the consequences of collagen degradation … · the consequences of collagen degradation on bone...

244
THE CONSEQUENCES OF COLLAGEN DEGRADATION ON BONE MECHANICAL PROPERTIES by Chrystia Wynnyckyj A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Materials Science and Engineering University of Toronto © Copyright by Chrystia Wynnyckyj (2010)

Upload: nguyenquynh

Post on 08-Apr-2018

219 views

Category:

Documents


4 download

TRANSCRIPT

THE CONSEQUENCES OF COLLAGEN

DEGRADATION ON BONE MECHANICAL

PROPERTIES

by

Chrystia Wynnyckyj

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Graduate Department of Materials Science and Engineering

University of Toronto

© Copyright by Chrystia Wynnyckyj (2010)

ii

The consequences of collagen degradation on bone mechanical

properties

Doctor of Philosophy

Chrystia Wynnyckyj

Materials Science and Engineering

University of Toronto

2010

Abstract

The mechanisms underlying the effect of alterations in Type I collagen on bone mechanical

properties are not well defined. Clinical tools for evaluating fracture risk, such as dual energy

x-ray absorptiometry (DXA) and quantitative ultrasound (QUS) focus on bone mineral and

cannot detect changes in the collagen matrix. The mechanical response tissue analyzer

(MRTA) is a potential tool for evaluating fracture risk. Thus, the focus of this work was to

investigate the effects of collagen degradation on bone mechanical properties and examine

whether clinical tools can detect these changes.

Female and male emu tibiae were endocortically treated with 1 M potassium hydroxide

(KOH) solution for 1-14 days and then either mechanically tested in three-point bending,

fatigued to failure or fatigued to induce stiffness loss. Computed Tomography scans, DXA,

QUS, MRTA and three-point bend testing in the elastic region were performed on emu tibiae

iii

before and after either KOH treatment or fatigue to induce stiffness loss. Fracture surfaces

were examined to determine failure mechanisms. Bone mineral and bone collagen were

characterized using appropriate techniques. Bone mineral-collagen interface was investigated

using Raman spectroscopy and atomic force microscopy (AFM).

Endocortical KOH treatment does not affect bone mineral however, it causes in situ collagen

degradation, rather than removal and may be weakening the mineral-collagen interface.

These changes result in significantly compromised mechanical properties. Emu tibiae show

significant decreases in failure stress and increased failure strain and toughness, with

increasing KOH treatment time. The significant increase in toughness of KOH treated bones

is due to structural alterations that enhance the ability of the microstructure to dissipate

energy during the failure process, thereby slowing crack propagation, as shown by fracture

surface analysis. KOH treated samples exhibit a lower fatigue resistance compared to

untreated samples at high stresses only for both sexes. Partial fatigue testing results in similar

decreases in modulus for all groups and sexes. The MRTA detected these changes whereas

DXA and QUS did not. MRTA detects changes in bone mechanical properties induced by

changes in collagen quality and fatigue and could be a more effective tool for predicting

fracture risk.

iv

Acknowledgements

This thesis could not have been accomplished without the input from a great number of

people, whose contribution in assorted ways deserves special mention.

First and foremost, I would like to thank Dr. Marc Grynpas for his supervision and guidance

from the very early stages of this research. Above all, you provided me with encouragement

and support in various ways that were needed throughout the bumpy road of completing this

thesis. Your ideas and passion for science constantly inspired and enriched my growth as a

student and as a researcher. I would like to thank my committee members, Dr. Harvey

Goldberg, Dr. Robert Pilliar, Dr. Eli Sone and Dr. Zhirui Wang for their endless questions

and suggestions that guided and challenged my thinking, substantially improving the finished

product. Finally, thank you to my external examiner, Dr. Mitchell B. Schaffler, for reading

my thesis in record time and for the great discussion.

This thesis would not have been possible without the help of several individuals who in one

way or another contributed and extended their valuable technical assistance in the preparation

and completion of my research. My deepest gratitude is extended to Richard Cheung, Doug

Holmyard, Jackie Razik, Kerri Tupy and Gabriela Voiet. A special thank you to Dr. Mircea

Dumitriu for mineralization analysis but more importantly, for his unique sense of humour. I

must also acknowledge the support of Fanny Strumas, Maria Fryman and Jody Prentice from

the Materials Science Engineering department. I would like to thank current and past

members of the Grynpas and Kandel labs, particularly Sidney Omelon, Lisa Wise-Milestone,

Tanya Hunt and Kimberly Kyle for our numerous stimulating discussions, emotional support,

entertainment and general advice, which helped enrich this wonderful experience.

