shown to play an

essential role in mammalian embryogenesis, muscle development, nervous system

formation, hematopoietic cell differentiation, bone remodeling, angiogenesis and

organization of three-dimensional tubular structures (e.g. renal tubular cells) (Birchmeier

et al., 1998; Comoglio et al., 2001). In embryogenesis, HGF mediated MET signaling

produces instructions critical for growth and survival of hepatocytes and trophoblast

cells. Hgf and Met knockout embryos display underdeveloped liver and placental

labyrinths caused by defects in epithelial-mesenchymal transition (Schmidt et al., 1995).

This results in early embryonic lethality due to compromised placental mediated

exchange between maternal and fetal blood (Schmidt et al., 1995; Trusolino, Bertotti, &

Comoglio, 2010). These mice also exhibit reduced hypaxial muscles, muscles ventral to

the horizontal septum of the vertebrae, due to loss of MET mediated proliferation and

motility (Lesko, 2007). Lastly, loss of MET signalling’s anti-apoptotic effects leads to

increased apoptosis in neuronal cells leading to impaired axon bundling and neuronal

defects (Trusolino et al., 2010).

1.7 CBL-Dependent Regulation of MET Signaling In addition to downstream pathway activation and signalling post-ligand binding, the

HGF/MET receptor complex is quickly internalized post-phosphorylation of Tyr1003 in

the juxtamembrane domain of the receptor into clathrin-coated vesicles (Abella et al.,

Page 30: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

18

2005). This phosphorylation modification recruits the Casitas B-lineage Lymphoma

(CBL) protein, an E3 ubiquitin-protein ligase, which facilitates polyubiquitination of the

receptor complex creating ubiquitin recognition motifs for trafficking and degradation of

the internalized receptor (Baldanzi et al., 2014; Sattler et al., 2009). CBL also acts as an

adaptor for endophilin, a protein crucial for receptor internalization, trafficking and

degradation (Kjaerulff et al., 2011). Internalized receptors continue on to two distinct

fates: they are recycled back to the plasma membrane or degraded via the lysosomal

pathway (Teis et al., 2003). This negative feedback mechanism allows for modulation of

MET signaling and associated downstream signalling. Disruption of this mechanism

could result in cellular transformation through increased recycling of the receptor back to

the plasma membrane causing hyper-signalling or increased degradation of the

internalized receptor leading to hypo-signalling (Abella et al., 2005).

1.8 Other Means of Regulating MET Signaling Ubiquitin mediated lysosomal degradation of the MET receptor is the major determinant

of receptor sensitivity and regulator of signal activity, though it is not the only regulatory

mechanism (Trusolino et al., 2010). Numerous cytokines such as interleukin-1, -6 and

transforming growth factor β (TGFβ) have been shown to enhance transcriptional

activity of HGF in fibroblasts and macrophages and MET in epithelial cells. These

factors also upregulate the production of plasmin and matriptase proteases, to enhance

cleavage of pro-HGF to its active form (Bhowmick, Neilson, & Moses, 2004;

Michalopoulos & DeFrances, 2005). Proteasomal degradation can also occur through

CBL-interacting protein 85 (CiN85 or SH3KbP1) endophilins, which acts as a scaffold

molecule for CBL to promote internalization of the MET receptor through contact

dependent means (Oved & Yarden, 2002). Lastly, the MET receptor can be regulated

through proteolytic means as well. A disintegrin and metalloprotease (AdAM) mediated

shedding of the extracellular domain can also regulate receptor stability, releasing a

soluble N-terminal fragment with a cytoplasmic tail remaining anchored to the

membrane. The remaining cytoplasmic tail is proteolytically cleaved by γ-secretase and

degraded by proteasomal means (Foveau et al., 2009). Proteolytic degradation does not

require activation of the receptor like CBL-mediated degradation does. In addition, it

leaves behind the extracellular domain of the receptor, allowing for sequestration of

Page 31: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

19

HGF and impairs dimerization with full-length MET receptors causing inhibition of MET

signaling (Trusolino et al., 2010; Zhang, Graveel, Shinomiya, & Vande Woude, 2004).

1.9 Mouse Model of Osteofibrous Dysplasia The regulatory juxtamembrane domain found in the human MET receptor is coded by

exon 15 in mice (exon 14 in humans). In order to study OFD, we created a MetΔ15-HET

mouse model (through the CRISPR-Cas9 system), with an in-frame deletion of the

juxtamembrane segment containing Y1003 to replicate the mutation seen in human

disease (Fig 1B). Mice homozygous for the MetΔ15 mutation are embryonically lethal at

around embryonic day 10.5 (E10.5) and are not the scope of this project. Heterozygous

MetΔ15 mice are used in this study. MetΔ15-HET mice are viable and fertile and do not

develop liver abnormalities or tumors. A semi-stabilized tibial fracture model is used to

induce representative mid-diaphyseal fractures seen in humans since Met Δ15-HET mice

do not develop spontaneous fractures.

Page 32: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

20

Figure 1H. MET Exon 14 Skipping Mutation Schematic diagram of MET showing a point mutation of G>T at the exon-intron boundary of Exon

14 resulting in the skipping of Exon 14. Exon 14 codes for a regulatory juxtamembrane domain crucial in negative feedback control of MET signaling post-activation. This mutation was found in family 2, 3 and 4 as described in Figure 1F.

Page 33: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

21

1.10 Chondrocyte Differentiation

Chondrogenesis is a critical process in the development and maintenance of healthy

skeletal physiology in vertebrates during embryonic and postnatal development

(Kozhemyakina, Lassar, & Zelzer, 2015). Though chondrocytes also play a critical role

in bone physiology and endochondral ossification, this study is focused on the effects of

aberrant MET signaling on osteoblasts and therefore the review of chondrogenesis is

more rudimentary in nature.

Chondrogenesis involves the creation of chondrocytes through progressive steps in the

commitment and differentiation of mesenchymal stem cells (MSCs) to produce new

cartilage tissue. MSCs are unique in that they are pluripotent, meaning they have the

ability to differentiate into a limited variety of cell types, such as adipocytes, osteoblast,

chondrocytes and endothelial cells (Pez Ponte et al., 2007). Chondrogenesis starts with

the migration and condensation of mesenchymal stem cells (mediated by cell-cell

interactions) at prospective skeleton sites (Oberlender & Tuan, 1994; Sun & Beier,

2014). Members of the SOX transcription factor family commit MSCs to immature

chondrocytes after cellular condensation has occurred. This is followed by proliferation

and hypertrophy of the cells. At the end of chondrogenesis, terminal chondrocytes can

either undergo apoptosis or transdifferentiate to osteoblasts. This transdifferentiation

process provides an additional source of osteoblasts distinct from the periosteum

derived pool during endochondral ossification (Yang, Tsang, Tang, Chan, & Cheah,

2016; Zhou et al., 2014).

Page 34: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

22

1.11 Osteoblast Differentiation MSCs have been shown to be heavily involved in the fracture healing process and

regeneration of bone post-injury (Garg et al., 2017). Under the influence of injury-

mediated chemokines, MSCs found in bone marrow, periosteum, walls of blood vessels

and circulation, and muscle are recruited to the fracture site. The divergent

differentiation MSCs undergo is strongly influenced and regulated by several key growth

factors though the full understanding of the mechanism behind this recruitment remains

unclear (Ito, 2011). (Kratchmarova et al., 2005).

Osteoblast are cells of mesenchymal origin which play a critical role in maintaining bone

homeostasis. Their involvement can be broken down into two main categories involving

the formation and resorption of bone. In their formative role, osteoblasts produce

extracellular matrix proteins, regulate matrix mineralization and control bone remodeling

(Corrado et al., 2017). Osteoblasts fulfil their role in bone resorption through expression

of RANKL and OPG. RANKL is an essential factor for the recruitment, differentiation,

activation and survival of osteoclastic cells through its interactions with the RANK

receptor found on immature and mature osteoclastic cells (Langdahl, Ferrari, &

Dempster, 2016). OPG is a soluble receptor for RANKL, which inhibits the promotion of

osteoclastic differentiation and activity, by sequestering and preventing RANK-RANKL

interactions (Langdahl et al., 2016; Wei Liu & Zhang, 2015).

Osteoblast commitment, differentiation and growth is controlled by several local and

systemic pathways which act through paracrine and/or autocrine manners in order to

regulate the activity of key transcriptional factors (Aubin et al., 1995). Some of these

factors include bone morphogenetic proteins (BMPs), hedgehog proteins, cell growth

factors, hormones, cytokine modulators and canonical Wnt/β-catenin signaling (Vimalraj

et al., 2015). The effect of these factors in promoting or inhibiting osteoblast

differentiation relies on the stage of maturation and development (Canalis et al., 1992).

This complex differentiation process can largely defined by three distinct stages:

proliferation, maturation and extra-cellular matrix synthesis, and matrix mineralization

(Neve, Corrado, & Cantatore, 2011) (Fig 1I).

Page 35: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

23

Proliferation:

Mesenchymal stem cells become committed to the osteoblastic differentiation pathway

under Runx2 expression. RUNX2 directs mesenchymal progenitor cells to begin

differentiation to pre-osteoblasts, while inhibiting their differentiation to adipocytes and

chondrocytes (Toshihisa Komori, 2006). Proliferating osteoblasts begin expressing

fibronectin, TGFβ Receptor 1 and osteopontin (Rutkovskiy, Stensløkken, & Vaage,

2016), markers of early stage osteoblast differentiation.

Maturation and Extracellular Matrix Synthesis:

As pre-osteoblasts mature, they exit the cell cycle and differentiate to immature

osteoblasts which produce extracellular matrix proteins such as alkaline phosphatase

and type 1 collagen (Col1a1). The immature osteoblast stage is characterized by

maximal expression of alkaline phosphatase (ALP) a membrane-bound enzyme

required for bone matrix mineralization (Huang, 2007). The expression of ALP occurs

post type 1 collagen production and reaches maximal levels just prior to mineralization.

ALP can also be found in other tissues such as the liver, intestines, spleen and kidneys.

In bone, it is a marker of late-stage osteoblast differentiation (Heino & Hentunen, 2008).

Matrix Mineralization:

Mature osteoblast are cuboidal in shape, with enlarged Golgi complexes and a well-

defined endoplasmic reticulum (Heino & Hentunen, 2008). These cells begin expression

and secretion of the following matrix markers: osteopontin, bone sialoprotein and

osteocalcin (Huang, 2007).

1.11 Osteocyte Differentiation Osteoblasts which have embedded themselves within the bone matrix differentiate to

osteocytes. Osteocytes are spider-shaped cells derived from MSCs which have

undergone the osteoblast differentiation process. During the transition from osteoblast

to osteocyte, the once prominent Golgi complexes and endoplasmic reticulum decrease

in size (Capulli, Paone, & Rucci, 2014). Osteoblastic markers such as BSP, OC, ALP

and type 1 collagen which were once abundantly expressed are downregulated and

replaced with osteocytic markers (Bellido, 2014). Osteocytes are the most abundant

Page 36: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

24

cells found in bone, making up 90 – 95% of the total number of cells. Osteocytes

possess branched processes which extend throughout the mineralized matrix, thus

allowing them to form an interconnected network. This allows them to sense and

respond to stimuli to regulate bone remodelling and adaptation. These effects are

mediated through both cell-cell interactions and signaling with osteoclasts and

osteoblasts (Goldring, 2015).

1.12 Osteoclast Differentiation Osteoclasts are multinucleated bone reabsorbing cells differentiated from hematopoietic

stem cells found in the bone marrow. Osteoclasts rely on the presence of osteoblast

and osteocyte secreted macrophage-colony stimulating factor (M-CSF) and RANKL for

their survival, proliferation, differentiation and activation (Park-Min, 2018). In the

presence of M-CSF, expression of the RANK receptor is increased in osteoclast

precursor cells allowing for increased sensitivity to RANKL activation (Katagiri &

Takahashi, 2002). RANKL-RANK signaling promotes the differentiation and fusion of

preosteoclasts to multinucleated mature osteoclasts. Osteoclasts exert their resorptive

function through attachment to the bone matrix and forming an integrin mediated seal

around the resorption zone, to separate the resorptive microenvironment from the

extracellular space (Soysa & Alles, 2016). This is followed by the targeted secretion of

HCl to dissolve the hydroxyapatite crystals allowing for secreted proteolytic enzymes to

degrade the collagenous bone matrix. Degraded products are removed via a

transcytosis pathway from the ruffled border of the sealing zone to the secretory

domain, where it is released into the extracellular matrix (Väänänen, Zhao, Mulari, &

Halleen, 2000).

Page 37: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

25

Figure 1I. Stages of Osteoblast Differentiation Brief overview of the stages of osteoblast differentiation and select markers secreted at the indicated stages. Progenitor cells are

committed to the osteo-progenitor fate by RUNX2 expression. Osteo-progenitor cells transition to immature osteoblast under the regulation of RUNX2 and OSX. Immature osteoblast begin secretion of extracellular proteins such as alkaline phosphatase, bone sialoprotein and type 1 collagen. These factors are critical for matrix mineralization. Mature osteoblast express and secret Osteonectin, osteopontin, and

osteocalcin during the mineralization stage. As mature osteoblast become embedded within the bone matrix, they differentiate to

osteocytes.

Page 38: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

26

1.13 Selected Molecular Factors Affecting Osteoblast Differentiation

RUNX2 is a key transcription factor in bone formation

The RUNX family of transcription factors are defined by a runt DNA binding domain.

RUNX2 functions as a platform protein to regulate bone specific genes, coregulatory

proteins, and chromatin remodelling factors. RUNX2 has the ability to enhance DNA

binding of itself, through upregulation of PI3K and AKT expression (Thomas et al.,

2004). RUNX2 transcriptional activity can also be regulated through phosphorylation

events, of note, ERK1/2 mediated phosphorylation of RUNX2 enhances its association

with coactivators to upregulate osteoblast specific genes (Greenblatt et al., 2010).

Mice deficient in Runx2 exhibit a complete lack of bone (T. Komori et al., 1997; Otto et

al., 1997). In contrast, Runx2 over-expression induces non-osteogenic cells to express

osteoblastic markers (Yamaguchi et al., 2000). In addition to its role in bone formation,

RUNX2 regulates RANKL expression, an inducer of osteoclastogenesis (Vimalraj et al.,

2015). RUNX2 can also form complexes with STAT1, TWIST and HEY1 to disrupt

binding of key downstream osteogenic gene promoter regions, resulting in inhibition of

osteoblast differentiation (Bialek et al., 2004; S. Kim et al., 2003; Thomas et al., 2004).

Osterix is important in the differentiation of pre-osteoblasts to immature

osteoblasts

The commitment of pre-osteoblasts to immature osteoblasts is partly regulated by

osterix (OSX). Osterix null mice are completely void of osteoblasts, indicating the

essential role of this transcription factor in osteoblast differentiation (Huang, 2007; Koga

et al., 2005). Osterix’s role in osteoblast differentiation is downstream of RUNX2; Runx2

expression is present in the mesenchymal stem cells of osterix null mice, but osterix

expression is absent in Runx2 null mice (Huang, 2007; Toshihisa Komori, 2006).

1.14 Key Signaling Pathways Involved in Bone WNT Pathway

The canonical WNT pathway is important in the formation of bone and a tight regulation

of its activity is required to ensure the terminal differentiation of osteoblasts. Canonical

Page 39: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

27

WNT signaling involves the stabilization of β-catenin by inhibiting glycogen synthase

kinase 3 beta (GSK-3β), a critical component of the β-catenin degradation complex

(Nelson & Nusse, 2004; Stamos & Weis, 2013). GSK-3β is also inhibited by

adrenomedullin (ADM); ADM is produced in bone tissues and acts as an important

regulator of osteoblast cells through modulation of Wnt signaling (Lausson & Cressent,

2011). Inhibition of GSK-3β mediated β-catenin phosphorylation results in the

accumulation of unphosphorylated β-catenin in the cytoplasm, translocation of β-catenin

to the cytosol and activation of downstream gene targets such as Axin2. Axin2 is a

scaffolding protein which promotes the phosphorylation of β-catenin by GSK-3β and its

subsequent degradation (Jho et al., 2002).

β-catenin is essential for osteoblast differentiation in both endochondral and

intramembranous ossification. While the inactivation of β-catenin blocks osteoblast

differentiation and results in a chondrocyte fate(Day et al., 2005), β-catenin appears not

to be crucial in the initial commitment stages of osteoblast differentiat ion. β-catenin

regulates the pre-osteoblast to immature osteoblast transition stages (Hill et al., 2005;

Hu, 2004) by enhancing Runx2 expression, thus promoting commitment to the

osteoblastic lineage and inhibiting the adipocyte and chondrocyte fates (Gaur et al.,

2005).

RAS-MEK1/2-ERK1/2 Signaling

RAS proteins act as molecular switches for many signaling cascades involved in cell

biology. RAS activation occurs downstream of MET receptor signaling and has been

previously shown to upregulate osteoprogenitor cell proliferation resulting in increased

osteoblastic and stromal cell descendants (Papaioannou, Mirzamohammadi, &

Kobayashi, 2016). RAS activation and associated activation of PI3K, regulates AKT and

ERK1/2 signaling activity to stimulate osteoblast differentiation (Ghosh-Choudhury,

Mandal, & Choudhury, 2007; Yamashita et al., 2008). RAS protein activator like 3

(RASAL3) is a part of the Ras protein family and a negative regulator of RAS signaling.

RASAL3 inactivates RAS proteins by converting active RAS-GTP to inactive RAS-GDP.

Knockdown of RASAL3 increases RAS-GTP levels and phosphorylated ERK, a marker

of RAS activity (Muro et al., 2015).

Page 40: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

28

Downstream effectors of RAS signaling such as mitogen-activated protein kinase kinase

(MEK1/2 or MAP2K) and mitogen-activated protein kinase 1 (ERK1/2 or MAPK1),

regulate Runx2 expression through modulation of canonical Wnt signaling (Li, Ge, &

Franceschi, 2017; R.T., C., G., H., & D., 2009). RAS signaling can also increase

stabilization and modulation of β-catenin directly (Jeong, Ro, & Choi, 2018). Through

these studies, it is clear that the regulation of RAS and its downstream effectors is

crucial for the osteoblast differentiation process.

PI3K-AKT-mTOR Signaling

Activation of phosphoinositide 3-kinase (PI3K) occurs directly downstream of the MET

receptor, its activation creates a docking site for protein kinase B (AKT) thus allowing

access for PI3K to activate AKT through phosphorylation. AKT directly phosphorylates

and activates mammalian target of rapamycin (mTOR), in the mammalian target of

rapamycin complex 1 (mTORC1) (Hemmings & Restuccia, 2012). PI3K-AKT-mTOR

kinase pathway activation directly promotes osteoblast differentiation and proliferation

through. There is also a connection with β-catenin since activated AKT signalling

phosphorylates β-catenin at Serine 552 to enhance its stabilization and nuclear

translocation, where it upregulates the expression of Runx2 and Osterix (Raucci,

Bellosta, Grassi, Basilico, & Mansukhani, 2008). AKT activation of mTOR also

upregulates Runx2 expression(Dai et al., 2017). Thus, it is clear that regulation of PI3K

signaling and its downstream effectors is crucial for osteoblast differentiation.

Page 41: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

29

Page 42: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

30

Figure 1J. Summary of Genes and Factors Affecting Osteoblast Differentiation

Effects of selected published genes and factors affecting osteoblast differentiation and the model utilized to determine their effect.

Page 43: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

31

1.15 Bone Development

The adult skeleton consists of 206 bones subcategorized into the axial skeleton (skull,

ribs and spine) and appendicular skeleton. Though morphologically and functionally

diverse, all bones form through either intramembranous and endochondral ossification.

Intramembranous ossification (IO) forms the flat bones which make up the skull, the

mandible and the clavicles. IO is characterized by the direct differentiation of

mesenchymal stem cells into osteoblasts, eventually producing bone (Fig 1K) (E. M.

Thompson, Matsiko, Farrell, Kelly, & O’Brien, 2015). Appendicular bones such as the

tibia, the preeminent site affected by OFD, are formed through endochondral

ossification (EO) and therefore will be the main focus of this section. EO relies on the

formation of an intermediate cartilaginous template which is replaced by bone through a

series of processes (Shapiro, 2008) (Fig 1K). EO begins with the condensation and

differentiation of mesenchymal stem cells to chondrocytes, which secrete proteoglycans

and extracellular matrix forming the hyaline cartilaginous template (E. M. Thompson et

al., 2015). Once the cartilaginous structure is completed, chondrocytes stop secretion

and while embedded in the cartilage template, begin to proliferate (Farrell et al., 2011).