In addition to the assistance above, I received equally important assistance from family and

friends. My mother instilled in me, from an early age, the desire and skills to be the never-

ending student. A special thank you is due to my brother, Marco, for his support. Finally,

words fail to express my gratitude to my husband, Yurij, whose love, dedication,

understanding and patience made it possible for me to finish my thesis.

v

Table of Contents

Abstract.................................................................................................................................... ii Acknowledgements ................................................................................................................ iv Table of Contents .................................................................................................................... v List of Tables ........................................................................................................................... x List of Figures......................................................................................................................... xi List of Abbreviations ............................................................................................................ xv CHAPTER 1: INTRODUCTION.......................................................................................... 1 1.1 Rationale and Background........................................................................................... 2 1.1.1. Hierarchical structure of bone.............................................................................. 2 1.1.2. Bone composition.................................................................................................. 5 1.1.3. Bone mineral : Hydroxyapatite............................................................................. 5 1.1.4. Bone matrix : Collagen......................................................................................... 6 1.1.5. Noncollagenous proteins (NCPs) ......................................................................... 8 1.1.6. Mineral-collagen interface ................................................................................... 9 1.1.7. Bone mineralization ............................................................................................ 10 1.2 How to Affect Bone Collagen? ................................................................................... 12 1.3 Bone Mechanics........................................................................................................... 13 1.4 Fatigue.......................................................................................................................... 16 1.4.1 Bone fatigue ......................................................................................................... 17 1.4.2 Bone creep strain ................................................................................................. 20 1.5 Bone Toughening Mechanisms .................................................................................. 20 1.6 Tools for Fracture Risk Assessment.......................................................................... 21 1.6.1 Dual Energy X-ray Absorptiometry (DXA).......................................................... 22 1.6.2 Quantitative Ultrasound (QUS)........................................................................... 22 1.6.3 Mechanical Response Tissue Analyzer (MRTA) .................................................. 23 1.7 Animal Models - Emu................................................................................................. 26 1.8 Objectives..................................................................................................................... 26 1.9 Hypothesis.................................................................................................................... 27 1.10 References.................................................................................................................... 28 CHAPTER 2: EXPERIMENTAL APPROACH ............................................................... 37 2.1 Bone Samples............................................................................................................... 38 2.2 Potassium Hydroxide (KOH) Treatment.................................................................. 43 2.3 Statistical Analysis ...................................................................................................... 44 2.4 References.................................................................................................................... 45

vi

CHAPTER 3: INITIAL STUDY ......................................................................................... 46 3.1 Introduction................................................................................................................. 47 3.2 Experimental Details .................................................................................................. 49 3.2.1. Emu bone samples............................................................................................... 49 3.2.2. Bone composition................................................................................................ 50 3.2.3. KOH treatment.................................................................................................... 51 3.2.4. CT........................................................................................................................ 53 3.2.5. DXA..................................................................................................................... 53 3.2.6. QUS..................................................................................................................... 53 3.2.7. MRTA .................................................................................................................. 53 3.2.8. Mechanical testing .............................................................................................. 54 3.2.9. Statistical analysis .............................................................................................. 55 3.3 Results .......................................................................................................................... 55 3.3.1. Emu bone composition........................................................................................ 55 3.3.2. Sex differences in untreated and control bones .................................................. 56 3.3.3. Effect of KOH treatment on bone composition ................................................... 57 3.3.4. Effect of KOH treatment on BMD, SOS, EI ........................................................ 59 3.3.5. Effect of KOH treatment on structural and mechanical properties.................... 59 3.4 Discussion..................................................................................................................... 63 3.4.1. Emu bone composition........................................................................................ 63 3.4.2. KOH treatment and collagen degradation ......................................................... 63 3.4.3. Effect of KOH treatment on bone ductility and toughness ................................. 64 3.4.4. Composite material behaviour............................................................................ 65 3.4.5. Sex differences .................................................................................................... 67 3.4.6. Clinical tools....................................................................................................... 68 3.5 Conclusions.................................................................................................................. 69 3.6 Chapter Summary ...................................................................................................... 70 3.7 References.................................................................................................................... 71 CHAPTER 4: FATIGUE ..................................................................................................... 76 4.1 Introduction................................................................................................................. 77 4.2 Experimental Details .................................................................................................. 79 4.2.1. Emu bone samples and KOH treatment.............................................................. 79 4.2.2. Fatigue testing to failure..................................................................................... 80 4.2.3. Partial fatigue testing ......................................................................................... 83 4.2.4. CT........................................................................................................................ 83 4.2.5. DXA..................................................................................................................... 84 4.2.6. QUS..................................................................................................................... 84 4.2.7. MRTA .................................................................................................................. 84 4.2.8. Three-point bend testing ..................................................................................... 84 4.2.9. Microdamage ...................................................................................................... 85 4.2.10. Statistical analysis ............................................................................................ 86 4.3 Results .......................................................................................................................... 86 4.3.1. Effect of KOH treatment on fatigue properties................................................... 86