Chondrocytes then align into lacunae columns, mature and undergo hypertrophy

followed by apoptosis or transdifferentiation to osteoblast (Long & Ornitz, 2013; Yang et

al., 2016). This is followed by vascularization and calcification of the transitory cartilage

template (Long & Ornitz, 2013). While the cartilage template continues to grow

interstitially, mesenchymal stem cells of the perichondrium begin differentiation to

osteoblasts, secreting osteoid to make a cortical bone collar, transforming the

perichondrium into the periosteum (E. M. Thompson et al., 2015). This cortical bone

acts as a dense protective outer layer of bone surrounding the inner cavity (E. M.

Thompson et al., 2015). Osteoprogenitor cells occupying the cavity at the center of the

cartilage template formed by chondroclasts will now differentiate to osteoblasts and form

a primary ossification center. These osteoblasts produce and secrete osteoid forming

spongy trabecular bone.

Page 44: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

32

Figure 1K. Overview of Intramembranous and Endochondral Ossification Intramembranous ossification occurs with the direct differentiation of mesenchymal stem cells to osteoblasts and subsequent f ormation of an ossification center and bone formation (A). Endochondral ossification begins with the condensation and differentiation of mesenchymal stem cells to chondrocytes which forms a cartilaginous template resembling the final bone’s structure. In the final stages of endochondral

ossification osteoblasts mineralize this cartilaginous template resulting in the formation of new bone (B).

Page 45: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

33

1.16 Fracture Healing

A fracture is defined by a lack of continuity in the bone. Fracture healing under the right

circumstances is a robust process. However, approximately 4 to 10% of all fractures

develop in a non-union fashion (Hak et al., 2014). A fracture non-union is the incomplete

consolidation of a fracture, without the formation of a boney and cartilaginous bridge

across the disconnected ends of bone also known as a callus, after approximately 4 – 8

months post-injury (Tall, 2018).

Fracture non-unions present in two distinct forms; hypertrophic and atrophic non-unions

(Kostenuik & Mirza, 2017). Hypertrophic non-unions are caused by an unstable fracture

site due to excessive fracture site motion resulting in large amounts of non-bridging

callus at the fracture site (Hak, 2011; Tall, 2018). Atrophic non-unions develop due to a

deficiency in vascularization post-fracture caused by injury to the soft-tissue envelope

depriving fractures of normal blood supply and reduced vascular in-growth resulting in

fibrous tissue at the fracture site (Hak, 2011; Hankenson, Dishowitz, Gray, & Schenker,

2011; Tall, 2018).

Fracture non-unions result in poor mobility compromising ones activities of daily living,

resulting in a loss of productivity associated mental health burdens (Hak et al., 2014).

While the management of a fracture non-union is largely surgical there have been

promising results utilizing recombinant human bone morphogenetic proteins (BMPs) to

enhance bone formation and improve bridging of fracture callus (Kostenuik & Mirza,

2017). There remains significant challenges associated with sequestering BMPs at the

fracture site and an opportunity to develop new treatment modalities (Marsell & Einhorn,

2011).

Primary, or direct healing is extremely rare and only typically occurs with rigid support of

the fracture and formation of cortical bone without a transitory cartilaginous intermediate

(Z. Thompson, Miclau, Hu, & Helms, 2002). Secondary, or indirect healing is much more

common as it does not rely on stabilization and rigidity of the fracture site (Marsell &

Einhorn, 2011). Fracture healing of bones in the appendicular skeleton, such as the

Page 46: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

34

tibia, heal through secondary healing; a combination of intramembranous and

endochondral ossification (Runyan & Gabrick, 2017).

Secondary fracture healing can be described in the 3 following phases (Fig 1L):

1) Reactive Phase: Inflammation occurs directly after a fracture breaks the blood vessels

and disrupts the blood supply to the bone marrow, periosteum, bone and surrounding

soft tissue (Jha, Blau, & Bhattacharyya, 2016). First, a hematoma forms secondary to

the activation of the plasma coagulation cascade and exposure of platelets to the

extravascular environment (Loi et al., 2016). These events in addition to the local

microenvironment, trigger the condensation of mesenchymal cells from multiple sources

(periosteum, endosteum, and bone marrow) and their differentiation to chondrocytes

and osteoblasts (Loi et al., 2016). Since muscle and periosteum overlying the bone are

sources of mesenchymal stem cells, fractures with poor soft tissue coverage may result

in delayed or inhibited healing (Chan, Harry, Williams, & Nanchahal, 2012).

2) Reparative Phase: The hematoma, acts as a scaffold for callus formation, in which the

recently recruited mesenchymal stem cells differentiate to chondrocytes and osteoblasts

(Jha et al., 2016). Chondrocytes begin laying down cartilage, forming a soft callus which

is crucial in bridging the gap and restoring some of the bone’s original strength and

temporary stability. In the last stages of soft callus formation, chondrocytes mineralize

the soft matrix. IO occurs at regions of healthy vasculature, directly adjacent to the distal

and proximal ends of the fracture, resulting in formation of a hard callus (Marsell &

Einhorn, 2011). In parallel, EO occurs between the ends of the fracture and external to

periosteal sites (Marsell & Einhorn, 2011).

3) Remodeling Phase: This phase is characterized by osteoblast and osteoclast

cooperation and communication mediated through RANK-RANKL interactions and

osteoprotegerin (OPG) expression. The hard callus is gradually remodeled by

osteoclasts and osteoblasts until normal bone geometry and integrity is re-established

(Kostenuik & Mirza, 2017).

Page 47: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

35

Figure 1L. Overview of Stages of Fracture Repair

Brief overview of the stages of fracture repair. There are four stages in the repair of a broken bone: (A) Formation of a hematoma post injury. (B) Formation of a fibrocartilaginous soft callus. (C) The formation of a bony hard callus and (D) Remodelling and addition of

compact bone.

Page 48: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

36

Figure 1M. Genes Involved in Fracture Repair Schematic diagram of the cells important in endochondral ossification mediated fracture repair; chondrocytes, osteoblasts and osteoclast.

Osteoblasts and osteoclasts have the ability to express MET and its one and only ligand, HGF. Furthermore, previous studies have implicated downstream effectors of MET, RAS, MAPK, and AKT to be involved in the regulation of key transcription factors involved in the differentiation process of these cells.

Page 49: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

37

1.17 What is known about the MET-HGF pathway in Bone MET and its only ligand HGF are expressed in osteoblasts and osteoclasts. HGF may

be a coupling factor between osteoblasts and osteoclasts (Fig 1M) (Grano et al., 2002;

Lee et al., 2018). The activation of the MET receptor in osteoblasts results in cell cycle

progression. In vitro studies show loss of MET in osteoblast resulted in suppressed

osteoclastogenesis (Lee et al., 2018). Furthermore, exogenous HGF stimulation of

human mesenchymal stem cells inhibited bone-morphogenic protein-2 (BMP-2) and

osteopontin, and other markers of osteoblast differentiation, via the MET receptor and

its downstream effectors (Standal et al., 2007). Others have reported HGF stimulation of

human mesenchymal stem cells promotes osteogenic marker expression and the

inhibition of it, results in reduced matrix mineralization (Aenlle, Curtis, Roos, & Howard,

2014; H. Te Chen, Tsou, Chang, & Tang, 2012; Tsai, Huang, Yang, & Tang, 2012).

While these data imply a regulatory role for MET signaling and HGF in the osteoblast

differentiation process, the explanation for the contradictory results remains unclear.

Page 50: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

38

2 Research Aims, Hypothesis, and Summary Plan

MET signaling regulates a diverse number of functions including but not limited to

proliferation and cell survival. MET protein levels have been shown to be elevated in the

fracture non-union tissue of OFD patients (Gray et al., 2015). Previously published data

also indicate abnormalities of osteoblast differentiation in OFD patient cells. MET has

previously been shown to be expressed at the periosteum of long bones, a crucial

source of mesenchymal stem cells during fracture repair (Gray et al., 2015; Mara et al.,

2011). Its’ expression can be found in osteoblasts, osteoclasts, chondrocytes and

fibroblasts, which all play their own crucial role in fracture repair (Fig 1M) (Grano et al.,

2002; Vallet et al., 2016). Our group and others have identified three different germline

MET mutations having a causative role in OFD; these novel autosomal dominant gain-

of-function mutations all result in the loss of MET’s juxtamembrane domain, which is

crucial for downregulation of MET signaling. The cause of the reduced fracture healing

seen in OFD patients is still largely unknown therefore investigating the MET receptor

and osteoblast behaviour can provide insight into new mechanisms by which long bones

grow and heal. Taken together, these findings suggest MET signaling may have a

critical role in normal bone biology and fracture repair. We hypothesize that gain-of-

function MET mutations result in delayed bone repair ability due to reduced

osteoblast differentiation. Since the cause of the reduced fracture healing seen in

OFD patients is still largely unknown this study will focus on the role of MET in

osteoblasts and during fracture repair.

The following experiments will be completed to test the hypothesis. Pathway analysis at

the protein level will be conducted to visualize any dysregulations in key signaling

pathways downstream of MET such as RAS, AKT, and β-catenin which have been

previously shown to regulate osteoblast differentiation (Y. Chen et al., 2007;

Papaioannou et al., 2016). This is followed by examination of osteoblast differentiation

in vitro through colony forming unit (CFU) assays to determine osteogenic potential of

WT and Met Δ15-HET bone marrow stromal cells (BMSCs). The length of the long bones

and epiphyseal growth plates of WT and MetΔ15-HET mice will then be examined to

expose any gross skeletal phenotypes present in mutants. We will shift our focus to

Page 51: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

39

examining fracture healing in vivo. We will utilize a semi-stabilized tibial fracture model

to induce fractures since our genetic mouse model does not develop spontaneous

fractures. Fractures generated through this method have been shown to heal through

intramembranous and endochondral ossification (Y. Chen et al., 2007).

2.1 Rationale We have identified a unique human condition in which gain-of-function MET mutations

result in non-healing fractures with increased MET expression at the bone fracture site.

In mice MET is expressed at the periosteum, which is a thin membranous structure

involved in the maintenance of cortical bone and fracture healing. Surprisingly, little else

is known about MET in fracture repair or its role in bone cell differentiation. Given that

many of the downstream pathways of MET have been previously implicated in delayed

fracture healing and osteoblast differentiation, dysregulation of MET signaling may be

responsible for the delays in fracture healing seen in OFD patients (Y. Chen et al., 2007;

Gray et al., 2015). In addition, the inhibition of the MET receptor has previously been

shown to stimulate osteoblast differentiation and bone regeneration in vitro by multiple

groups (Fioramonti et al., 2017; J. W. Kim et al., 2017; Kokabu, 2013; Shibasaki et al.,

2015). However, the effects of aberrant and upregulated MET signaling on osteoblast

differentiation and bone repair is still unclear.

2.2 Hypothesis Previous research in the Kannu lab has implicated MET exon 14 skipping mutations to

cause osteofibrous dysplasia in humans. Forced induction of this exon-exclusion event

results in retarded osteoblast differentiation and reduced bone-matrix mineralization in

vitro but the mechanism underlying this phenotype is still largely unknown (Gray et al.,

2015). We hypothesize that gain-of-function MET mutations result in delayed bone

repair ability due to reduced osteoblast differentiation.

Page 52: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

40

2.3 Objectives

Objective 1: Investigating MET pathway irregularities in Met Δ15-HET mice

The first aim of this study was to investigate whether there are similar MET pathway

irregularities in Met Δ15-HET mice as what is seen in human OFD. The utilization of a

mouse model was critical as patient samples and cells are difficult to obtain. Pathway

analysis was conducted utilizing WT and Met Δ15-HET mouse embryonic fibroblast cell

lines as they share many similarities with osteoblasts: both are a part of the connective-

tissue cells family, secrete the MET receptor and HGF, collagenous extracellular matrix

rich in type-1 collagen and are responsible for the structural framework of the body

(Grano et al., 2002; Vallet et al., 2016). In addition, patient samples are few and far in-

between, and are extremely difficult to come by. For these reasons we felt mouse

embryonic fibroblasts were a suitable alternative in place of osteoblasts to perform the

MET signalling analysis reported here. Protein was extracted from WT and Met Δ15-HET

murine embryonic fibroblast for protein expression analysis of downstream effectors of

MET signaling, along of key players which have been previously implicated in osteoblast

differentiation defect phenotypes. WT and Met Δ15-HET murine embryonic fibroblasts

were also treated with exogenous HGF in timed experiments to examine whether the

mutation was ligand (HGF) dependent.

Gene expression analysis was also conducted on WT and Met Δ15-HET osteoblasts

during the terminal stages of differentiation. This was done as Gray et al., (2015)

reported aberrant MET signaling to dysregulate late stage markers of osteoblast

differentiation while leaving early stage markers unaffected. Since this was completed in

a cell line, we wanted to confirm these results in primary osteoblasts.

Objective 2: Skeletal Characterization of WT and in Met Δ15-HET mice

Human OFD patients develop tibial bowing which may progress to non-healing

fractures. The second aim of this study was to determine if there were any

endochondral or intramembranous bone differences between WT and Met Δ15-HET

littermate mice. We first examined the osteogenic potential of bone marrow stromal cells

(BMSCs), which are mesenchymal stem cells originating from the bone marrow stroma.

Page 53: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

41

We investigated the osteoblast differentiation process since it is involved in both forms

of ossification.

While Met Δ15-HET mice appear to be phenotypically normal, they have not been

extensively studied and characterized to examine bone differences. Full-body

dissections and skeletal preparations of male mice at the pup age (P5) and the more

skeletally mature age (P21) were performed to visualize any gross defects or

differences. These timepoints were chosen as P5 mice acted as a pre-pubescent young

immature skeleton, representing the age group (<10 years old) which is most commonly

affected by OFD. P21 mice acted as the post-pubescent adult mature skeletons,

(representing adult humans) and used to visualize longer term effects of the MET

mutation. Stained skeletons were photographed and further dissected to isolate long

bones (femur, tibia, humerus, radius) for manual and digital measurements by two-

blinded individuals. Measurements were analyzed to observe any statistical

significance. This was followed by examination of the long bone epiphyseal growth plate

zones of WT and Met Δ15-HET P21 littermate mice to examine if there were any

differences in zone heights.

Objective 3: Examination of Fracture Healing in WT and Met Δ15-HET mice in vivo

The third aim of this study was to examine the effects of the Met exon 15 skipping

mutation on fracture healing in vivo. This was done to test our hypothesis that the

phenotype seen in humans may be a stress induced one, as healthy tissue surrounding

OFD lesions do not experience defects in osteoblast differentiation. Transverse tibial

fractures were induced in 12-week old male mice utilizing a semi-stabilized tibial fracture

model. Progression of fracture healing was followed at 14-days post-fracture induction.

This timepoint was chosen as osteoblasts are involved in three key pathways occurring

simultaneously at this point in fracture repair: intramembranous ossification,

endochondral ossification, and callus remodeling. In addition, the 14-days post fracture

timepoint marks the peak of hard callus formation (Marsell & Einhorn, 2011). Quality of

healing was examined by analyzing callus size and mineralization at the aforementioned

time-point.

Page 54: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

42

2.4 Clinical Significance

There is currently no cure for osteofibrous dysplasia, a quality of life impacting disease,

with treatment limited to pain management, stabilization of weakened tibias, or

osteotomies to correct severe deformities. Since OFD patients present with reduced

fracture healing abilities, understanding the underlying mechanisms responsible for their

phenotype may enable the identification of therapeutic targets and development of

novel therapeutic treatments. Furthermore, the knowledge gained from this study could

also be applied to other diseases displaying similar pseudarthrosis phenotypes such as

neurofibromatosis-1, or in the elderly where fracture healing abilities are reduced. By

improving the clinical outcomes of fracture healing, we hope to improve quality of life by

enabling patients to walk independently and perform their everyday activities.

Page 55: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

43

3 Methods

3.1 Generation of a Met exon 15 splice donor mutant allele by CRISPR/Cas9-mediated genome editing Cas9 guide RNAs targeting the splice donor sequence of mouse Met exon 15 were

identified using the crispr.mit.edu site. Two guide RNAs targeting the exon 15 splice

donor sequence were selected for activity testing. Guide RNAs were cloned into a T7

promoter vector followed by in vitro transcription and spin column purification. Guide

RNA functional testing was performed by an in vitro cleavage assay incubating Cas9

protein and guide RNA with PCR-amplified target site. The guide RNA selected for

genome editing in embryos was Met-g68B (protospacer sequence 5’-

GTAAACTGAATTATACCTTC-3’). The donor oligonucleotide used to insert the splice

site mutation was Met-KI-B (5’-

AGCTCACAGAGGTCTATGTATAGATATTTCTCAGGATAGTAAACTGAATTATAACTT

CTGGAAAAGTAGCTCTGTAGTCTACAGACTCATTTGAAACCATCTCTGTAG-3’).

Cas9 mRNA was produced by T7 in vitro transcription.

3.2 Embryo Microinjection C57BL/6J zygotes were microinjected with Mix1 (10 ng/ul Cas9 mRNA, 5 ng/ul guide

RNA and 100 ng/ul donor oligonucleotide) or Mix2 (10 ng/ul Cas9 mRNA, 5 ng/ul guide

RNA and 10 ng/ul donor oligonucleotide) and implanted in recipient pseudopregnant

females. Resulting pups were screened by PCR and sequencing for the presence of the

mutant allele. One male founder from Mix1 and 1 male and 1 female founder from Mix2

were found to harbor the desired mutation.

3.3 Confirmation of Exon 15 Skipping The exon 15 skipping event was confirmed in our genetic mouse model by examining

the Met transcript for the presence or absence of exon 15. This was done by designing

PCR primers targeted to Exon 14 and 16 of the Met transcript. Met transcripts

containing exon 15 would result in formation of a 469 bp PCR product while the

exclusion of exon 15 would result in a 319 bp PCR product. Gel electrophoresis

Page 56: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

44

mediated resolution was utilized to distinguish the differently sized PCR products.

Primers are as follows: Forward Primer: GGAAGCAAGCAGTCTCTTCAAC, Reverse

Primer: TGCTGAACTGCTTGGACCAG.

3.4 Maintenance and Genotyping of Mice Mice used in this study were housed in standardized cages at The Centre for

Phenogenomics in Toronto, ON according to their standards and guidelines. Mouse

genotyping was completed DNA extraction from mice tail clips done using DNA

extraction reagent from QuantaBio’s Extracta DNA Prep for PCR-Tissue (84158). 50 ul

of DNA extraction reagent is used per tail clipping. Incubated at 95oC for 35 minutes. 50

ul of DNA stabilizing buffer from QuantaBio’s Extracta DNA Prep for PCR-Tissue

(84159) was added post incubation. Concentration of the DNA extracted was

determined by nanodrop. Concentrations are used to calculate the dilution required of

extracted DNA to 5ng/mL. qPCR Master Mix Composition Per Sample: 10 uL of

TaqMan 2X Universal PCR Master Mix (4324018), 0.5 uL of Custom Taqman™ SNP

Genotyping Assay (4332077), 0.5 uL H2O. Primers are as follows: Forward Primer:

CCAACTACAGAGATGGTTTCAAATGAGT, Reverse Primer:

CACAGCTCACAGAGGTCTATGTATAGATAT. Probes are as follows: Probe 1:

CTTTTCCAGAAGGTATAATT, Probe 2: CTTTTCCAGAAGTTATAATT

Page 57: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

45

3.5 Murine Embryonic Fibroblast Generation WT and Met Δ15-HET murine embryonic fibroblasts were derived following WiCell’s

“Derivation of Mouse Embryonic Fibroblasts (MEFs)” protocol. In short, timed-pregnant

mice were sacrificed on postnotum day E11.5, where 0.5 is the day of detection of a

copulation plug, by cervical dislocation. Pregnant mice were prepped by dissection by

sterilization of mice abdomen. An incision in the peritoneal wall was made until

exposure of the uterine horns. Uterine horns were extracted from the mother’s carcass,

washed with PBS. Embryos were released from embryonic sacs and washed with PBS,

followed by separation of visceral tissue from the embryos. Embryos were minced using

curved dissecting scissors into grain sized pieces (minced for approximately 5 – 1

minutes), followed by 2 ml of trypsin and additional mincing to ensure pieces were

further reduced in size. Lastly, add 5 ml more trypsin, pipetting cells vigorously up and

down to freeing of the cells from minced tissue. Incubate cells in a T75 flask, allowing

for growth and adherence, when cells reached 90% the culture were harvested to be

Heterozygous Mutant: GT

Homozygous Mutant: TT

Negative Control

(Water)

Homozygous

WT: GG

Typical allelic discrimination plot. Real-time PCR instrument software displaying the results of the

allelic discrimination data as a plot wild-type (GG) versus Met homozygous (TT).

Page 58: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

46

stored in canonical tubes with Cryopreservation media. Cells were aliquoted into 1.5 ml

cryovials and flash-frozen with dry ice to be stored in -80oC freezer overnight and

moved to liquid nitrogen storage the following day for long term storage. Vials were

defrosted, re-suspended and cultured as required with Dulbecco’s Modified Eagle

Medium F/12 (Gibco) with 5% anti-anti and 10% fetal bovine serum.

3.6 MET Pathway Analysis WT and Met Δ15-HET murine embryonic fibroblasts were allowed to be cultured until

90% confluency with Dulbecco’s Modified Eagle Medium F/12 (Gibco) with 5% anti-anti

and 10% fetal bovine serum in a T75 flask. Culture media was removed and washed

with PBS to ensure removal of all residual media. Cells were detached and freed with

trypsin to form cell pellets in conical tubes for RNA and protein extraction.

3.7 Activation of MET pathway WT and Met Δ15-HET murine embryonic fibroblasts were allowed to be cultured until

50% confluency with Dulbecco’s Modified Eagle Medium F/12 (Gibco) with 5% anti-anti

and 10% fetal bovine serum in a T25 flask. 10 ng/ml HGF treatment was directly applied

to the culture media for 0, 5, and 30 minutes. Following these timepoints, HGF culture

media was removed and cells were washed with PBS to ensure removal of all residual

media. Cells were detached and freed with trypsin to form cell pellets in conical tubes

for RNA and protein extraction.

3.8 Western Blot Analysis

To examine protein levels of MET and its downstream effectors, protein was extracted

from WT and Met Δ15-HET MEFs. Cell cultures were pelletized, and protein extracted

using radioimmunoprecipitation assay buffer (RIPA buffer). Equal amounts of total

protein were resolved utilizing SDS-polyacrylamide gel-based electrophoresis, then

transferred to nitrocellulose membranes and immunoblotted at 4oC overnight with the

indicated primary antibodies: MET (Santa Cruz Biotechnology), phospho-MET (Cell

Signaling Technology), AKT (Cell Signaling Technology), phospho-AKT (Cell Signaling

Technology), mTOR (Cell Signaling Technology), phospho-mTOR (Cell Signaling

Technology), MEK (Cell Signaling Technology), phospho-MEK, ERK (Cell Signaling

Page 59: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

47

Technology), phospho-ERK (Cell Signaling Technology), RAS (Cell Signaling

Technology), and β-catenin (BD Transduction Laboratories). Horseradish Peroxidase

(HRP) tagged secondary antibodies and ChemiLuminescence were utilized to detect

hybridization. Expression of the targeted proteins were quantified relative to the

expression of Actin or Vinculin as the housekeeping proteins. Quantification of Met Δ15-

HET protein expression was performed utilizing Image Lab quantification tools relative

to WT levels. Western blot analysis was carried out in triplicates to make sure the

experiments are reproducible.

3.9 Colony Forming Unit – Fibroblast Femora and tibiae from 12-week old male WT and Met Δ15-HET mice were isolated and

debrided of soft tissue. Bone marrow was flushed into BMSC culture media containing

α-modification of eagle’s medium (α-MEM) (Wisent Inc, Cat# 310-012-CL, St. Bruno,

CA), 10% FBS (Gibco Life Technology, Cat# 16000-044, Burlington, ON) and 1x

antibiotic and antimycotic (Wisent Inc, Cat# 450-115-EL). The cell suspensions were

passed through an 18G needle and 70μm cell strainer to dissociate clumps of cells.

Single cell suspensions were plated at a density of 1x106cells/cm2 and 5x106cells/cm2

surface area on 6-well plates, in BMSC culture medium for 7 days, with 50% of the

media changed at day 4. At day 7, plates were stained with crystal violet in 25%

methanol for 15 minutes. CFUs were counted manually using a microscope. CFU-F

colonies were defined as a discrete colony that contained 30 or more cells with the

majority of cells in each colony staining positively for crystal violet. All quantification

performed with 3 wells per mouse, and minimum of 3 mice per genotype.

3.10 Colony Forming Unit – ALP and Osteoblast Femora and tibiae from 12-week old male WT and Met Δ15-HET mice were isolated and

debrided of soft tissue. Bone marrow was flushed into BMSC culture media containing

α-modification of eagle’s medium (α-MEM) (Wisent Inc, Cat# 310-012-CL, St. Bruno,

CA), 10% FBS (Gibco Life Technology, Cat# 16000-044, Burlington, ON) and 1x

antibiotic and antimycotic (Wisent Inc, Cat# 450-115-EL). The cell suspensions were

passed through an 18G needle and 70μm cell strainer to dissociate clumps of cells.

Single cell suspensions were plated at a density of 1x106cells/cm2 and 5x106cells/cm2

Page 60: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

48

surface area on 6-well plates, in BMSC culture medium for 7 days, with 50% of the

media changed at day 4. At day 7, the media was changed for osteoblastic

differentiation media, which consisted of BMSC culture media, as described above, with

the addition of 50μg/ml Ascorbic Acid (Sigma, Cat# A4544), 10-8M Dexamethasone

(Sigma, Cat# D8893), 8mM β-Glycerol phosphate (Sigma, Cat# G9891). The

osteoblastic differentiation media was subsequently changed every 2-3 days for the

duration of the experiment. On days 14 and 21 with OB differentiation media cells were

rinsed with PBS and fixed with 10% formalin for 30 mins and rinsed twice with water.

Alkaline phosphatase staining (Fast Red TR/Naphthol AS-MX Tablets: Sigma#F4523,

manufactory protocol) and Von Kossa staining (2.5% (w/v) Silver Nitrate: Sigma

#S8157, 30 minutes at room temperature on light box) were performed to assess for

osteoblastic differentiation and mineralization. CFUs were counted manually using a

microscope. CFU-ALP and CFU-OB colonies were defined as a discrete colony that

contained 30 or more cells with the majority of cells in each colony staining positively for

alkaline phosphatase and Von Kossa respectively. All quantification performed with 3

wells per mouse, and minimum of 3 mice per genotype.

3.11 RT-PCR: Gene Expression Analysis RNA was extracted from cell pellets using Monarch Total RNA Miniprep Kit (T2010S)

and used to create cDNA using LunaScript RT SuperMix Kit (E3010L), following

manufacturer’s protocol and instructions. cDNA was targeted with primers targeting

genes of interest from RNA-Seq and western blot analysis data. qPCR analysis was

carried out in triplicates to make sure the experiments are reproducible.

3.12 Skeletal Prep Staining For staining of bone and cartilage, whole skeletons were dissected from mice and fixed

in 95% ethanol for 48 hours in room temperature. Murine skeletons were submerged in

Alcian blue staining solution (2.5ml 0.3% Alcian Blue SGS (Sigma), 10ml glacial acetic

acid, and 40ml ethanol) for 48 hours at 37°C. Alcian blue staining solution was replaced

with 95% ethanol daily over 3 days. Skeletons were then submerged in Alizarin red

staining solution (0.5ml 0.2% Alizarin Red S (Sigma), 5ml 10% KOH, and 45ml distilled

water) for 24 hours in room temperature. Alizarin red staining solution was then

replaced with 20% glycerol, 1% KOH for 3 days in room temperature and then 50%

Page 61: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

49

glycerol, 1% KOH until muscle and fat tissue was dissolved. This dissolving step varied

in time depending on the mouse size and age. Skeletons were then stored in 80%

glycerol for 24 hours and then 100% glycerol for long-term storage.

3.13 Epiphyseal Growth Plate Staining Mouse limbs were harvested in PBS and stored in 70% ethanol prior to paraffin

embedding and sectioning. Limbs were then paraffin embedded and sectioned. Slides

were rehydrated in the series of following ethanol concentrations: 100%, 90%, 70%,

50%, then washed twice in distilled water for 3 minutes. For H&E staining, slides were

then immersed in 0.3% ammonium hydroxide for 20 dips and rinsed twice in water for 1

minute, followed by immersion in Eosin Y certified biological stain (Fisher Scientific) for

10 dips. For Toluidine blue staining, slides were then immersed in 1% Toluidine Blue O

(Sigma) for 10 minutes. For Safranin-O staining, slides were then immersed in Weigert’s

iron hematoxylin solution (Sigma) for 10 minutes and rinsed in distilled water followed by

immersion in 1% Safranin O (Sigma) for 5 minutes. After all staining procedures, slides

were washed in water thrice for 1 minute, dehydrated in a 95%, 100% ethanol and

xylene series, then mounted and covered.

3.14 Visualization and Measurement of Long Bones and Epiphyseal Growth Plates Skeletal staining preparations of mice were photographed with a Canon DSLR over a

backlight while utilizing a custom cardboard aperture to maintain the same

magnification. A ruler was placed in each shot for measurement references, and images

were rotated in ImageJ and PowerPoint. Measurements of limbs were performed using

a Mastercraft digital caliper. The measurements were performed blinded by two

individuals and with each bone measured twice and its averaged value reported for

statistical analyses.

For P5 mice, humerus bones were measured from the most proximal staining point (at

the articulation with the distal glenoid cavity) to the most distal staining point (at the

articulation with the proximal ulnar aspect). Radius bones were measured from the most

proximal staining point (at the head of the radius) to the most distal staining point (at the

condyle). Femur bones were measured from the most proximal staining point (at the

Page 62: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

50

articulation with the distal acetabulum) to the most distal staining point (preceding the

knee femoral cartilage). Tibia bones were measured from the most proximal staining

point (proceeding the knee tibial cartilage) to the most distal staining point (at the

articulation with the talus bone).

For P21 mice, humerus bones were measured from the most proximal staining point (at

the head of the humerus) to the most distal staining point (at the condyle). Following

suit, radius bones were measured from the most proximal staining point (at the head of

the radius) to the most distal staining point (at the condyle). Femur bones were

measured from the most proximal staining point (the femoral head and greater

trochanter) to the most distal staining point (at the femoral condyles). Tibia bones were

measured from the most proximal staining point (the head and tuberosity) to the most

distal staining point (at the malleoli).

Epiphyseal plates were visualized and photographed microscopically, then transferred

to ImageJ and PowerPoint for cropping. Measurements of epiphyseal plate zones and

cellularity were performed using ImageJ, then recorded using the Region of Interest

(ROI) Manager. The measurements were also performed blinded by two individuals and

each replicate was measured twice with its averaged value reported for statistical

analyses.

3.15 Tibia Fracture Generation Semi-stabilized tibial fractures were generated on 12-week-old male WT and Met Δ15-

HET mice following a protocol previously described (Y. Chen et al., 2007). Briefly, the

mice were anaesthetized using inhalational isoflurane general anesthesia. The left hind

limb of each mouse was surgically prepared by shaving and cleaning with povidone-

iodine as disinfectant. A small anterior midline incision was made over the knee joint

and proximal tibia. A 0.7 mm pilot hole was made at the proximal tibial epiphysis, medial

to the insertion of the patella tendon using a hollow 27G needle. Intramedullary fixation

was completed via a 0.7mm Anticorro insect pin (Fine Science Tools,

http://www.finescience.ca) was then inserted into the medullary cavity of the intact tibia

and advanced to the distal tibia. A transverse fracture was then induced at the mid-shaft

of the tibia using blunt scissors. The insect pin was cut 5-7mm proud of the proximal

tibial epiphysis and the skin incision closed with a series of absorbable vicryl rapide

Page 63: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

51

sutures and metallic wound clips. Analgesic given subcutaneously (buprenorphine,

0.1mg/kg/twice per day) was administered for 3 days post-surgery. Previous data shows

that a fracture generated in this manner heals through both intramembranous and

endochondral ossification (Hiltunen, Vuorio, & Aro, 1993; Le, Miclau, Hu, & Helms,

2001). The animals were allowed to maintain free and full weight bearing in their cages

following surgery. At specific time points (14 days) after the fracture, samples of

fractured and unfractured tibiae were harvested following euthanasia with inhaled

carbon dioxide. A minimum of 3 mice per genotype, per time point, were used for further

analysis.

3.16 Histological Staining of Fracture Callus Fractured tibiae were harvested on post-fracture day 14, fixed in 10% formalin following

by decalcification with 20% (w/v) EDTA pH 8.0 or formic acid bone decalcifier (Decal

Chemical Corp, Tallman, NY) and embedded in paraffin. Serial 5 μM sections of paraffin

embedded tissues were deparaffinized and rehydrated through an alcohol gradient to

water. Sections were stained with Safranin-O and counter stained with fast

green/Mayer’s haematoxylin. Here red staining confirms the presence of proteoglycans,

which indicates cartilaginous tissue, and green staining indicates bone (Camplejohn &

Allard, 1988), or stained with Tartrate Resistant Acid Phosphatase (TRAP)

(manufacturers protocol, TRAP Staining- 387A-1KT, Sigma, St. Louis, MO) to evaluate

osteoclast activity. Here red staining confirms the presence of proteoglycans, which

indicates cartilaginous tissue, and green staining indicates bone (Camplejohn & Allard,

1988), or stained with Tartrate Resistant Acid Phosphatase (TRAP) (manufacturers

protocol, TRAP Staining- 387A-1KT, Sigma, St. Louis, MO) to evaluate osteoclast

activity.

3.17 Statistical Analyses Data (mouse weights, bone measurements, epiphyseal plate measurements,

chondrocyte experiments, etc.) were analyzed using the Student’s two-tailed

heteroscedastic t test by comparing all test groups (heterozygous mutant) to

corresponding control groups (wildtype). The ANOVA test was not utilized as multiple

group comparisons were not performed in any analyses. Significance was defined as

Page 64: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

52

the p-value (P), with *P <0.05. Means and error bars were graphed using Microsoft

Excel chart and error bar formatting tools.

Page 65: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

53

4 Results

4.1 Met Δ15-HET mouse embryonic fibroblast exhibit higher levels of MET protein and upregulated MET signaling WT and Met Δ15-HET mouse embryonic fibroblasts (MEFs) were subjected to protein

expression analysis to observe modulations and activity of downstream effectors of the

MET signaling pathway. This experiment was designed to test whether our mouse

model demonstrated similar disruptions in MET signaling seen in OFD patients. Murine

embryonic fibroblasts were cultured until 90% confluency in a T75 flask then prepared

for protein extraction. Protein was extracted and separated by SDS-PAGE and probed

for MET and its downstream effectors: AKT, mTOR MEK, ERK and their activated

phosphorylated versions.

MET was chosen to test if similar changes to the protein level as seen in human OFD

(Figure 1A) was replicated in the Met Δ15-HET mouse model, which mimics the identical

loss of regulatory control as humans. Protein kinase B (AKT) is a downstream MET

effector and was chosen as it has been implicated in the regulation of Runx2, a key

marker of early osteoblast differentiation. AKT regulates RUNX2 through Wnt/β-catenin

dependent and independent signaling pathways. Mammalian target of rapamycin

(mTOR), a downstream effector of AKT, was chosen as it is also involved in the

regulation of Runx2 and Wnt/β-catenin signaling through the mTORC1 complex.

Mitogen-activated protein kinase kinase (MEK1/2 or MAP2K) and mitogen-activated

protein kinase 1 (ERK1/2 or MAPK1), downstream effectors of RAS signaling (which is

connected to MET as shown previous in Figure 1A), have also been previously been

shown to directly regulate Runx2 expression and osteoblast differentiation. This

experiment tested whether signaling changes in Met Δ15-HET mice are similar to those

observed in human OFD. The utilization of a mouse model was required to overcome

the difficulties in obtaining OFD patient bone samples and cells.

Met Δ15-HET MEFs exhibited higher phosphorylation at tyrosine residues (Y1234 and

Y1235) found in the catalytic domain of the receptor, indicative of greater MET catalytic

Page 66: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

54

signaling activity in comparison to WT MEFs (Fig 4.1 A). Accordingly, downstream

effectors of MET: AKT, mTOR, MEK1/2, ERK1/2 followed suit and also exhibited

upregulated signaling activity with increased levels of phosphorylation in Met Δ15-HET

MEFs (Fig 4.1 B & C). Surprisingly, Met Δ15-HET MEFs also displayed higher levels of

unphosphorylated MET, mTOR, MEK, ERK1/2 in comparison to their WT counterparts

(Fig 4.1 A – C).

Page 67: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

55

WT HET WT HET WT HET

Y1234/Y1235

Figure 4.1. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts

Representative images of western blot protein analysis of MET (A) and its downstream effectors: AKT, mTOR, MEK, ERK1/2 (B & C) and their phosphorylated versions (B & C) in WT and MetΔ15-HET mouse embryonic fibroblasts (MEFs). MetΔ15-HET MEFs display

upregulated MET protein along with phosphorylation indicative of increased signaling activity. Accordingly, downstream effectors of MET display also display upregulated

signaling in MetΔ15-HET MEFs. Actin and Vinculin are the loading controls. N=3

A B C

Page 68: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

56

4.2 Met Δ15-HET mouse embryonic fibroblast exhibit upregulated and prolonged MET signaling after HGF stimulation

WT and Met Δ15-HET mouse embryonic fibroblasts (MEFs) were treated with 10ng/ml

HGF for 0, 5, and 30 minutes. HGF is the only known ligand for MET. The experimental

timepoints were chosen as the half-life of HGF is approximately 3 – 5 minutes, therefore

peak signaling activity should be seen at 5 minutes and returned to basal levels at the

30-minute timepoint (Chang et al., 2016). Treated cells were subjected to protein

expression analysis to observe modulation, activity, and duration of activity in

downstream effectors of the MET signaling pathway. The results of this experiment

would allow us to examine if the dysregulation in signaling is ligand dependent.

Murine embryonic fibroblasts were cultured until 50% confluency in a T25 flask then

treated with 10ng/ml HGF for the aforementioned durations of time and prepared for

protein extraction. Protein was extracted and separated by SDS-PAGE and probed for

MET and its downstream effectors: AKT, mTOR, RAS, ERK1/2, β-catenin and their

activated phosphorylated versions.

MET, AKT, mTOR, RAS, and ERK1/2 were chosen for their involvement in regulation of

Runx2 expression and the osteoblast differentiation process as described in the

previous experiment. We chose to investigate β-catenin since the tight spatiotemporal

regulation of β-catenin signaling is crucial during osteoblast differentiation and

deviations compromise the differentiation process (Y. Chen et al., 2007). β-catenin’s

Serine 552 phosphorylation activity was examined as it is a phosphorylation target of

AKT. Phosphorylation at this site enhances β-catenin stability and transcriptional

activity.

Consistent with our previous data Met Δ15-HET MEFs exhibited greater phosphorylation

at tyrosine residues (Y1234 and Y1235) found in the catalytic domain of the receptor

prior to HGF stimulation (Fig 4.2 A & D). The higher levels of unphosphorylated MET

protein in comparison to WT MEFs may be due to reduced ability for the degradation

and terminal of signalling. At 5-minutes post HGF treatment, Met Δ15-HET MEFs

Page 69: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

57

exhibited greater upregulation in phosphorylation at the Y1234 and Y1235 catalytic

tyrosine residues indicating increased MET signaling activity in comparison to WT MEFs

(Fig 4.2 A). Accordingly, downstream effectors of MET: AKT, mTOR, ERK and β-catenin

followed suit and also exhibited greater upregulated activity with increased levels of

phosphorylation 5-minutes post HGF treatment (Fig 4.2 A - M). This is further proof of

upregulated MET signaling pathway activity. Surprisingly, Met Δ15-HET MEFs displayed

higher unphosphorylated protein levels of MET, AKT, mTOR, RAS, ERK, β-catenin post

HGF treatment in comparison to their WT counterparts as well (Fig 4.2 A - M). This is

surprisingly as we previously believed the Met Δ15-HET mutation would only alter

phosphorylation activity at downstream effectors of MET. This indicates aberrant MET

signaling must modulation expression of these proteins as well, in addition to changes in

signaling activity.

Met Δ15-HET MEFs exhibit prolonged MET signaling following HGF stimulation; we

found prolonged phosphorylation activity at the MET receptor and its aforementioned

downstream effectors 30-minutes post HGF stimulation (Fig 4.2 A & B). This contrasts

the behaviour of WT MEFs, which returned to basal levels of phosphorylation activity at

the 30-minute timepoint, exhibiting similar activity to the 0-minute timepoint where no

HGF stimulation was applied (Fig 4.2 A & B).

Page 70: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

58

MET

p-MET

AKT

p-AKT

Actin

mTOR

p-mTOR

p-AKT

AKT

MET

p-MET

Actin

Y1234/Y1235

Figure 4.2. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes

Representative images of western blots of MET and its downstream effectors: AKT, mTOR and their activated phosphorylated versions (A) in WT and MetΔ15-HET MEFs. When HGF stimulation is applied MetΔ15-HET MEFs display upregulated MET protein along with

phosphorylation indicative of increased signaling activity. In addition, MetΔ15-HET MEFs

display prolonged signaling in comparison to their WT counterparts. Accordingly, downstream effectors of MET display also display upregulated and prolonged signaling in MetΔ15-HET MEFs. Actin is the loading control. N=3

A

Page 71: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

59

β-Catenin

p-β-Catenin (Ser 552)

Actin

ERK

RAS

Figure 4.2 Contd. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes

Representative images of western blots of downstream effectors of MET: RAS, ERK, β-catenin and their phosphorylated versions (B) in WT and MetΔ15-HET MEFs. When HGF stimulation is applied MetΔ15-HET MEFs display upregulated MET protein along with

phosphorylation indicative of increased signaling activity. In addition, MetΔ15-HET MEFs

display prolonged signaling in comparison to their WT counterparts. Accordingly, downstream effectors of MET display also display upregulated and prolonged signaling in

MetΔ15-HET MEFs. Actin is the loading control. N=3

B

Page 72: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

60

0

1

2

3

4

5

0 5 30

Re

lati

ve E

xpre

ssio

n v

s W

T

Minutes with HGF

MET

WT HET

0

1

2

3

4

5

0 5 30

Re

lati

ve E

xpre

ssio

n v

s W

T

Minutes with HGF

pMET

WT HET

0

1

2

3

4

5

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

pAKT

WT HET

0

1

2

3

4

5

0 5 30

Re

lati

ve E

xpre

ssio

n v

s W

T

Minutes with HGF

mTOR

WT HET

0

1

2

3

4

5

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

pMTOR

WT HET

C D

E F

G H

0

1

2

3

4

5

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

AKT

WT HET

* *

* *

* * * * *

* *

* *

N.S. *

*

*

*

Page 73: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

61

0

1

2

3

4

5

0 5 30

Re

lati

ve E

xpre

ssio

n v

s W

T

Minutes with HGF

RAS

WT HET

0

1

2

3

4

5

6

7

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

ERK1/2

WT HET

0

1

2

3

4

5

6

7

0 5 30

Re

lati

ve E

xpre

ssio

n v

s W

T

Minutes with HGF

pERK1/2

WT HET

0

1

2

3

4

5

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

β-Catenin (Ser 552)

WT HET

I J

K L

M

0

1

2

3

4

5

0 5 30

Rel

ativ

e Ex

pre

ssio

n v

s W

T

Minutes with HGF

β-Catenin

WT HET

Figure 4.2 Contd. Protein expression analysis of WT and Met Δ15-HET mouse embryonic fibroblasts treated with 10ng/ml HGF stimulation for 0, 5, and 30 minutes Quantification of western blots of downstream effectors of MET: AKT, mTOR, RAS, ERK, β-

catenin and their phosphorylated versions normalized to WT levels (C – M). Student’s two-tailed t test was performed for statistical analysis comparing WT to MET Δ15-HET

measurements, with significance level set at *P<0.05. N=3

*

* *

*

*

*

* *

*

N.S.

N.S.

N.S. N.S.

N.S. N.S.

Page 74: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

62

4.3 Met Δ15-HET MEFs exhibit upregulated levels of β-catenin

RNA-sequencing and protein expression analysis uncovered candidate genes of

interest for reverse transcription polymerase chain reaction (RT-PCR) validation

analysis. The following genes were chosen to be examined as previous experiments

revealed the potential for them to be a part of the explanation for delayed fracture

healing in OFD. WT and Met Δ15-HET MEFs were cultured until 80% confluency and

processed for RNA extraction followed by cDNA synthesis and RT-PCR analysis for the

following genes: Met, Axin2, β-catenin, Erk, Akt.

Since we saw higher levels of MET at the protein level in Met Δ15-HET MEFs, we wanted

to confirm this was caused by a failure to degrade the receptor rather than upregulated

Met expression caused by aberrant MET signaling. Therefore, Met expression was

examined between WT and Met Δ15-HET MEFs to observe any potential differences.

Here we saw no significant changes in Met expression between WT and Met Δ15-HET

mutants, indicating the differences in MET at the protein level may be due to the failure

to degrade the receptor rather than changes in gene expression (Fig 4.3 B). β-catenin,

Erk, and Akt were chosen as we saw upregulation of their protein levels in Met Δ15-HETs

during western blot analysis (Fig 4.2 A & B) and therefore wanted to check whether

there were associated gene expression changes. These three proteins are also

important in the osteoblast differentiation process requiring tight regulation of their

expression and activity (Ghadakzadeh et al., 2016; Greenblatt et al., 2010). Here we

saw a 6.43-fold, 6.88-fold, and 7.4-fold upregulation in expression of β-catenin, Erk, and

Akt respectively confirming our previous results indicating the MET mutation also affects

gene expression (Fig 4.3 D – F).

While examining the dysregulation of β-catenin levels, one must also examine if there is

any aberrant degradation of it as well, therefore Axin2 was chosen as it is a critical

scaffold protein of the β-catenin degradation complex consisting of GSKβ and APC

(Shang, Hua, & Hu, 2017). Here we saw a 12-fold decrease in Axin2 expression (Fig 4.3

C), indicating reduced degradation of β-catenin in Met Δ15-HET MEFs along with the

increased expression of β-catenin.

Page 75: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

63

Overall these results indicate aberrant MET signaling may be caused by the failure to

degrade the MET receptor in a negative-feedback fashion and not caused by elevated

Met gene expression. As we observed higher levels of MET protein and activation of

downstream effectors in Met Δ15-HETs despite no changes in Met expression. They also

show the dysregulation of β-catenin levels in Met Δ15-HET MEFs occurs through multiple

diverse pathways.

Page 76: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

64

A

Figure 4.3 Met Δ15-HET MEFs exhibit gene expression dysregulation resulting in upregulation of β-catenin RT-PCR gene expression analysis of Met, Axin2, β-catenin, Erk, Akt. Expression

changes shown as log2 fold change.

-1

-0.5

0

0.5

1

1.5

2

2.5

3

3.5

MET AXIN2 β-CATENIN ERK AKT

log

2 F

old

Ch

an

ge

Log2 Fold Change Gene Expression

MET

AXIN2

β-CATENIN

ERK

AKT

Page 77: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

65

0

1

2

3

4

5

6

7

8

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

T

AKT

0

1

2

3

4

5

6

7

8

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

T

ERK

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Re

lati

ve

Ex

pre

ssio

n v

s W

T

AXIN2

0

1

2

3

4

5

6

7

8

WT HET

Re

lati

ve E

xpre

ssio

n v

s W

T

β-CATENIN

B

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Re

lati

ve

Ex

pre

ssio

n v

s W

T

MET C

D E

F

N.S. *

* *

*

Figure 4.3 Contd. RT-PCR gene expression analysis of Met, Axin2, β-catenin, Erk, Akt. Expression

changes shown as fold-change normalized to WT expression (B – F). Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET

measurements, with significance level set at *P<0.05. N=3

Page 78: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

66

4.4 Met Δ15-HET bone marrow stromal cells exhibit reduced osteoblast differentiation and bone mineralization

In order to visualize differences in osteogenic capacity in vitro, we utilized bone marrow

stromal cells (BMSCs) isolated from 10-week-old male WT and Met Δ15-HET mice’s

femurs. Only male mice were used in this experiment to account for sex-specific

differences in osteogenic potential. 10-week-old aged mice were used as 8 – 12-week-

old mice provide the most reliable source of bone marrow stromal cells (Beane,

Fonseca, Cooper, Koren, & Darling, 2014; Nadri et al., 2007). BMSCs were plated at

1x106 cells/well in a 6-well dish and induced to undergo osteoblast differentiation after 7

days of rest using osteoblast (OB) differentiation media consisting of ascorbic acid,

dexamethasone, and β-Glycerol phosphate.

Colony forming units – fibroblasts (CFU-F) were quantified at day 7 of BMSC rest by

staining with crystal violet and examining colonies staining positively for crystal violet.

This assay provides a means to assess the proliferation and cologenic capacity of the

cells when expanded in culture (Nadri et al., 2007). Osteoblast cultures were stained for

alkaline phosphatase (ALP) 14 days after addition of OB differentiation media. ALP acts

as a marker of bone matrix secretion and early stage marker of osteoblast

differentiation. At day 21, cultures were stained with Von Kossa, a late stage marker of

osteoblast differentiation, to visualize matrix mineralization. Colony forming units –

alkaline phosphatase (CFU-ALP) were quantified as colonies staining positively for

alkaline phosphatase. Colony forming units – osteoblast (CFU-OB) were quantified as

colonies staining positively for Von Kossa on day 21 of osteoblast differentiation. The

number of bone nodules formed, represents the number of osteoprogenitors in a bone

marrow sample as each colony forms from a single cell precursor (Aubin, 1998). CFU-F,

CFU-ALP, and CFU-OB colonies were defined as discrete colonies which contained 30

or more cells.

Surprisingly, Met Δ15-HET BMSC colonies exhibited a reduced ability to form CFU-

fibroblast. Indicating Met Δ15-HET BMSCs may have a reduced capacity for proliferation,

as each colony is derived from a single-cell precursor (Bianco & Robey, 2000). Met Δ15-

HET BMSCs also formed 23% less CFU-F colonies in comparison to WT BMSCs. They

Page 79: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

67

also exhibited 35% less CFU-ALP colonies and 80% less CFU-OB colonies in

comparison to WT. This indicates Met Δ15-HET BMSCs display reduced osteoblast

differentiation and bone mineralization capacity caused by aberrant MET signaling.

These results also show our Met Δ15-HET mouse model mimics the same osteoblast

differentiation defect seen in human OFD patients and may thus exhibit the same

reduced fracture healing ability as well.

Page 80: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

68

Met Δ15-HET

WT

Figure 4.4. Met Δ15-HET bone marrow stromal cells exhibit reduced mineralization ability Bone marrow stromal cells were obtained from 12-week old male mice. Representative wells of WT and Met Δ15-HET bone marrow stromal cells plated at 1x106 cells/well treated

with osteoblast differentiation media and stained for alkaline phosphatase and Von Kossa 21 days after initiation of the differentiation process (A). N=3

A

Page 81: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

69

Nu

mb

er

of

No

rmali

zed

Co

lon

ies

Nu

mb

er

of

No

rmali

zed

Co

lon

ies

Nu

mb

er

of

No

rmali

zed

Co

lon

ies

CFU-F CFU-ALP CFU-OB

** * ***

Figure 4.4 Contd. Met Δ15-HET bone marrow stromal cells exhibit reduced mineralization ability

Quantification of colony forming units (CFU) plated at 1x106 cells/well normalized to WT to demonstrate reduced ability of Met Δ15-HET BMSCs to form CFU-Fibroblast (CFU-F;

p=0.011), CFU-Alkaline Phosphatase (CFU-ALP; p=0.005) measured on day 14, and CFU-Osteoblast (CFU-OB; p=2.95x10-5) measured on day 21 (B). Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET measurements, with

significance level set at *P<0.05. N=3

B

Page 82: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

70

4.5 Mature Met Δ15-HET osteoblasts display dysregulation of osteoblast specific gene markers

In order to gain a better understanding of the underlying mechanistic changes in the

osteoblast differentiation defect observed in the previous experiment, we examined the

expression of osteoblast specific gene markers at day 21 of differentiation. This

timepoint was chosen as we saw a more severe and striking reduction in osteoblast

differentiation at day 21 of our osteoblast assay experiments, suggesting the Met Δ15-

HET mutation may be interfering with the late stages of the differentiation process.

Here we utilized BMSC induced osteoblast differentiation cell cultures as described

previously. RNA was extracted and cDNA synthesized on day 14 of osteoblast

differentiation, followed by RT-PCR analysis for the following genes: β-catenin, Runx2,

Alp. The same was done for day 21 of osteoblast differentiation, followed by RT-PCR

analysis for the following genes: β-catenin, Axin2, Runx2, Osteocalcin, Adrenomedullin,

Rasal3.

β-catenin, Axin2 were chosen for their involvement in the osteoblast differentiation

process as previously described. At day 14 of osteoblast differentiation we see no

differences in β-catenin expression while we found a 22-fold increase in β-catenin

expression at day 21 of osteoblast differentiation (Fig 4.5 A & B). This indicates there is

a dysregulation of β-catenin expression in day 21 Met Δ15-HET osteoblasts. In addition,

the 6.4-fold reduction in Axin2 expression indicates reduced degradation of β-catenin in

day 21 Met Δ15-HET osteoblasts as well (Fig 4.5 C). Together, these two results

correspond with the upregulation of β-catenin at the protein level in Met Δ15-HET

mutants seen during our western blot analyses (Fig 4.2 B).

Runx2 was chosen as the persistence of its expression during the later stages of

differentiation inhibits terminal differentiation of osteoblasts. In addition, β-catenin

positively regulates Runx2 expression. Therefore, if there is dysregulation in β-catenin

levels, Runx2 should follow suit as well. Alkaline phosphatase and osteocalcin

expression were examined they act as early and late markers of osteoblast maturation

Page 83: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

71

respectively. Thus, we would expect to see a reduction in alkaline phosphatase and

osteocalcin expression in Met Δ15-HET osteoblasts if there was a defect in terminal

differentiation. Met Δ15-HET osteoblasts exhibited a non-significant 1.03-fold increase in

Runx2 expression at day 14 of differentiation, while a 27-fold increase in Runx2

expression was seen at day 21 of differentiation (Fig 4.5 D & E). This upregulation

occurs at a stage at which Runx2 expression should be at its minima. As expected, we

saw a 1.2-fold reduction in alkaline phosphatase expression in day 14 Met Δ15-HET

osteoblasts (Fig 4.5 F). This reduction was present at day 21 of differentiation as well,

with Met Δ15-HET osteoblast exhibiting a 1.3-fold reduction in alkaline phosphatase (Fig

4.5 G). In addition, a 4-fold reduction in osteocalcin expression was seen in day 21 Met

Δ15-HET osteoblasts, indicative of a failure in the terminal osteoblast differentiation

process (Fig 4.5 H).

Adm and Rasal3 were chosen for their involvement in the regulation of β-catenin

stability. (Vagin & Beenhouwer, 2016). We saw a 24.6-fold increase in Adm expression

in day 21 Met Δ15-HET osteoblasts, adrenomedullin acts as a regulator of the inhibition

of GSK3β activity (Fig 4.5 I) (Lausson & Cressent, 2011). In addition, a 1.29-fold

decrease in Rasal3 expression is seen in day 21 Met Δ15-HET osteoblasts, which

functions in the inactivation of RAS activity and β-catenin (Fig 4.5 J) (Muro et al., 2015).

Overall these results indicate aberrant MET signaling secondary to the gain-of-function

MET mutation results in dysregulation of β-catenin mediated Runx2 expression, leading

to inhibition of terminal differentiation in Met Δ15-HET osteoblasts.

Page 84: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

72

C

A

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

T

OB Day 21: AXIN2

*

0

5

10

15

20

25

WT HET

Re

lati

ve E

xpre

ssio

n v

s W

T

OB Day 21: β-CATENIN

*

E

0

5

10

15

20

25

30

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

T

OB Day 21: RUNX2*

B

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Re

lati

ve E

xpre

ssio

n v

s W

T

OB Day 14: β-CATENIN

N.S

D

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

T

OB Day 14: RUNX2

N.S

F

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Re

lati

ve E

xpre

ssio

n v

s W

T

OB Day 14: ALP

*

Page 85: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

73

0

0.2

0.4

0.6

0.8

1

1.2

WT HET

OB Day 21: RASAL3I

0

5

10

15

20

25

30

WT HET

OB Day 21: ADM

G

H

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Re

lati

ve E

xpre

ssio

n v

s W

T

OB Day 21: OSTEOCALCIN

*

Figure 4.5 Met Δ15-HET mature osteoblasts display dysregulation of osteoblast specific gene markers Bone marrow stromal cells were obtained from 12-week old male mice. RT-PCR gene expression analysis of β-catenin, Runx2, Alp on day 14 of osteoblast differentiation and β-catenin, Axin2, Runx2, Alp, Osteocalcin, Adm, Rasal3 on day 21 of osteoblast differentiation.

Expression changes shown as fold-change normalized to WT expression (A – E). Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET

measurements, with significance level set at *P<0.05. N=3

* *

J

0

0.2

0.4

0.6

0.8

1

1.2

1.4

WT HET

Rel

ativ

e Ex

pre

ssio

n v

s W

TOB Day 21: ALP

*

Page 86: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

74

4.6 Body weight and preliminary comparisons of WT and Met Δ15-HET mice

We examined the body weights of WT and Met Δ15-HET mice to rule out a generalized

growth defect secondary to the gain-of-function MET mutation. Mice were selected at

the pup/neonatal age (P5), juvenile age (P21) and adult age (3 month) for imaging and

examination of body weight. These timepoints were chosen to represent pre-pubescent,

adult and adult mice respectively.

At the P5 timepoint, mice were not distinguished based on sex, as sex determination

has been described to be unreliable at this time point (Schlomer et al., 2013). At the P21

and 3-month timepoints, only male mice were used to avoid confounding factors relating

to differences in body weight due to sex.

At P5, decapitated WT and Met Δ15-HET mice do not appear to differ significantly in

body size or average weight (Fig 4.6 A & D). At P21, WT and Met Δ15-HET mice also

appear similar superficially (Fig 4.6 B & D). The same is true at the 3 months old time

point when examining WT and Met Δ15-HET mice (Fig 4.6 C & D). Weights charted over

time also revealed a consistent linear body weight velocity between WT and Met Δ15-

HET mice (Fig 4.6 E). These results preliminarily suggest the Met Δ15-HET mutation

does not cause a developmental phenotype affecting overall size.

Page 87: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

75

WT Met Δ15-HET

Figure 4.6 Body weight and preliminary comparisons of WT and Met Δ15-HET mice

Side-by-side comparison of physical appearance of WT and Met Δ15-HET at P5 to observe

gross superficial differences at pup stage (A). N=6

A 5cm

Page 88: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

76

WT Met Δ15-HET

Figure 4.6 Contd. Side-by-side comparison of physical appearance of WT and Met Δ15-HET at P21 to observe

gross superficial differences at the juvenile stage (B). N=6

B 10cm

Page 89: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

77

WT Met Δ15-HET

Figure 4.6 Contd. Side-by-side comparison of physical appearance of WT and Met Δ15-HET at 3 months old to

observe gross superficial differences at the juvenile stage (C). N=4

C 10cm

Page 90: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

78

`

0

5

10

15

20

25

30

35

WT HET

Av

era

ge W

eig

ht

(g)

3 Months Old Mice

0

0.5

1

1.5

2

2.5

3

3.5

WT HET

Ave

rag

e W

eig

ht

(g)

P5 Mice

0

2

4

6

8

10

12

WT HET

Av

era

ge W

eig

ht

(g)

P21 MiceN.S. N.S.

N.S.

0

5

10

15

20

25

30

p5 p21 3 mos

Ave

rag

e W

eig

ht

(g)

Mouse Age

WT

METΔ15-HET

Figure 4.6 Contd. Representative images of WT vs Met Δ15-HET mice at P5 (A), P21 (B), 3 months old (C). Body weight

measurements of WT and Met Δ15-HET mice at P5, P21, 3 months of age. N=6, 6, and 4 respectively (D). Average body weights of WT and Met Δ15-HET plotted over time (E). Student’s two-tailed t test was

performed for statistical analysis, with significance level set at *P<0.05.

D

E

Page 91: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

79

4.7 Skeletal Staining of P5 WT and Met Δ15-HET mice

We previously showed Met Δ15-HET BMSCs exhibit reduced ability to form osteoblasts

and mineralization nodules. Osteoblasts are heavily involved in the endochondral and

intramembranous ossification processes. Therefore, we determined if there were any

discrepancies in bones formed through these processes. In addition, we examined for

gross skeletal phenotypic differences present in 5-day old mice. P5 mice were sacrificed

by cervical dislocation and dissected for whole-mount skeletal staining. The staining

protocol was used to digest excess muscle and fat tissue, stain bone with Alizarin red

and stain cartilage with Alcian blue to delineate skeletal regions for greater accuracy

when measuring.

No gross body skeletal defects were seen between WT and Met Δ15-HET mice at 5 days

old (Fig 4.7 A). Isolation of the upper limbs showed no significant differences in humerus

and radius lengths between the two (Fig 4.7 C & E). Isolation of the lower limbs

displayed no significant differences in femur and tibia lengths in WT and Met Δ15-HET

mice (Fig 4.7 D & E). These results indicate this mutation does not affect endochondral

ossification and long bone development at 5-days old. No gross differences in skull

length or size were seen at P5 (Fig 4.7 B & F). This suggests the rate of

intramembranous ossification, the process in which skulls are formed, is not affected by

the Met Δ15-HET mutation at this timepoint.

Page 92: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

80

50 m

m

50 m

m

WT Met Δ15-HET A

Page 93: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

81

WT Met Δ15-HET

WT Met Δ15-HET

B

C

10 mm

10 mm

Page 94: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

82

WT Met Δ15-HET

D

10 mm

Figure 4.7. 5-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates

Representative images of 5-day old WT and Met Δ15-HET whole-body, condylobasal lengths,

humerus and radius bones, and femur and tibia bones skeletal preps stained with Alcian Blue

and Alizarin Red for cartilage and bone identification respectively (A-D). Red lines indicate observable ossified regions in each body part.

Page 95: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

83

Figure 4.7. 5-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates

Average measurements of P5 WT and MET Δ15-HET femur, tibia, humerus, radius (E). Average measurements of P5 WT and Met Δ15-HET condylobasal lengths (F). Student’s two-tailed t test

was performed for statistical analysis comparing WT to Met Δ15-HET measurements, with

significance level set at *P<0.05. N = 6.

0

1

2

3

4

5

6

7

8

Avg Femur Avg Tibia Avg Humerus Avg Radius

Ave

rag

e L

en

gth

(m

m)

WT

HET

N.S.

N.S.

N.S.

N.S. E

0

5

10

15

20

25

WT HET

Av

era

ge

Co

nd

ylo

ba

sa

l L

en

gth

(m

m)

F

N.S.

Page 96: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

84

4.8 Skeletal Staining of P21 WT and Met Δ15-HET mice

No gross body skeletal defects were seen between WT and Met Δ15-HET mice at 21-

days-old (Fig 4.8 A). Isolation of the upper limbs showed no significant differences in

humerus and radius lengths between the two (Fig 4.8 C & E). Isolation of the lower

limbs displayed no significant differences in femur and tibia lengths in WT and Met Δ15-

HET mice (Fig 4.8 D & E). These results indicate this mutation does not affect

endochondral ossification and long bone development at 21-days old either. No gross

differences in skull length or size were seen at P21 (Fig 4.8 B & F). This suggests the

rate of intramembranous ossification, the process in which skulls are formed, is not

affected by the MetΔ15-HET mutation at this timepoint as well.

Page 97: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

85

WT Met Δ15-HET 10

0 m

m

10

0 m

m

A

Page 98: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

86

WT Met Δ15-HET

WT Met Δ15-HET

B

C

20 mm

10 mm

Page 99: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

87

WT Met Δ15-HET

D

10 mm

Figure 4.8. 21-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates

Representative images of 21-day old WT and Met Δ15-HET whole-body, condylobasal lengths, humerus

and radius bones, and femur and tibia bones skeletal preps stained with Alcian Blue and Alizarin Red

for cartilage and bone identification respectively (A-D). Red lines indicate observable ossified regions in each body part.

Page 100: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

88

0

2

4

6

8

10

12

14

Avg Femur Avg Tibia Avg Humerus Avg Radius

Ave

rag

e L

en

gth

(m

m)

WT

HET

Figure 4.8. 21-day old Met Δ15-HET mice exhibit no differences in long bone lengths in comparison to WT littermates Average measurements of P21 WT and MET Δ15-HET femur, tibia, humerus, radius (E). Average

measurements of P21 WT and Met Δ15-HET condylobasal lengths (F). Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET measurements, with significance level

set at *P<0.05. N = 6.

N.S.

N.S.

N.S.

N.S.

E

0

5

10

15

20

25

WT HET

Av

era

ge

Co

nd

ylo

ba

sa

l L

en

gth

(m

m)

F N.S.

Page 101: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

89

4.9 Histological analysis of P21 WT and Met Δ15-HET mice epiphyseal growth plate

We previously discovered no differences in endochondral and intramembranous bone

lengths, despite osteoblast being critical in both processes. The epiphyseal growth plate

is where chondrocytes progressively enlarge, hypertrophy and either differentiate to

osteoblasts or undergo apoptosis and mineralization. This is very similar to the

processes which occur in endochondral ossification. Thus, we analyzed the growth plate

to determine if there were irregularities in growth plate morphology between WT and

Met Δ15-HET mice.

Proximal tibia epiphyseal growth plates of 21-day old male WT and MET Δ15-HET

littermates were sectioned and stained with Safranin-O and fast green to visualize

cartilage and bone respectively.

Here we find Met Δ15-HET mice exhibit no significant differences in height or cellular

organization at the resting, proliferating, prehypertrophic and hypertrophic zones (Fig

4.9 A & B). No significant differences in overall epiphyseal growth plate height was seen

either between WT and Met Δ15-HET mice at 21-days old. Overall these results confirm

and further suggest this mutation does not affect long bone development at 21-days old.

Page 102: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

90

A

Figure 4.9 Met Δ15-HET mice exhibit no differences in epiphyseal growth plate lengths in comparison to WT littermates

Safranin-O staining of P21 WT vs Met Δ15-HET tibial epiphyseal growth plates (composed of the

resting, proliferative, prehypertrophic and hypertrophic zones). Representative images shown

here (A).

Resting Zone

Proliferating Zone

Prehypertrophic Zone

Hypertrophic Zone

Page 103: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

91

0

50

100

150

200

250

RZ PZ PHZ HZ Full

Len

gth

m)

Epiphyseal Growth Plate Zone

B

Figure 4.9 Contd. Met Δ15-HET mice exhibit no differences in epiphyseal growth plate lengths in comparison to WT littermates

Average measurements of resting zone (RZ), proliferating zone (PZ), prehypertrophic zone (PHZ), hypertrophic zone (HZ) and full zone (Full) (B). Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET measurements, with significance level set

at *P<0.05. N = 3

N.S.

Zones: RZ = Resting

PZ = Proliferating PHZ = Prehypertrophic HZ = Hypertrophic

N.S.

N.S.

N.S.

N.S.

Page 104: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

92

4.10 Overactivation of MET signaling impairs fracture healing

The lack of a skeletal phenotype despite a reduction in osteoblast differentiation ability

in Met Δ15-HET mice led us to believe the phenotype seen in OFD patients may be

stress induced. We then investigated the fracture healing ability of WT and Met Δ15-HET

12-week-old mice. Since our Met Δ15-HET mouse model do not develop spontaneous

fractures, here we utilized a semi-stabilized tibial fracture induction model which has

been shown to heal through a combination of endochondral and intramembranous

ossification (Y. Chen et al., 2007). Mice were sacrificed 14-days post-fracture for

histological analysis to evaluate fracture healing. This time-point was chosen as it is the

mid-way point of fracture repair where multiple processes (intramembranous

ossification, endochondral ossification, and callus remodelling) involving osteoblasts can

be examined.

At 14-days post-fracture we found significantly more unmineralized cartilage at the

callus site, stained by Safranin-O in red, in Met Δ15-HET mice (Fig 4.10 A). Accordingly,

histomorphometric analysis showed Met Δ15-HET mice fracture sites had significantly

less bone volume/total callous tissue volume (BV/TV) and more osteoid volume/bone

volume (OV/BV) in comparison to their timepoint matched WT counterparts (Fig 4.10 B

& C). This indicates Met Δ15-HET possessed less mineralized bone and more

unmineralized tissue (osteoid) at the fracture site, alluding to a delay in fracture healing.

We also see significantly less tartrate resistant acid phosphatase (TRAP) positive

staining, the histochemical marker of osteoclasts, in Met Δ15-HET mice 14-days post

fracture (Fig 4.10 D) (Minkin, 1982). This was confirmed through colony counting

quantification of the TRAP positive osteoclasts, indicating Met Δ15-HET mice had

significantly reduced number of osteoclasts per defined region of interest (Fig 4.10 E).

Overall, these results indicate Met Δ15-HET mice are at an earlier stage in the fracture

repair timeline in comparison to WT mice 14-days post-fracture.

Page 105: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

93

Safranin-O staining for Cartilage (red)

A

Figure 4.10 Aberrant MET signaling impairs fracture healing in 12-week old mice Representative images of tibial fractures of WT and Met Δ15-HET mice stained with

Safranin-O and fast green to visualize cartilage and bone respectively. Images display the region of interest at the fracture callus. N = 3 mice per group

WT Met Δ15-HET

Page 106: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

94

0

10

20

30

40

WT HET

Perc

en

tag

e

Bone Volume/Tissue Volume [%]

* B

Figure 4.10 Contd.

Quantification of the relative proportion of mineralized bone cartilage and total callous tissue volume at the fracture site using histomorphometric analysis. Student’s two-tailed t test was performed for statistical analysis comparing WT to

Met Δ15-HET measurements, with significance level set at *P<0.05. N=3 mice per

group.

Page 107: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

95

*

0

3

6

9

12

15

WT HET

Perc

en

tag

e

Osteoid Volume/Bone Volume [%]C

Figure 4.10 Contd. Quantification of the relative proportion of unmineralized tissue (osteoid) and total

mineralized bone tissue at the fracture site using histomorphometric analysis. Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET measurements, with significance level set at *P<0.05. N=3 mice per

group.

*

Page 108: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

96

WT Met Δ15-HET D

Figure 4.10 Contd. Representative images of WT and Met Δ15-HET tibial fracture callous sites stained

for tartrate-resistant acid phosphatase (TRAP) positive cells (osteoclasts). Arrows indicate TRAP positive cells. N = 3 mice per group.

Page 109: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

97

0

5

10

15

20

25

30

35

40

WT HET

Nu

mb

er

of

Oste

ocla

sts

/Reg

ion

of

Inte

rest

* E

Figure 4.10 Contd.

Quantification of the number of osteoclasts per region of interest. Student’s two-tailed t test was performed for statistical analysis comparing WT to Met Δ15-HET

measurements, with significance level set at *P<0.05. N=3 mice per group.

Page 110: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

98

5 Discussion

Fracture healing is a highly regulated and complex process. Dysregulation at any one of

the many steps may result in delayed healing and predispose to the non-union seen in

osteofibrous dysplasia. Fracture healing is dependent on osteoblasts to mineralize the

cartilaginous matrix and restore bone function and structural integrity (Marsell &

Einhorn, 2011). Therefore, we believe that a defect in osteoblast differentiation, as seen

in MetΔ15-HET mice, results in delayed fracture healing. Determining and defining the

mechanisms underlying this process in osteofibrous dysplasia may lead to the discovery

of novel therapeutic targets for non-invasive therapies to augment fracture repair.

5.1 MetΔ15-HET mice exhibit higher MET protein levels and upregulated MET signaling

In this project, we used a mouse model specifically created to mimic one of the human

OFD mutations which results in a MET gain-of-function mutation. We first determined if

the model exhibited a similar dysregulation in signalling pathways previously observed

in osteofibrous dysplasia. To accomplish this, we tested MET and its immediate

downstream effectors’ activity at the protein level using MEFs. MetΔ15-HET MEFs

exhibited elevated basal phosphorylation activity at MET catalytic phosphorylation sites

and its downstream effectors: AKT, mTOR, MEK1/2 and ERK1/2 in comparison to WT

MEFs (Fig 4.1A – C). Higher levels of basal signaling activity at these proteins was

consistent with what was previously identified in OFD patients and previously reported

by Gray et al. (2015) ; they demonstrated the competency of the mutant OFD MET

receptor to initiate phosphorylation and downstream signal transduction despite the loss

of the juxtamembrane domain. We also observed higher MET protein levels in mutant

MEFs, but no changes in Met gene expression between WT and MetΔ15-HET MEFs (Fig

4.3 B). Together this data indicates the higher levels of MET protein may be secondary

to diminished receptor internalization and degradation since Met gene expression was

unchanged. Since our results demonstrate similar changes in MET and its downstream

effectors as what was previously reported in human OFD patients, we were satisfied

with using this model for our other proposed experiments.

Page 111: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

99

Next, we determined if the MetΔ15-HET mutation was ligand dependent by following

receptor activation after stimulation with its one and only ligand, HGF. Since the half-life

of HGF is approximately 3 – 5 minutes, peak signalling activity should be observed at

the 5-minute time-point (Chang et al., 2016). When the MET receptor was exogenously

activated, we observed a greater upregulation in catalytic phosphorylation activity at

MET and its downstream effectors (RAS, AKT, mTOR, ERK and β-catenin) 5-minutes

post HGF addition in mutant MEFs versus WT (Fig 4.2A – B). This activity level was

also maintained for a longer duration in MetΔ15-HET MEFs. There was a persistence of

phosphorylation activity 30-minutes post-addition of HGF by which time WT MEFs had

returned to basal activity levels. The reduction in WT MEF signaling activity 30-minutes

post-HGF is consistent with the expected degradation of the MET receptor and signal

termination post receptor activation. Together, these two results suggest the

upregulation of the signaling is due to a failure to internalize and degrade the receptor

post-activation, and consistent with a deletion of the ubiquitin ligase CBL binding site.

The MetΔ15-HET mutation appears to be ligand dependent as increasing HGF

concentrations above basal levels resulted in upregulation of signaling activity. This

observation is consistent with what has been reported by other groups who describe

mutant MET-mediated transformations to be ligand-dependent and inhibited by HGF

antagonists (Michieli et al., 1999). Overall, these results indicate the MetΔ15-HET

mutation is a gain-of-function mutation which is ligand dependent. Our data is also

consistent with the MetΔ15-HET receptor exhibiting reduced degradation.

Mutant MEFs also exhibited higher protein levels of downstream MET effectors: AKT,

mTOR, MEK, ERK1/2 and RAS, replicating what was seen in human OFD lesional

tissue (Fig 1A). In addition, RT-PCR analysis confirmed upregulation of AKT and

ERK1/2 expression in METΔ15-HET MEFs suggesting modulation at the gene

expression level. We have shown that a key osteoblast transcriptional factor, Runx2, is

upregulated in MetΔ15-HET MEFs. Others have shown that an upregulation of Runx2

expression can increase PI3K, AKT and mTOR expression (Cohen-Solal, Boregowda, &

Lasfar, 2015; Fujita et al., 2004). We speculate that Runx2 expression may be changing

AKT and ERK1/2 expression in our animal model but agree that the mechanism behind

Page 112: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

100

these gene expression changes is not yet well understood and will be a critical point of

interest for future experiments aimed at deciphering the relationship between MET

signaling and downstream effectors.

5.2 Overactivation of MET signaling causes osteoblast differentiation defects

After confirming our mouse model recapitulates the dysregulations in signaling seen in

human OFD patients, we tested if the MetΔ15-HET mutation produces an osteoblast

differentiation defect. Here we compared the osteogenic potential of our WT and MetΔ15-

HET mice utilizing colony forming unit assays. We showed MetΔ15-HET BMSCs have a

reduced ability to form fibroblast colonies in comparison to WT bone marrow (Fig 4.4 B).

CFU-F are used as an approximation of skeletal stem cell concentration in bone marrow

(Kuznetsov, Mankani, Bianco, & Robey, 2009). Thus, this data indicates MetΔ15-HET

mice have reduced skeletal stem cells (a subset of BSMCs) in comparison to their WT

counterparts. We then showed MetΔ15-HET mice formed a reduced number of CFU-ALP

and CFU-OB consistent with a reduced ability for osteoblast differentiation and

mineralization (Fig 4.4 B). This finding was similar to what was described by Gray et al.,

(2015) in OFD patients. The number of bone nodule per number of cultured cells

provides an estimation and measurement of the bone forming capacity of the bone

marrow sample (Aubin, 1998). We began to see the reductions in osteoblast

differentiation capacity at day 14 of differentiation through the CFU-ALP assay, but the

reductions in colony number were much more severe at day 21. This suggests the

MetΔ15-HET mutation must have a greater impact on differentiation between day 14 and

21 than between day 1 and 14 of the differentiation timeline. Osteoblasts undergo

maturation between day 14 and 21, suggesting the MetΔ15-HET mutation may be

targeting its effects at the maturation stage of the differentiation process.

5.3 MetΔ15-HET MEFs exhibit dysregulation of β-catenin activity

Through pathway analysis at the protein and gene expression level, we found

dysregulation in multiple pathways leading to the upregulation of β-catenin and Runx2

expression levels in MetΔ15-HET MEFs.

Page 113: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

101

MetΔ15-HET MEFs exhibited higher levels of phosphorylated ERK1/2, indicative of

increased ERK1/2 signaling activity (Fig 4.1C & 4.2 B). ERK1/2 signaling can

phosphorylate LDL-related protein 6 (LRP6), a coreceptor of WNT proteins and key

positive regulator of Wnt/β-catenin signaling. This phosphorylation event enhances the

cellular response to WNT proteins, upregulating Wnt/β-catenin signaling (Cervenka et

al., 2010). We also saw upregulated AKT signaling and corresponding phosphorylation

of the Ser552 residue in β-catenin (Fig 4.2 A & B). Phosphorylation at this residue is

AKT dependent and promotes β-catenin stability and translocation to the nucleus where

it regulates downstream target gene expression (Raucci et al., 2008). Increased AKT

signaling also upregulates downstream pathways such as mTOR, which acts through

the mTORC1 complex to directly upregulate the key osteoblast transcription factor

Runx2 (Dai et al., 2017) (Fig 4.2 A). We also identified a down-regulation in Axin2

expression. Axin2 is required for GSK-3β to phosphorylate β-catenin and target it for

degradation (Jho et al., 2002). Thus, we expect a reduction in Axin2 to increase the

stability of β-catenin as was shown in our data.

In conclusion, we believe that the combination of events affecting ERK1/2 and AKT

signaling along with dysregulations in gene expression of regulators of β-catenin

stability result in the constitutively higher β-catenin levels observed in MetΔ15-HET mice.

Osteoblast differentiation requires tight regulation of β-catenin to achieve terminal

differentiation (Ghadakzadeh et al., 2016). Thus, changes to the spatial and temporal

control of β-catenin levels may result in the osteoblast differentiation seen in MetΔ15-

HET osteoblasts.

5.4 Overactivation of MET signaling does not affect postnatal skeletal development

MetΔ15-HET mice appear to be phenotypically normal in our breeding facility but have

not been thoroughly examined previously. Therefore, we determined their skeletal

phenotype using standard skeletal preparations. Despite seeing reductions in osteoblast

differentiation potential in vitro, MetΔ15-HET mice exhibited no gross skeletal

Page 114: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

102

abnormalities in comparison to WT at 5-days old and 21-days old. We did not identify

any changes in long bones lengths or abnormalities of the cranial skeleton (formed

through endochondral and intramembranous ossification respectively) (Fig 4.7 E & F,

4.8 E & F). In addition, there were no differences in tibial epiphyseal growth plate zone

lengths or overall growth plate height between WT and MetΔ15-HET mice. Overall, this

indicates murine skeletal development and growth does not appear to be affected by the

MetΔ15-HET.

In OFD, there is abnormal osteoblast differentiation at the lesional tissue (shown by the

presence of cells positive for both osteoblast and fibroblast cell markers). The healthy

tissue surrounding these lesions and the rest of the skeleton are normal (Gray et al.,

2015). It is plausible that the osteoblast differentiation defect in MetΔ15-HET BMSCs

does not affect normal development but does affect bone repair. Another plausible

explanation for the lack of a skeletal phenotype in mice may be due to their more robust

regenerative ability (Haffner-Luntzer, Kovtun, Rapp, & Ignatius, 2016). Lastly, the

absence of a skeletal phenotype could be because MetΔ15-HET mice still exhibit partial

regulatory control of Met signaling and not a complete block in osteoblast differentiation.

This notion is supported by the embryonic lethality seen in MetΔ15-HOMO mice, which

may be caused by the inability to form an organized cartilaginous skeletal precursor

during embryogenesis (Ben-Zvi, Shilo, & Barkai, 2011).

5.5 MetΔ15-HET mice experience delayed fracture healing ability

To test whether the in vitro osteoblast differentiation defect in MetΔ15-HET mice impairs

bone healing, we examined fracture healing in vivo. Our MetΔ15-HET mice do not

develop spontaneous fractures and we thus induced fractures using a stabilized mid-

diaphyseal tibia fracture model. In mice fractures are typically healed 28 days post-

fracture, with peak soft callus formation occurring approximately 7 – 9 days post-

fracture, accompanied by a peak in proteoglycan content while peak hard callus

formation is reached 14 days post-fracture (Marsell & Einhorn, 2011).

Page 115: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

103

We decided to first assess the healing process 14-days post fracture, as this timepoint

allows for visualization of intramembranous and endochondral ossification and

remodelling which both require osteoblasts. In comparison to WT mice, MetΔ15-HET

mice still exhibit large amounts of cartilage callus 14-days post fracture, indicated by the

greater amount of Safranin-o staining (Fig 4.10 A). This suggests MetΔ15-HET fracture

sites have yet to transition from a soft to hard boney callus, a process in which

osteoblast are heavily involved. Accordingly, MetΔ15-HET fractures also displayed less

bone volume and more unmineralized matrix (osteoid) volume at the fracture site when

compared to WT mice at the same timepoint (Fig 4.10 B & C). Endochondral ossification

and osteoblasts are critical for the transition from soft callus to hard callus therefore a

defect in osteoblast differentiation would hinder this transition. In summary our data

indicates delayed fracture healing in MetΔ15-HET mice in comparison to their WT

counterparts.

Osteoclasts are intimately involved in the remodeling of the hard callus, a late stage

process in fracture healing, which starts approximately 13 – 14 days post-fracture.

Osteoclasts are typically in abundance at the fracture site 14-days post-fracture.

However, the MetΔ15-HET callus contained 50% less osteoclasts in comparison to WT

counterparts 14-days post fracture (Fig 4.10 D & E). We believe that this is further proof

of delayed fracture repair in the MetΔ15-HET mouse.

There is an intimate relationship between osteoclasts and osteoblasts, in which co-

regulation occurs through RANK-RANKL interactions. As previously mentioned,

osteoblasts secrete RANKL which binds the RANK receptor on osteoclast to facilitate

their maturation. Therefore, a defect in terminal osteoblast differentiation reducing

osteoblast formation will also reduce RANKL expression which is needed to induce

osteoclastogenesis (Katagiri & Takahashi, 2002).

The reduction in osteoclast differentiation could also be attributed to the upregulation of

adrenomedullin expression we previously identified. Along with ADM’s role in the

inhibition of GSK3β, it is also involved in the inhibition of RANK-RANKL interactions (Y.

Liu et al., 2017). Therefore, diminished osteoclast numbers seen in MetΔ15-HET fracture

Page 116: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

104

calluses could be the result of reduced terminally differentiated osteoblasts (available to

produce RANKL) and increased ADM-mediated suppression of RANKL-RANK

interactions (Y. Liu et al., 2017).

5.6 MetΔ15-HET osteoblasts display dysregulation of β-catenin

When examining the gene expression profile of osteoblasts at day 21 of differentiation,

we found a dysregulation of osteoblast specific gene markers in MetΔ15-HETs. We saw

a 22-fold increase in β-catenin along with a 6.4-fold reduction in Axin2 expression, a

critical scaffold protein in the β-catenin degradation complex (Fig 4.5 B & C) (Jho et al.,

2002; Yan et al., 2009). The combination of increased expression of β-catenin and

reduced ability to degrade the β-catenin protein would explain the higher levels of β-

catenin and its phosphorylated version seen in MetΔ15-HETs. The elevated levels of β-

catenin occur at a timepoint at which β-catenin requires tight regulation for normal

osteoblast differentiation. Any deviations from this would result in terminal osteoblast

differentiation defect (Ghadakzadeh et al., 2016).

One consequence of elevated β-catenin levels is an upregulation of downstream target

genes such as Runx2. Here we report MetΔ15-HET osteoblasts exhibit a 27-fold increase

in Runx2 expression at the end of in vitro osteoblast differentiation (day 21) (Fig 4.5 E).

In addition, we saw a 1.3-fold reduction in alkaline phosphatase and 4-fold reduction in

osteocalcin, late stage osteoblastic markers, in MetΔ15-HET osteoblasts at day 21 (Fig

4.5 G & H). This suggests that the MET gain of function mutation signaling causes

inhibition of terminal osteoblast differentiation through dysregulation of β-catenin and

Runx2. This is consistent with what has been reported by multiple groups; the

overexpression of Runx2 during the maturation stages of osteoblast differentiation

prevents terminal differentiation (T. M. Liu & Lee, 2012; Wenguang Liu et al., 2001). In

addition, this also fits with what was described by Gray et al., (2015) who found no

changes in early markers of osteoblast differentiation but observed reductions in late

stage osteoblastic markers when the MET exon 15 skipping mutation was induced in an

osteoblastic cell line. In summary, these results provide a potential mechanism

Page 117: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

105

explaining the observed in vitro osteoblast differentiation defect and the diminished

callus mineralization in MetΔ15-HET mice fractures (Fig 4.4 A & B, Fig 4.10 A – E).

Day 21 MetΔ15-HET osteoblasts were also found to exhibit a 24.6-fold increase in

adrenomedullin (ADM) expression (Fig 4.5 I), which is involved in the inactivation of

GSK3β. Since GSK3β is involved in the degradation of β-catenin, we would expect

higher levels of β-catenin and RAS, which we did indeed observe in MetΔ15-HET MEFs

(Fig 4.2 B) (Jeong et al., 2018). The upregulation of RAS was also accentuated by

downregulation of Rasal3 expression (Fig 4.5 J), an inactivator of RAS signaling (Muro

et al., 2015). The elevated levels of RAS protein and activity promotes increased

stabilization of β-catenin through ERK1/2-dependent and independent pathways (Jeong

et al., 2018).

Through comprehensive protein and gene expression analysis, we were able to

implicate a dysregulation of β-catenin as a potential reason behind the osteoblast

differentiation defect seen in our in vitro and in vivo experiments. Our findings are

consistent with the idea that β-catenin requires tight spatiotemporal regulation during

osteoblast differentiation and fracture healing. In the early stages of fracture repair,

precise regulation of β-catenin is required for mesenchymal stem cell differentiation to

osteoblasts. In the early stages of osteoblast differentiation, β-catenin positively

regulates osteoblasts through upregulation of Runx2 expression (Y. Chen et al., 2007).

In the late maturation stages, Runx2 expression is inhibited and terminal differentiation

to mature osteoblasts ensures. We propose, β-catenin levels are constitutively elevated

in MetΔ15-HET mice, resulting in upregulation of Runx2 during the maturation stages

hindering terminal differentiation.

Interestingly, others have also described a dysregulation of RAS and β-catenin levels in

pseudarthrosis secondary to neurofibromatosis-1 (NF-1) and fibrous dysplasia

(Ghadakzadeh et al., 2016; Regard et al., 2011). These bone diseases present similar

phenotypes to osteofibrous dysplasia and we propose they also share similar causative

mechanisms; through the dysregulation of β-catenin mediated Runx2 expression. Liu et

al., (2001) showed osteoblast specific overexpression of Runx2 in vivo resulted in a

Page 118: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

106

maturational blockage in osteoblasts in vivo; mice exhibited reduced number of

terminally differentiated osteoblasts and osteocalcin expression. We report similar

findings in our mature MetΔ15-HET osteoblasts, which displayed upregulations in Runx2

expression, downregulation of osteocalcin, and a reduced ability to terminally

differentiate. Additionally, when Liu et al., 2001 overexpressed Runx2 in osteoblasts in

vivo, it resulted in osteopenia and the development of fractures. The loss of the

spatiotemporal control of the regulatory relationship between Wnt/β-catenin signaling

and Runx2 expression could be the commonality between diseases presenting with

pseudarthrosis phenotypes.

Accordingly, we believe manipulation of the Wnt/β-catenin signaling pathway could be a

potential therapeutic target in treating OFD patients exhibiting non-healing tibial

fractures. This approach has been previously tested in vivo in treating NF-1 related tibial

fracture. Dickkopf-1 mediated inhibition of Wnt/β-catenin signaling led to the union of

fractures and improved osteoblast differentiation (Ghadakzadeh et al., 2016).

In this study we were able to characterize and validate the use of a MetΔ15-HET genetic

mouse model to replicate the aberrant signaling seen in human osteofibrous dysplasia.

Through pathway analysis at the protein and gene expression level, we were able to

confirm aberrant MET signaling is caused by the failure to degrade the MET receptor

post-activation. We were also able to replicate the osteoblast differentiation defect seen

in human disease in vitro utilizing our MetΔ15-HET mouse model. In addition, we

propose a potential mechanism explaining the osteoblast differentiation defect seen in

OFD. MetΔ15-HET mouse skeletal characterization revealed an absence of gross

defects. When examining the fracture healing ability of WT and MetΔ15-HET mice, we

found MetΔ15-HETs displayed delayed fracture repair in comparison to WT mice.

Overall, these results support the hypothesis that gain-of-function MET mutations result

in diminished bone repair ability due to reduced osteoblast differentiation. We believe

this defect is caused by upregulated signaling at downstream effectors of aberrant MET

signaling resulting in constitutively elevated levels of β-catenin leading to loss of

spatiotemporal regulation of Runx2 expression and inhibition of osteoblast terminal

differentiation.

Page 119: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

107

Patients currently affected by OFD experience severe pain and disabilities due to non-

healing fractures, resulting in gait deformation and punishing reduction in quality of life.

Since the role of MET in bone formation is largely unstudied, we hope our data will pave

the way for further interventions to improve fracture healing in these patients.

5.7 Limitations to the Study

Through in vitro and in vivo experiments, we have been able to implicate MET’s role in

bone development and fracture repair, however, there were limitations in this study

which are discussed here.

Firstly, we used a model where the MET mutation was expressed in all cell types. This

makes it difficult to delineate and define the clear effects of the mutation on a specific

cell type such as osteoblasts. This limitation is especially important when examining

complex processes such as fracture repair which involve numerous cell types and

signaling pathways acting in concert. In the face of this limitation, the MetΔ15-HET

mouse model we have created and utilized does provide compelling evidence for the

role of MET in bone development and fracture repair. The use of an osteoblast targeted

conditional MetΔ15 mouse model would allow for more precise conclusions of the effect

of the osteoblast differentiation defect on fracture healing, it may not allow for us to

examine any broad changes in the fracture repair process.

This leads to the next limitation of this study; we did not extensively explore the role of

osteoclasts but rather focused on osteoblasts function. The harmonic relationship

between osteoblasts and osteoclasts has been well defined; they are known to co-

regulate and communicate through RANK-RANKL interactions. In addition, MET and

HGF also are involved in reciprocal signaling between osteoblasts and osteoclasts.

Osteoclast function may be impaired by the MetΔ15-HET mutation independent of

RANK-RANKL interactions. Investigation into the effects of the MetΔ15-HET mutation on

osteoclast differentiation and function may provide alternative explanations and

Page 120: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

108

mechanisms for why we see reduced osteoclast number at the fracture site 14-days

post-fracture.

Various animals models with enhanced angiogenesis have shown improved bone

regeneration abilities post-fractures, making modulation of vascularization post-fracture

a viable therapeutic approach in improving fracture outcomes (Hankenson et al., 2011).

We did not examine the effects of the MET mutation on vascularization, a key

component of the fracture repair process (Hankenson et al., 2011). Vascular invasion of

the fracture site supplies oxygen, nutrients and inflammatory cells to the fracture site

and contributes to the formation of a hematoma which acts as the template for the

formation of a vascular callus. Failure to vascularize the fracture site has been shown to

result in diminished repair and greater incidence of non-unions (Dickson, Katzman,

Delgado, & Conteras, 1994; Hankenson et al., 2011). This crucial process is regulated

by numerous growth factors, but most notably by VEGF signaling which has been

shown to be regulated by MET’s downstream pathways: PI3K/AKT, MAPK and STAT3

(Matsumura et al., 2013). Therefore, it is highly plausible the MET mutation may affect

vascularization during the fracture repair process.

Many of the pathway analysis experiments utilized MEFs instead of osteoblasts;

although the results may be representative of what is seen in osteoblasts, there still may

be subtle differences which cannot be ignored. This is despite the fact they share many

similarities such as both being a part of the connective-tissue cell family, secrete type-1

collagen and are responsible for the structural framework of the body (Mackie, Ahmed,

Tatarczuch, Chen, & Mirams, 2008). Osteoblast differentiation is a dynamic process

with multiple key stages such as the commitment of MSCs to pre-osteoblasts and

maturation of immature osteoblasts. These stages involve distinctive regulation of key

genes and signaling pathways therefore utilizing MEFs results in loss of temporal

information and cannot capture the dynamic properties of the process. Nonetheless,

MEFs allow for a broad overview of the potential differences in signalling and

expression between WT and MetΔ15-HET mutants, and therefore still provide valuable

knowledge on the effects of aberrant MET signaling. In addition, the technical benefits

and advantages of using fibroblasts due to their resilient characteristics and ability to

Page 121: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

109

thrive, making them one of the easiest cells to grow in cell culture, were why they were

chosen in favour of osteoblasts (Alberts et al., 2002).

Page 122: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

110

5.8 Proposed Mechanism for the Dysregulation of β-catenin Signaling in Osteoblast Differentiation

During the maturation stages of osteoblast differentiation:

1) Loss of juxtamembrane domain results in diminished degradation of the MET

receptor causing upregulation of signaling

2) Upregulation of MET signaling upregulates downstream pathways: AKT-mTOR

and RAS-MEK-ERK1/2 through upregulated phosphorylation events

3) Aberrant MET signaling results in upregulated ADM mediated inhibition of

GSK3β activity and downregulation of RASAL3 mediated inactivation of RAS

activity

Overall Effect: inappropriate elevated expression of Runx2 at a timepoint at which

Runx2 inhibits terminal osteoblast maturation resulting in a differentiation defect.

I. Upregulated AKT signaling results in

upregulation of phosphorylation of β-catenin

at Ser552, enhancing its stability and

promoting translocation to the nucleus to

upregulate Runx2 expression through

interactions with LEF/TCF transcription factors

II. Upregulated AKT activity increases mTOR

activity

III. Upregulated mTOR signaling directly

upregulates Runx2 expression through the

mTORC1 complex

I. At the same time, upregulation of RAS

signaling enhances β-catenin stability and

translocation to the nucleus resulting

upregulation of Runx2

II. Upregulated RAS activity increases ERK1/2

activity

III. Upregulated ERK1/2 signaling activity, can

result in upregulation of LRP6 phosphorylation

leading to inhibition of β-catenin degradation

complex activity. Therefore β-catenin stability

and translocation to the nucleus is enhanced

resulting in upregulation of Runx2

Pathway 1 Pathway 2

Page 123: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

111

Figure 5.1. Proposed Mechanism for the Dysregulation of β-catenin Signaling in Osteoblast Differentiation

Page 124: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

112

6 Conclusions

We initially hypothesized that gain-of-function MET mutations result in diminished bone

repair due to reduced osteoblast differentiation. To test this hypothesis, we 1)

determined MET pathway irregularities between WT and MetΔ15-HET mice, 2)

characterized the skeletal phenotype of WT and MetΔ15-HET, and 3) compared fracture

healing between WT and MetΔ15-HET mice in vivo.

We first validated the pathway signaling profile of our MetΔ15-HET mouse model to

ensure it captured the same dysregulations in signaling seen in OFD patients. MetΔ15-

HET MEFs exhibit upregulated protein levels of MET, its downstream effectors and their

phosphorylated versions in comparison to WT MEFs (Fig 4.1A – C). The absence of an

upregulation in MET gene expression suggests the elevated protein levels is due to a

failure in degradation of the receptor, while elevated levels of downstream effectors

indicated elevated basal signaling activity.

When HGF stimulation was applied, MetΔ15-HET MEFs exhibited higher upregulation of

signaling activity and longer sustained signaling activity in comparison to WT MEFs.

This is indicative of a diminished ability for MetΔ15-HETs to degrade the MET receptor

and terminate signaling activity. Overall these results confirm the MetΔ15-HET mouse

model replicates the dysregulations in signalling seen in humans as described by Gray

et al., (2015).

Next, we examined osteoblast differentiation through CFU assays of primary BMSCs

induced to differentiate osteoblasts. Here we found MetΔ15-HET BMSCs displayed a

reduced ability for differentiation and mineralization in comparison to WT counterparts,

forming 35% less CFU-ALP colonies and 80% less CFU-OBs at day 21 (Fig 4.4B). This

demonstrates an osteoblast differentiation defect in MetΔ15-HET BMSCs.

We then examined if there were any differences in gene expression profiles between

WT and MetΔ15-HET mutants. We found multiple pathways and factors with

dysregulated signaling and expression resulting in the upregulation of β-catenin levels

and Runx2 expression during the late stages of osteoblast differentiation (osteoblast

maturation). These two factors require strict spatiotemporal regulation of expression and

Page 125: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

113

activity in order to achieve terminal osteoblast differentiation. β-catenin mediated

upregulation of Runx2 expression during a time (day 21 of osteoblast differentiation) at

which Runx2 inhibits the progression of osteoblast differentiation leads us to believe

dysregulation of β-catenin signaling activity is the cause of the differentiation defect

seen. This conclusion is supported by the literature, as diseases with similar bone

phenotypes have also been shown to be caused by dysregulations in spatiotemporal

regulation of β-catenin (Ghadakzadeh et al., 2016).

Since we found dysregulation of multiple signaling pathways and an osteoblast

differentiation defect in MetΔ15-HET mutants, we examined if these mutant mice

exhibited any gross skeletal phenotypes. Through whole skeletal prep and epiphyseal

growth plate analysis, we found there were no differences in intramembranous and

endochondral bone size (Fig 4.7 & 4.8). These results suggest perhaps the defect seen

in humans is a stress induced defect caused by fractures. This fits in line with the

characterization of OFD in which osteoblast differentiation defects are only seen in

lesional tissue but not the surrounding healthy tissue resulting in a distinctive zonal

architecture.

We then moved on to examination of fracture healing in vivo and found 12-week old

male MetΔ15-HET mice exhibited delayed fracture healing in comparison to WT

counterparts at 14-days post-fracture. We saw increased cartilage at the callus site

during a timepoint at which there should be minimal, as the soft callus should have been

converted to hard callus. Furthermore, we saw minimal osteoclast number at the callus

in MetΔ15-HET mice confirming the remodeling process has yet to begin, while WT

calluses were abundant in osteoclasts already. This points to MetΔ15-HET mice being at

a less advanced stage of fracture repair due to a delay in the mineralization of the hard

callus, a process in which osteoblasts are intimately involved in. This leads us to

conclude overactivation of MET signaling delays fracture healing due the inability for

osteoblasts to terminally differentiate and mineralize the soft callus.

Overall, this study demonstrates overactive MET signaling results in stabilization of β-

catenin and upregulated Runx2 expression causing loss of the tight spatiotemporal

regulation required for terminal osteoblast differentiation. The inability for osteoblasts to

Page 126: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

114

terminally differentiate results in the delays in fracture healing seen in our mouse model

and OFD patients. Further interventional studies against the MET and β-catenin

signaling pathway may result in the development of novel therapeutic strategies against

OFD and similar diseases before resorting to invasive surgical means.

Page 127: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

115

7 Future Directions

Osteofibrous Dysplasia (OFD) is a rare benign non-neoplastic condition, caused by

mutations in the MET gene, in which patients experience reduced fracture healing and

osteoblast differentiation ability. The MET receptor is a regulator of numerous

proliferation and differentiation pathways critical for osteoblast differentiation. In this

study, we demonstrate the loss of spatiotemporal control of Runx2 expression due to

dysregulation of β-catenin levels results in an osteoblast differentiation defect and

delayed fracture healing. This further supports and reiterates the need for tight control of

β-catenin levels during osteoblast differentiation and coordinated expression of Runx2

depending on the stage of osteoblast differentiation (Ghadakzadeh et al., 2016).

To better understand the dysregulated pathways in osteoblast differentiation some

future directions would include pathway analysis of MET and its downstream effectors

throughout the differentiation process. These results may reveal the exact stage of

osteoblast differentiation the dysregulation of β-catenin occurs at. This would be done

by examining protein and gene expression levels at the following timepoints: day 2 of

osteoblast differentiation (commitment to osteoblast differentiation), day 7 (pre-

osteoblast to immature osteoblast transition), day 14 (maturation of immature osteoblast

begins) and day 21 (terminally differentiated). The results of these experiments would

provide a comprehensive timeline of the modulations and dysregulations which occur

due to the MetΔ15-HET mutation.

To further confirm the role of MET signaling in osteoblast differentiation, rescue

experiments would need to be performed using selective MET inhibitors such as ARQ

197 (Jeay et al., 2007). MET inhibitors would be applied during the osteoblast

differentiation process to determine if the defect seen in Met Δ15-HET mice can be

improved. This would be measured by colony forming unit assays, protein expression,

and gene expression analysis of osteoblastic markers such as Runx2, ALP and

osteocalcin. A successful rescue would show improvements in CFU-ALP and CFU-OB

number and restoration of the aforementioned osteoblast differentiation markers to WT

levels. The ligand-dependency of aberrant MET signaling could be further examined by

treatment of Met Δ15-HET MEFs with NK4. NK4 is a HGF specific antagonist which

Page 128: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

116

competitively inhibits the binding of HGF to the MET receptor and prevents

transphosphorylation post-activation (Date, Matsumoto, Shimura, Tanaka, & Nakamura,

1997; Du et al., 2007). Reduction of the elevated levels of phosphorylated MET and its

downstream effectors seen in Met Δ15-HET MEFs, would reconfirm the ligand-dependent

nature of this mutation.

Fracture repair is a complex process involving a delicate balance between osteoblast

and osteoclast activity. Though the main focus of this study was focused on the role of

osteoblasts, the role of osteoclasts cannot be ignored. In this study we demonstrate the

MET exon 15 skipping mutation may result in an osteoclast differentiation defect as well

with a reduction in osteoclast number at the callus site 14-days post fracture in Met Δ15-

HET mice. This is not so far-fetched as the primary method of osteoclastogenesis

induction is through the communication of osteoblast and osteoclasts through RANK-

RANKL interactions (Park-Min, 2018). In addition, MET and HGF participate in

reciprocal signaling between osteoblasts and osteoclasts as well, therefore it is not

unlikely osteoclast function and differentiation may also be impacted by this mutation.

This would be done by performing the pathway analysis experiments highlighted in this

study utilizing BMSC induced osteoclast cultures (Bradley & Oursler, 2008). In addition,

a bone resorption functional assay would be performed to examine any potential

differences in osteoclast function. The results of these experiments could also provide

alternative explanations and mechanisms explaining the reduction in osteoclast number

seen at the callus site of Met Δ15-HET mice 14-days post fracture.

Lastly, a crucial and necessary step in this line of research would be to transition from

mouse and animal model experiments to human clinical trials. The transability and

practicality of targeting some of these elucidated pathway effectors would need to be

examined as well. Only then would we be closer to developing novel therapeutic

treatment plans against reduced fracture healing in osteofibrous dysplasia patients.

Page 129: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

117

8 Supplementary Results 8.1 Confirmation of Exon 15 Skipping Event

The occurrence of the Exon 15 skipping event was confirmed in our Met Δ15-HET mouse

model utilizing PCR primers targeting Exon 14 and Exon 16 followed by PCR

amplification. Therefore, in complete Met transcripts (with Exon 15), Exon 14 – 16 would

be amplified by PCR, resulting in the formation of a 469 bp product. The exclusion of

Exon 15 would result in the formation of a 319 bp PCR product. Through gel

electrophoresis mediated resolution, we were able to confirm the exon skipping event by

differentiating PCR product sizes. Here we found WT osteoblasts to exhibit both

complete Met transcripts along with a small proportion of Met transcripts excluding Exon

15 as well. This coincides with what was reported by Gray et al., (2015) who found the

Met exon 15 skipping event to occur during development in mice. In addition, as

expected we see MetΔ15-HET osteoblasts exhibiting both complete Met and MetΔ15

transcripts confirming the exon skipping event in our genetic mouse model.

WT MetΔ15-HET

100

200

300

400

500

Page 130: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

118

9 References

Abella, J. V., Peschard, P., Naujokas, M. A., Lin, T., Saucier, C., Urbe, S., & Park, M.

(2005). Met/Hepatocyte Growth Factor Receptor Ubiquitination Suppresses Transformation and Is Required for Hrs Phosphorylation. Molecular and Cellular Biology. https://doi.org/10.1128/mcb.25.21.9632-9645.2005

Aenlle, K. K., Curtis, K. M., Roos, B. A., & Howard, G. A. (2014). Hepatocyte Growth Factor and p38 Promote Osteogenic Differentiation of Human Mesenchymal Stem Cells. Molecular Endocrinology. https://doi.org/10.1210/me.2013-1286

Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., & Walter, P. (2002). Fibroblasts and Their Transformations: The Connective-Tissue Cell Family. In Molecular Biology of the Cell.

Aubin, J. E. (1998). Advances in the osteoblast lineage. Biochemistry and Cell Biology =

Biochimie et Biologie Cellulaire.

Aubin, J. E., Liu, F., Malaval, L., & Gupta, A. K. (1995). Osteoblast and chondroblast differentiation. Bone. https://doi.org/10.1016/8756-3282(95)00183-E

Baldanzi, G., & Graziani, A. (2014). Physiological Signaling and Structure of the HGF Receptor MET. Biomedicines. https://doi.org/10.3390/biomedicines3010001

Beane, O. S., Fonseca, V. C., Cooper, L. L., Koren, G., & Darling, E. M. (2014). Impact of aging on the regenerative properties of bone marrow-, muscle-, and adipose-derived mesenchymal stem/stromal cells. PLoS ONE.

https://doi.org/10.1371/journal.pone.0115963

Bellido, T. (2014). Osteocyte-driven bone remodeling. Calcified Tissue International.

https://doi.org/10.1007/s00223-013-9774-y

Ben-Zvi, D., Shilo, B. Z., & Barkai, N. (2011). Scaling of morphogen gradients. Current Opinion in Genetics and Development. https://doi.org/10.1016/j.gde.2011.07.011

Bhowmick, N. A., Neilson, E. G., & Moses, H. L. (2004). Stromal fibroblasts in cancer initiation and progression. Nature. https://doi.org/10.1038/nature03096

Bialek, P., Kern, B., Yang, X., Schrock, M., Sosic, D., Hong, N., … Karsenty, G. (2004). A twist code determines the onset of osteoblast differentiation. Developmental Cell.

Bianco, P., & Robey, P. G. (2000). Marrow stromal stem cells. Journal of Clinical

Investigation. https://doi.org/10.1172/JCI10413

Birchmeier, C., & Gherardi, E. (1998). Developmental roles of HGF/SF and its receptor, the c-met tyrosine kinase. Trends in Cell Biology. https://doi.org/10.1016/S0962-

8924(98)01359-2

Boon, E. M. J., Van der Neut, R., Van de Wetering, M., Clevers, H., & Pals, S. T. (2002). Wnt signaling regulates expression of the receptor tyrosine kinase Met in colorectal cancer. Cancer Research.

Bradley, E. W., & Oursler, M. J. (2008). Osteoclast culture and resorption assays. Methods in Molecular Biology. https://doi.org/10.1007/978-1-59745-104-8_2

Camplejohn, K. L., & Allard, S. A. (1988). Limitations of safranin “O” staining in

proteoglycan-depleted cartilage demonstrated with monoclonal antibodies.

Page 131: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

119

Histochemistry. https://doi.org/10.1007/BF00489922

Canalis, E., Varghese, S., McCarthy, T. L., & Centrella, M. (1992). Role of platelet derived growth factor in bone cell function. Growth Regulation.

Capulli, M., Paone, R., & Rucci, N. (2014). Osteoblast and osteocyte: Games without frontiers. Archives of Biochemistry and Biophysics.

https://doi.org/10.1016/j.abb.2014.05.003

Cervenka, I., Wolf, J., Masek, J., Krejci, P., Wilcox, W. R., Kozubik, A., … Bryja, V. (2010). Mitogen-Activated Protein Kinases Promote WNT/ -Catenin Signaling via Phosphorylation of LRP6. Molecular and Cellular Biology.

https://doi.org/10.1128/mcb.00550-10

Chan, J. K. K., Harry, L., Williams, G., & Nanchahal, J. (2012). Soft-tissue reconstruction of open fractures of the lower limb: Muscle versus fasciocutaneous flaps. Plastic and Reconstructive Surgery.

https://doi.org/10.1097/PRS.0b013e3182589e63

Chang, H. K., Kim, P. H., Cho, H. M., Yum, S. Y., Choi, Y. J., Son, Y. S., … Cho, J. Y. (2016). Inducible HGF-secreting human umbilical cord blood-derived MSCs produced via TALEN-mediated genome editing promoted angiogenesis. Molecular Therapy. https://doi.org/10.1038/mt.2016.120

Chen, H. Te, Tsou, H. K., Chang, C. H., & Tang, C. H. (2012). Hepatocyte growth factor

increases osteopontin expression in human osteoblasts through PI3k, akt, c-src, and AP-1 signaling pathway. PLoS ONE.

https://doi.org/10.1371/journal.pone.0038378

Chen, Y., Whetstone, H. C., Lin, A. C., Nadesan, P., Wei, Q., Poon, R., & Alman, B. A.

(2007). Beta-catenin signaling plays a disparate role in different phases of fracture repair: Implications for therapy to improve bone healing. PLoS Medicine.

https://doi.org/10.1371/journal.pmed.0040249

Christensen, J. G., Burrows, J., & Salgia, R. (2005). c-Met as a target for human cancer and characterization of inhibitors for therapeutic intervention. Cancer Letters.

https://doi.org/10.1016/j.canlet.2004.09.044

Cohen-Solal, K. A., Boregowda, R. K., & Lasfar, A. (2015). RUNX2 and the PI3K/AKT axis reciprocal activation as a driving force for tumor progression. Molecular Cancer. https://doi.org/10.1186/s12943-015-0404-3

Comoglio, P. M., & Boccaccio, C. (2001). Scatter factors and invasive growth. Seminars

in Cancer Biology. https://doi.org/10.1006/scbi.2000.0366

Corrado, A., Sanpaolo, E. R., Di Bello, S., & Cantatore, F. P. (2017). Osteoblast as a target of anti-osteoporotic treatment. Postgraduate Medicine.

https://doi.org/10.1080/00325481.2017.1362312

Dai, Q., Xu, Z., Ma, X., Niu, N., Zhou, S., Xie, F., … Zou, W. (2017). MTOR/Raptor signaling is critical for skeletogenesis in mice through the regulation of Runx2 expression. Cell Death and Differentiation. https://doi.org/10.1038/cdd.2017.110

Date, K., Matsumoto, K., Shimura, H., Tanaka, M., & Nakamura, T. (1997). HGF/NK4 is a specific antagonist for pleiotrophic actions of hepatocyte growth factore. FEBS Letters. https://doi.org/10.1016/S0014-5793(97)01475-0

Page 132: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

120

Day, T. F., Guo, X., Garrett-Beal, L., & Yang, Y. (2005). Wnt/β-catenin signaling in mesenchymal progenitors controls osteoblast and chondrocyte differentiation during vertebrate skeletogenesis. Developmental Cell.

https://doi.org/10.1016/j.devcel.2005.03.016

Dickson, K., Katzman, S., Delgado, E., & Conteras, D. (1994). Delayed Unions and Nonunions of Open Tibial Fractures Correlation With Arteriography Results. Clinical Orthopaedics and Related Research.

Du, W., Hattori, Y., Yamada, T., Matsumoto, K., Nakamura, T., Sagawa, M., … Ikeda, Y. (2007). NK4, an antagonist of hepatocyte growth factor (HGF), inhibits growth of multiple myeloma cells: Molecular targeting of angiogenic growth factor. Blood.

https://doi.org/10.1182/blood-2006-02-003103

Farrell, E., Both, S. K., Odörfer, K. I., Koevoet, W., Kops, N., O’Brien, F. J., … Van

Osch, G. J. V. M. (2011). In-vivo generation of bone via endochondral ossification by in-vitro chondrogenic priming of adult human and rat mesenchymal stem cells. BMC Musculoskeletal Disorders. https://doi.org/10.1186/1471-2474-12-31

Fioramonti, M., Santini, D., Iuliani, M., Ribelli, G., Manca, P., Papapietro, N., … Pantano, F. (2017). Cabozantinib targets bone microenvironment modulating human osteoclast and osteoblast functions. Oncotarget.

https://doi.org/10.18632/oncotarget.15390

Foveau, B., Ancot, F., Leroy, C., Petrelli, A., Reiss, K., Vingtdeux, V., … Tulasne, D. (2009). Down-Regulation of the Met Receptor Tyrosine Kinase by Presenilin-dependent Regulated Intramembrane Proteolysis. Molecular Biology of the Cell.

https://doi.org/10.1091/mbc.e08-09-0969

Frampton, G. M., Ali, S. M., Rosenzweig, M., Chmielecki, J., Lu, X., Bauer, T. M., … Miller, V. A. (2015). Activation of MET via diverse exon 14 splicing alterations

occurs in multiple tumor types and confers clinical sensitivity to MET inhibitors. Cancer Discovery. https://doi.org/10.1158/2159-8290.CD-15-0285

Fujita, T., Azuma, Y., Fukuyama, R., Hattori, Y., Yoshida, C., Koida, M., … Komori, T. (2004). Runx2 induces osteoblast and chondrocyte differentiation and enhances their migration by coupling with PI3K-Akt signaling. Journal of Cell Biology.

https://doi.org/10.1083/jcb.200401138

Furge, K. A., Zhang, Y. W., & Vande Woude, G. F. (2000). Met receptor tyrosine kinase: Enhanced signaling through adapter proteins. Oncogene.

https://doi.org/10.1038/sj.onc.1203859

Garg, P., Mazur, M. M., Buck, A. C., Wandtke, M. E., Liu, J., & Ebraheim, N. A. (2017).

Prospective Review of Mesenchymal Stem Cells Differentiation into Osteoblasts. Orthopaedic Surgery. https://doi.org/10.1111/os.12304

Gaur, T., Lengner, C. J., Hovhannisyan, H., Bhat, R. A., Bodine, P. V. N., Komm, B. S., … Lian, J. B. (2005). Canonical WNT signaling promotes osteogenesis by directly stimulating Runx2 gene expression. Journal of Biological Chemistry.

https://doi.org/10.1074/jbc.M500608200

Ghadakzadeh, S., Kannu, P., Whetstone, H., Howard, A., & Alman, B. A. (2016). β-

Catenin modulation in neurofibromatosis type 1 bone repair: Therapeutic implications. FASEB Journal. https://doi.org/10.1096/fj.201500190RR

Page 133: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

121

Ghosh-Choudhury, N., Mandal, C. C., & Choudhury, G. G. (2007). Statin-induced Ras activation integrates the phosphatidylinositol 3-kinase signal to Akt and MAPK for bone morphogenetic protein-2 expression in osteoblast differentiation. Journal of Biological Chemistry. https://doi.org/10.1074/jbc.M606706200

Gleason, B. C., Liegl-Atzwanger, B., Kozakewich, H. P., Connolly, S., Gebhardt, M. C.,

Fletcher, J. A., & Perez-Atayde, A. R. (2008). Osteofibrous dysplasia and adamantinoma in children and adolescents: A clinicopathologic reappraisal. American Journal of Surgical Pathology.

https://doi.org/10.1097/PAS.0b013e318150d53e

Goldring, S. R. (2015). The osteocyte: Key player in regulating bone turnover. RMD Open. https://doi.org/10.1136/rmdopen-2015-000049

Grano, M., Galimi, F., Zambonin, G., Colucci, S., Cottone, E., Zallone, A. Z., &

Comoglio, P. M. (2002). Hepatocyte growth factor is a coupling factor for osteoclasts and osteoblasts in vitro. Proceedings of the National Academy of Sciences. https://doi.org/10.1073/pnas.93.15.7644

Gray, M. J., Kannu, P., Sharma, S., Neyt, C., Zhang, D., Paria, N., … Robertson, S. P. (2015). Mutations Preventing Regulated Exon Skipping in MET Cause Osteofibrous Dysplasia. American Journal of Human Genetics.

https://doi.org/10.1016/j.ajhg.2015.11.001

Greenblatt, M. B., Shim, J. H., Zou, W., Sitara, D., Schweitzer, M., Hu, D., … Glimcher, L. H. (2010). The p38 MAPK pathway is essential for skeletogenesis and bone homeostasis in mice. Journal of Clinical Investigation.

https://doi.org/10.1172/JCI42285

Guévremont, M., Martel-Pelletier, J., Massicotte, F., Tardif, G., Pelletier, J. P., Ranger, P., … Reboul, P. (2003). Human Adult Chondrocytes Express Hepatocyte Growth

Factor (HGF) Isoforms but Not HGF: Potential Implication of Osteoblasts on the Presence of HGF in Cartilage. Journal of Bone and Mineral Research.

https://doi.org/10.1359/jbmr.2003.18.6.1073

Haffner-Luntzer, M., Kovtun, A., Rapp, A. E., & Ignatius, A. (2016). Mouse Models in Bone Fracture Healing Research. Current Molecular Biology Reports.

https://doi.org/10.1007/s40610-016-0037-3

Hak, D. J. (2011). Management of aseptic tibial nonunion. Journal of the American

Academy of Orthopaedic Surgeons. https://doi.org/10.5435/00124635-201109000-

00007

Hak, D. J., Fitzpatrick, D., Bishop, J. A., Marsh, J. L., Tilp, S., Schnettler, R., … Alt, V.

(2014). Delayed union and nonunions: Epidemiology, clinical issues, and financial aspects. Injury. https://doi.org/10.1016/j.injury.2014.04.002

Hankenson, K. D., Dishowitz, M., Gray, C., & Schenker, M. (2011). Angiogenesis in bone regeneration. Injury. https://doi.org/10.1016/j.injury.2011.03.035

Heino, T. J., & Hentunen, T. A. (2008). Differentiation of osteoblasts and osteocytes from mesenchymal stem cells. Current Stem Cell Research & Therapy.

Heist, R. S., Shim, H. S., Gingipally, S., Mino-Kenudson, M., Le, L., Gainor, J. F., … Iafrate, A. J. (2016). MET Exon 14 Skipping in Non-Small Cell Lung Cancer . The

Page 134: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

122

Oncologist. https://doi.org/10.1634/theoncologist.2015-0510

Hemmings, B. A., & Restuccia, D. F. (2012). PI3K-PKB/Akt pathway. Cold Spring

Harbor Perspectives in Biology. https://doi.org/10.1101/cshperspect.a011189

Hill, T. P., Später, D., Taketo, M. M., Birchmeier, W., & Hartmann, C. (2005). Canonical Wnt/β-catenin signaling prevents osteoblasts from differentiating into chondrocytes. Developmental Cell. https://doi.org/10.1016/j.devcel.2005.02.013

Hiltunen, A., Vuorio, E., & Aro, H. T. (1993). A standardized experimental fracture in the mouse tibia. Journal of Orthopaedic Research.

https://doi.org/10.1002/jor.1100110219

Hitachi, S., Hayashi, K., Watanuki, M., Hagiwara, Y., Yoshida, S., Watanabe, M., … Hatori, M. (2018). Osteofibrous dysplasia arising in the humerus: A case report. Rare Tumors. https://doi.org/10.1177/2036361318808852

Hu, H. (2004). Sequential roles of Hedgehog and Wnt signaling in osteoblast development. Development. https://doi.org/10.1242/dev.01564

Huang, W. (2007). Signaling and transcriptional regulation in osteoblast commitment and differentiation. Frontiers in Bioscience. https://doi.org/10.2741/2296

Ito, H. (2011). Chemokines in mesenchymal stem cell therapy for bone repair: A novel concept of recruiting mesenchymal stem cells and the possible cell sources. Modern Rheumatology. https://doi.org/10.1007/s10165-010-0357-8

Jeay, S., Munshi, N., Hill, J., Moussa, M., Ashwell, M., Leggett, D., & Li, C. (2007). ARQ

197, a highly selective small molecule inhibitor of c-Met, with selective antitumor properties in a broad spectrum of human cancer cells. Cancer Res.

Jeong, W.-J., Ro, E. J., & Choi, K.-Y. (2018). Interaction between Wnt/β-catenin and

RAS-ERK pathways and an anti-cancer strategy via degradations of β-catenin and RAS by targeting the Wnt/β-catenin pathway. Npj Precision Oncology.

https://doi.org/10.1038/s41698-018-0049-y

Jha, S., Blau, J. E., & Bhattacharyya, T. (2016). Normal and Delayed Fracture Healing: Symphony and Cacophony. Hormone and Metabolic Research.

https://doi.org/10.1055/s-0042-117636

Jho, E., Zhang, T., Domon, C., Joo, C.-K., Freund, J.-N., & Costantini, F. (2002).

Wnt/beta-catenin/Tcf signaling induces the transcription of Axin2, a negative regulator of the signaling pathway. Molecular and Cellular Biology.

Kahn, L. B. (2003). Adamantinoma, osteofibrous dysplasia and differentiated adamantinoma. Skeletal Radiology. https://doi.org/10.1007/s00256-003-0624-2

Karol, L. A., Brown, D. S., Wise, C. A., & Waldron, M. (2005). Familial osteofibrous dysplasia: A case series. Journal of Bone and Joint Surgery - Series A.

https://doi.org/10.2106/JBJS.D.02575

Katagiri, T., & Takahashi, N. (2002). Regulatory mechanisms of osteoblast and osteoclast differentiation. Oral Diseases. https://doi.org/10.1034/j.1601-

0825.2002.01829.x

Kim, J. W., Lee, M. N., Jeong, B. C., Oh, S. H., Kook, M. S., & Koh, J. T. (2017). Chemical inhibitors of c-Met receptor tyrosine kinase stimulate osteoblast

Page 135: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

123

differentiation and bone regeneration. European Journal of Pharmacology.

https://doi.org/10.1016/j.ejphar.2017.03.032

Kim, S., Koga, T., Isobe, M., Kern, B. E., Yokochi, T., Chin, Y. E., … Takayanagi, H. (2003). Stat1 functions as a cytoplasmic attenuator of Runx2 in the transcriptional program of osteoblast differentiation. Genes and Development.

https://doi.org/10.1101/gad.1119303

Kjaerulff, O., Brodin, L., & Jung, A. (2011). The Structure and Function of Endophilin Proteins. Cell Biochemistry and Biophysics. https://doi.org/10.1007/s12013-010-

9137-5

Kobayashi, T., Honke, K., Miyazaki, T., Matsumoto, K., Nakamura, T., Ishizuka, I., & Makita, A. (1994). Hepatocyte growth factor specifically binds to sulfoglycolipids. Journal of Biological Chemistry.

Kobayashi, Y., Uehara, S., Udagawa, N., & Takahashi, N. (2016). Regulation of bone metabolism by Wnt signals. Journal of Biochemistry.

https://doi.org/10.1093/jb/mvv124

Koga, T., Matsui, Y., Asagiri, M., Kodama, T., De Crombrugghe, B., Nakashima, K., &

Takayanagi, H. (2005). NFAT and Osterix cooperatively regulate bone formation. Nature Medicine. https://doi.org/10.1038/nm1270

Kokabu, S. (2013). BMP and osteoblastogenesis inhibitors. Osteoblasts: Morphology,

Functions and Clinical Implications.

Komori, T., Yagi, H., Nomura, S., Yamaguchi, A., Sasaki, K., Deguchi, K., … Kishimoto, T. (1997). Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts. Cell. https://doi.org/10.1016/S0092-

8674(00)80258-5

Komori, Toshihisa. (2006). Regulation of osteoblast differentiation by transcription factors. Journal of Cellular Biochemistry. https://doi.org/10.1002/jcb.20958

Kostenuik, P., & Mirza, F. M. (2017). Fracture healing physiology and the quest for therapies for delayed healing and nonunion. Journal of Orthopaedic Research.

https://doi.org/10.1002/jor.23460

Kozhemyakina, E., Lassar, A. B., & Zelzer, E. (2015). A pathway to bone: signaling

molecules and transcription factors involved in chondrocyte development and maturation. Development. https://doi.org/10.1242/dev.105536

Kratchmarova, I., Blagoev, B., Haack-Sorensen, M., Kassem, M., & Mann, M. (2005).

Cell Signalling: Mechanism of divergent growth factor effects in mesenchymal stem cell differentiation. Science. https://doi.org/10.1126/science.1107627

Kuznetsov, S. A., Mankani, M. H., Bianco, P., & Robey, P. G. (2009). Enumeration of

the colony-forming units-fibroblast from mouse and human bone marrow in normal and pathological conditions. Stem Cell Research.

https://doi.org/10.1016/j.scr.2008.07.007

Langdahl, B., Ferrari, S., & Dempster, D. W. (2016). Bone modeling and remodeling: potential as therapeutic targets for the treatment of osteoporosis. Therapeutic Advances in Musculoskeletal Disease. https://doi.org/10.1177/1759720X16670154

Lausson, S., & Cressent, M. (2011). Signal transduction pathways mediating the effect

Page 136: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

124

of adrenomedullin on osteoblast survival. Journal of Cellular Biochemistry.

https://doi.org/10.1002/jcb.23311

Le, A. X., Miclau, T., Hu, D., & Helms, J. A. (2001). Molecular aspects of healing in stabilized and non-stabilized fractures. Journal of Orthopaedic Research.

https://doi.org/10.1016/S0736-0266(00)00006-1

Lee, C., Whang, Y. M., Campbell, P., Mulcrone, P. L., Elefteriou, F., Cho, S. W., & Park,

S. I. (2018). Dual targeting c-met and VEGFR2 in osteoblasts suppresses growth and osteolysis of prostate cancer bone metastasis. Cancer Letters.

https://doi.org/10.1016/j.canlet.2017.11.016

Lesko, E. (2007). The biological role of HGF-MET axis in tumor growth and development of metastasis. Frontiers in Bioscience. https://doi.org/10.2741/2760

Li, Y., Ge, C., & Franceschi, R. T. (2017). MAP Kinase-Dependent RUNX2

Phosphorylation Is Necessary for Epigenetic Modification of Chromatin During Osteoblast Differentiation. Journal of Cellular Physiology.

https://doi.org/10.1002/jcp.25517

Liu, T. M., & Lee, E. H. (2012). Transcriptional Regulatory Cascades in Runx2-Dependent Bone Development. Tissue Engineering Part B: Reviews.

https://doi.org/10.1089/ten.teb.2012.0527

Liu, Wei, & Zhang, X. (2015). Receptor activator of nuclear factor-κB ligand

(RANKL)/RANK/osteoprotegerin system in bone and other tissues (Review). Molecular Medicine Reports. https://doi.org/10.3892/mmr.2015.3152

Liu, Wenguang, Toyosawa, S., Furuichi, T., Kanatani, N., Yoshida, C., Liu, Y., … Komori, T. (2001). Overexpression of Cbfa1 in osteoblasts inhibits osteoblast maturation and causes osteopenia with multiple fractures. Journal of Cell Biology.

https://doi.org/10.1083/jcb.200105052

Liu, X., Yao, W., Newton, R. C., & Scherle, P. A. (2008). Targeting the c-MET signaling pathway for cancer therapy. Expert Opinion on Investigational Drugs.

https://doi.org/10.1517/13543784.17.7.997

Liu, Y., Zuo, G., Meng, X., Gao, X., Zhang, L., & Tang, P. (2017). Adrenomedullin

inhibits osteoclast differentiation through the suppression of receptor activator of nuclear factor-κB ligand-induced nuclear factor-κB activation in glucocorticoid-induced osteoporosis. Experimental and Therapeutic Medicine.

https://doi.org/10.3892/etm.2017.5025

Loi, F., Córdova, L. A., Pajarinen, J., Lin, T. hua, Yao, Z., & Goodman, S. B. (2016). Inflammation, fracture and bone repair. Bone.

https://doi.org/10.1016/j.bone.2016.02.020

Long, F., & Ornitz, D. M. (2013). Development of the endochondral skeleton. Cold Spring Harbor Perspectives in Biology.

https://doi.org/10.1101/cshperspect.a008334

Mackie, E. J., Ahmed, Y. A., Tatarczuch, L., Chen, K. S., & Mirams, M. (2008). Endochondral ossification: How cartilage is converted into bone in the developing skeleton. International Journal of Biochemistry and Cell Biology.

https://doi.org/10.1016/j.biocel.2007.06.009

Page 137: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

125

Mara, C. S. de, Sartori, A. R., Duarte, A. S., Andrade, A. L. L., Pedro, M. A. C., & Coimbra, I. B. (2011). Periosteum as a source of mesenchymal stem cells: the effects of TGF-β3 on chondrogenesis. Clinics. https://doi.org/10.1590/s1807-

59322011000300022

Marsell, R., & Einhorn, T. A. (2011). The biology of fracture healing. Injury.

https://doi.org/10.1016/j.injury.2011.03.031

Matsumura, A., Kubota, T., Taiyoh, H., Fujiwara, H., Okamoto, K., Ichikawa, D., … Otsuji, E. (2013). HGF regulates VEGF expression via the c-Met receptor downstream pathways, PI3K/Akt, MAPK and STAT3, in CT26 murine cells. International Journal of Oncology. https://doi.org/10.3892/ijo.2012.1728

Michalopoulos, G. K., & DeFrances, M. (2005). Liver Regeneration. In Adv Biochem Eng Biotechnol. https://doi.org/10.1007/b99968

Michieli, P., Basilico, C., Pennacchietti, S., Maffè, A., Tamagnone, L., Giordano, S., … Comoglio, P. M. (1999). Mutant Met-mediated transformation is ligand-dependent and can be inhibited by HGF antagonists. Oncogene.

https://doi.org/10.1038/sj.onc.1202899

Mills, L. A., Aitken, S. A., & Simpson, A. H. R. W. (2017). The risk of non-union per fracture: current myths and revised figures from a population of over 4 million adults. Acta Orthopaedica. https://doi.org/10.1080/17453674.2017.1321351

Minkin, C. (1982). Bone acid phosphatase: Tartrate-resistant acid phosphatase as a marker of osteoclast function. Calcified Tissue International.

https://doi.org/10.1007/BF02411252

Most, M. J., Sim, F. H., & Inwards, C. Y. (2010). Osteofibrous dysplasia and Adamantinoma. Journal of the American Academy of Orthopaedic Surgeons.

https://doi.org/10.5435/00124635-201006000-00008

Muro, R., Nitta, T., Okada, T., Ideta, H., Tsubata, T., & Suzuki, H. (2015). The Ras

GTPase-activating protein Rasal3 supports survival of naive T cells supports survival of naive T cells. PLoS ONE. https://doi.org/10.1371/journal.pone.0119898

Nadri, S., Soleimani, M., Hosseni, R. H., Massumi, M., Atashi, A., & Izadpanah, R.

(2007). An efficient method for isolation of murine bone marrow mesenchymal stem cells. International Journal of Developmental Biology.

https://doi.org/10.1387/ijdb.072352ns

Naldini, L., Vigna, E., Ferracini, R., Longati, P., Gandino, L., Prat, M., & Comoglio, P. M.

(1991). The Tyrosine Kinase Encoded by the MET Proto-Oncogene Is Activated by Autophosphorylation. MOLECULAR AND CELLULAR BIOLOGY.

Nelson, W. J., & Nusse, R. (2004). Convergence of Wnt, β-Catenin, and Cadherin pathways. Science. https://doi.org/10.1126/science.1094291

Neve, A., Corrado, A., & Cantatore, F. P. (2011). Osteoblast physiology in normal and pathological conditions. Cell and Tissue Research. https://doi.org/10.1007/s00441-

010-1086-1

Oberlender, S. A., & Tuan, R. S. (1994). Expression and functional involvement of N-cadherin in embryonic limb chondrogenesis. Development (Cambridge, England).

Organ, S. L., & Tsao, M. S. (2011). An overview of the c-MET signaling pathway.

Page 138: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

126

Therapeutic Advances in Medical Oncology.

https://doi.org/10.1177/1758834011422556

Otto, F., Thornell, A. P., Crompton, T., Denzel, A., Gilmour, K. C., Rosewell, I. R., … Owen, M. J. (1997). Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell.

https://doi.org/10.1016/S0092-8674(00)80259-7

Oved, S., & Yarden, Y. (2002). Signal transduction: molecular ticket to enter cells. Nature. https://doi.org/10.1038/416133a

Papaioannou, G., Mirzamohammadi, F., & Kobayashi, T. (2016). Ras signaling regulates osteoprogenitor cell proliferation and bone formation. Cell Death & Disease. https://doi.org/10.1038/cddis.2016.314

Parikh, P. K., & Ghate, M. D. (2018). Recent advances in the discovery of small molecule c-Met Kinase inhibitors. European Journal of Medicinal Chemistry.

https://doi.org/10.1016/j.ejmech.2017.08.044

Park-Min, K. H. (2018). Mechanisms involved in normal and pathological osteoclastogenesis. Cellular and Molecular Life Sciences.

https://doi.org/10.1007/s00018-018-2817-9

Park, Y. K., Unni, K. K., McLeod, R. A., & Pritchard, D. J. (1993). Osteofibrous dysplasia: Clinicopathologic study of 80 cases. Human Pathology.

https://doi.org/10.1016/0046-8177(93)90268-L

Pez Ponte, A. L., Marais, E., Gallay, N., Langonné, A., Delorme, B., Hérault, O., … Domenech, J. (2007). The In Vitro Migration Capacity of Human Bone Marrow Mesenchymal Stem Cells: Comparison of Chemokine and Growth Factor Chemotactic Activities. STEM CELLS. https://doi.org/10.1634/stemcells.2007-0054

Ponzetto, C., Bardelli, A., Zhen, Z., Maina, F., dalla Zonca, P., Giordano, S., … Comoglio, P. M. (1994). A multifunctional docking site mediates signaling and transformation by the hepatocyte growth factor/scatter factor receptor family. Cell.

https://doi.org/10.1016/0092-8674(94)90318-2

R.T., F., C., G., G., X., H., R., & D., J. (2009). Transcriptional regulation of osteoblasts. Cells Tissues Organs.

Raucci, A., Bellosta, P., Grassi, R., Basilico, C., & Mansukhani, A. (2008). Osteoblast proliferation or differentiation is regulated by relative strengths of opposing signaling pathways. Journal of Cellular Physiology. https://doi.org/10.1002/jcp.21323

Regard, J. B., Cherman, N., Palmer, D., Kuznetsov, S. A., Celi, F. S., Guettier, J.-M., … Yang, Y. (2011). Wnt/ -catenin signaling is differentially regulated by G proteins and contributes to fibrous dysplasia. Proceedings of the National Academy of

Sciences. https://doi.org/10.1073/pnas.1114656108

Runyan, C. M., & Gabrick, K. S. (2017). Biology of bone formation, fracture healing, and distraction osteogenesis. Journal of Craniofacial Surgery.

https://doi.org/10.1097/SCS.0000000000003625

Rutkovskiy, A., Stensløkken, K.-O., & Vaage, I. J. (2016). Osteoblast Differentiation at a Glance. Medical Science Monitor Basic Research.

https://doi.org/10.12659/msmbr.901142

Page 139: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

127

Sattler, M., & Salgia, R. (2009). The MET axis as a therapeutic target. Update on Cancer Therapeutics. https://doi.org/10.1016/j.uct.2009.01.001

Schlomer, B. J., Feretti, M., Rodriguez, E., Blaschko, S., Cunha, G., & Baskin, L. (2013). Sexual differentiation in the male and female mouse from days 0 to 21: A detailed and novel morphometric description. Journal of Urology.

https://doi.org/10.1016/j.juro.2013.02.3198

Schmidt, C., Bladt, F., Goedecke, S., Brinkmann, V., Zschiesche, W., Sharpe, M., … Goedecke, S. (1995). Scatter factor/hepatocyte growth factor is essential for liver development. Nature. https://doi.org/10.1038/373699a0

Shang, S., Hua, F., & Hu, Z.-W. (2017). The regulation of &#x3B2;-catenin activity and function in cancer: therapeutic opportunities. Oncotarget.

https://doi.org/10.18632/oncotarget.15687

Shapiro, F. (2008). Bone development and its relation to fracture repair. The role of mesenchymal osteoblasts and surface osteoblasts. European Cells and Materials.

https://doi.org/10.22203/eCM.v015a05

Sharma, R., Wu, X., Rhodes, S. D., Chen, S., He, Y., Yuan, J., … Yang, F. C. (2013).

Hyperactive Ras/MAPK signaling is critical for tibial nonunion fracture in neurobromin-decient mice. Human Molecular Genetics.

https://doi.org/10.1093/hmg/ddt333

Shibasaki, S., Kitano, S., Karasaki, M., Tsunemi, S., Sano, H., & Iwasaki, T. (2015). Blocking c-Met signaling enhances bone morphogenetic protein-2-induced osteoblast differentiation. FEBS Open Bio. https://doi.org/10.1016/j.fob.2015.04.008

Soysa, N. S., & Alles, N. (2016). Osteoclast function and bone-resorbing activity: An overview. Biochemical and Biophysical Research Communications.

https://doi.org/10.1016/j.bbrc.2016.05.019

Stamos, J. L., & Weis, W. I. (2013). The β-catenin destruction complex. Cold Spring

Harbor Perspectives in Biology. https://doi.org/10.1101/cshperspect.a007898

Standal, T., Abildgaard, N., Fagerli, U. M., Stordal, B., Hjertner, Ø., Borset, M., & Sundan, A. (2007). HGF inhibits BMP-induced osteoblastogenesis: Possible implications for the bone disease of multiple myeloma. Blood.

https://doi.org/10.1182/blood-2006-07-034884

Sun, M. M. G., & Beier, F. (2014). Chondrocyte hypertrophy in skeletal development, growth, and disease. Birth Defects Research Part C - Embryo Today: Reviews.

https://doi.org/10.1002/bdrc.21062

Tall, M. (2018). Treatment of aseptic tibial shaft non-union without bone defect. Orthopaedics and Traumatology: Surgery and Research.

https://doi.org/10.1016/j.otsr.2017.05.028

Taylor, R. M., Kashima, T. G., Ferguson, D. J., Szuhai, K., Hogendoorn, P. C., & Athanasou, N. A. (2012). Analysis of stromal cells in osteofibrous dysplasia and adamantinoma of long bones. Modern Pathology.

https://doi.org/10.1038/modpathol.2011.141

Teis, D., & Huber, L. A. (2003). The odd couple: Signal transduction and endocytosis. Cellular and Molecular Life Sciences. https://doi.org/10.1007/s00018-003-3010-2

Page 140: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

128

Thomas, D. M., Johnson, S. A., Sims, N. A., Trivett, M. K., Slavin, J. L., Rubin, B. P., … Hinds, P. W. (2004). Terminal osteoblast differentiation, mediated by runx2 and p27 KIP1, is disrupted in osteosarcoma. Journal of Cell Biology.

https://doi.org/10.1083/jcb.200409187

Thompson, E. M., Matsiko, A., Farrell, E., Kelly, D. J., & O’Brien, F. J. (2015).

Recapitulating endochondral ossification: A promising route to in vivo bone regeneration. Journal of Tissue Engineering and Regenerative Medicine.

https://doi.org/10.1002/term.1918

Thompson, Z., Miclau, T., Hu, D., & Helms, J. A. (2002). A model for intramembranous ossification during fracture healing. Journal of Orthopaedic Research.

https://doi.org/10.1016/S0736-0266(02)00017-7

Trusolino, L., Bertotti, A., & Comoglio, P. M. (2010). MET signalling: Principles and functions in development, organ regeneration and cancer. Nature Reviews Molecular Cell Biology. https://doi.org/10.1038/nrm3012

Tsai, S. Y., Huang, Y. L., Yang, W. H., & Tang, C. H. (2012). Hepatocyte growth factor-

induced BMP-2 expression is mediated by c-Met receptor, FAK, JNK, Runx2, and p300 pathways in human osteoblasts. International Immunopharmacology.

https://doi.org/10.1016/j.intimp.2012.03.026

Väänänen, H. K., Zhao, H., Mulari, M., & Halleen, J. M. (2000). The cell biology of osteoclast function. Journal of Cell Science.

Vagin, O., & Beenhouwer, D. O. (2016). Septins: Regulators of Protein Stability. Frontiers in Cell and Developmental Biology.

https://doi.org/10.3389/fcell.2016.00143

Vallet, S., Bashari, M. H., Fan, F. J., Malvestiti, S., Schneeweiss, A., Wuchter, P., … Podar, K. (2016). Pre-osteoblasts stimulate migration of breast cancer cells via the HGF/MET pathway. PLoS ONE. https://doi.org/10.1371/journal.pone.0150507

Vimalraj, S., Arumugam, B., Miranda, P. J., & Selvamurugan, N. (2015). Runx2: Structure, function, and phosphorylation in osteoblast differentiation. International Journal of Biological Macromolecules.

https://doi.org/10.1016/j.ijbiomac.2015.04.008

Yamaguchi, A., Komori, T., & Suda, T. (2000). Regulation of osteoblast differentiation mediated by bone morphogenetic proteins, hedgehogs, and Cbfa1. Endocrine

Reviews. https://doi.org/10.1210/edrv.21.4.0403

Yamashita, M., Otsuka, F., Mukai, T., Otani, H., Inagaki, K., Miyoshi, T., … Makino, H. (2008). Simvastatin antagonizes tumor necrosis factor-α ihibition of bone morphogenetic proteins-2-induced osteoblast differentiation by regulating Smad signaling and Ras/Rho-mitogen-activated protein kinase pathway. Journal of Endocrinology. https://doi.org/10.1677/JOE-07-0532

Yan, Y., Tang, D., Chen, M., Huang, J., Xie, R., Jonason, J. H., … Chen, D. (2009).

Axin2 controls bone remodeling through the -catenin-BMP signaling pathway in adult mice. Journal of Cell Science. https://doi.org/10.1242/jcs.051904

Yang, L., Tsang, K. Y., Tang, H. C., Chan, D., & Cheah, K. S. E. (2016). Hypertrophic

chondrocytes can become osteoblasts and osteocytes in endochondral bone

Page 141: INSIGHTS INTO FRACTURE REPAIR USING A MURINE MODEL …...identified. Fracture repair is delayed Δ15in Met-HET mice, with decreased bone formation 2-weeks post fracture-inducing surgery

129

formation. World Review of Nutrition and Dietetics.

https://doi.org/10.1159/000441808

Zhang, Y. W., Graveel, C., Shinomiya, N., & Vande Woude, G. F. (2004). Met decoys: Will cancer take the bait? Cancer Cell. https://doi.org/10.1016/j.ccr.2004.07.003

Zhou, X., von der Mark, K., Henry, S., Norton, W., Adams, H., & de Crombrugghe, B. (2014). Chondrocytes Transdifferentiate into Osteoblasts in Endochondral Bone during Development, Postnatal Growth and Fracture Healing in Mice. PLoS Genetics. https://doi.org/10.1371/journal.pgen.1004820