vii

4.3.2. Microdamage ...................................................................................................... 92 4.3.3. Effect of partial fatigue testing on BMD, SOS and EI ........................................ 92 4.4 Discussion..................................................................................................................... 94 4.4.1. Fatigue behaviour............................................................................................... 94 4.4.2. Microdamage ...................................................................................................... 97 4.4.3. Sex differences .................................................................................................... 98 4.4.4. Clinical tools....................................................................................................... 99 4.5 Conclusions................................................................................................................ 100 4.6 Chapter Summary .................................................................................................... 100 4.7 References.................................................................................................................. 101 CHAPTER 5: FRACTOGRAPHY ................................................................................... 105 5.1 Introduction............................................................................................................... 106 5.2 Experimental Details ................................................................................................ 107 5.2.1. Emu bone samples and KOH treatment............................................................ 107 5.2.2. Fractography analysis ...................................................................................... 109 5.2.3. Surface roughness measurements ..................................................................... 113 5.2.4. Statistical analysis ............................................................................................ 113 5.3 Results ........................................................................................................................ 114 5.3.1. Tensile versus compressive areas ..................................................................... 114 5.3.2. Degree of roughness ......................................................................................... 115 5.3.3. Regions of interests ........................................................................................... 117 5.3.4. Correlations ...................................................................................................... 121 5.4 Discussion................................................................................................................... 124 5.4.1. Tensile versus compressive areas ..................................................................... 125 5.4.2. Degree of roughness ......................................................................................... 126 5.4.3. Failure mechanisms .......................................................................................... 127 5.5 Conclusions................................................................................................................ 129 5.6 Chapter Summary .................................................................................................... 130 5.7 References.................................................................................................................. 131 CHAPTER 6: COLLAGEN DEGRADATION ............................................................... 134 6.1 Introduction............................................................................................................... 135 6.2 Experimental Details ................................................................................................ 137 6.2.1. Reagents............................................................................................................ 137 6.2.2. Emu bone samples and KOH treatment............................................................ 137 6.2.3. Powder X-ray diffaction.................................................................................... 138 6.2.4. Quantitative backscattered electron imaging ................................................... 140 6.2.5. Microhardness testing....................................................................................... 141 6.2.6. Bone powder preparation ................................................................................. 142 6.2.7. α-Chymotrypsin ................................................................................................ 142 6.2.8. DSC................................................................................................................... 143

viii

6.2.9. SDS-PAGE ........................................................................................................ 145 6.2.10. Polarized light microscopy ............................................................................. 145 6.2.11. Statistical analysis .......................................................................................... 146 6.3 Results ........................................................................................................................ 146 6.3.1. Powder X-ray diffraction .................................................................................. 147 6.3.2. Microhardness .................................................................................................. 147 6.3.3. Quantitative backscattered electron imaging ................................................... 147 6.3.4. α-Chymotrypsin ................................................................................................ 150 6.3.5. DSC................................................................................................................... 152 6.3.6. SDS-PAGE ........................................................................................................ 157 6.3.7. Polarized light microscopy ............................................................................... 161 6.4 Discussion................................................................................................................... 163 6.4.1. Mineral characterization .................................................................................. 163 6.4.2. Collagen degradation ....................................................................................... 164 6.4.3. Partial debonding of the collagen-mineral interface........................................ 167 6.4.4. Sex differences .................................................................................................. 168 6.5 Conclusions................................................................................................................ 169 6.6 Chapter Summary .................................................................................................... 170 6.7 References.................................................................................................................. 171 CHAPTER 7: INTERFACE .............................................................................................. 177 7.1 Introduction............................................................................................................... 178 7.2 Experimental Details ................................................................................................ 181 7.2.1. Emu bone samples and KOH treatment............................................................ 181 7.2.2. Raman spectroscopy data acquisition .............................................................. 181 7.2.3. Raman spectroscopy data analysis ................................................................... 182 7.2.4. Atomic force microscopy imaging .................................................................... 183 7.2.5. Secondary surface roughness measurements.................................................... 184 7.2.6. Statistical analysis ............................................................................................ 185 7.3 Results ........................................................................................................................ 185 7.3.1. Raman spectroscopy analysis ........................................................................... 185 7.3.2. Atomic force microscopy................................................................................... 190 7.4 Discussion................................................................................................................... 194 7.4.1. Raman spectroscopy ......................................................................................... 195 7.4.2. Atomic force microscopy................................................................................... 196 7.5 Conclusions................................................................................................................ 199 7.6 Acknowledgement ..................................................................................................... 199 7.7 Chapter Summary .................................................................................................... 200 7.8 References.................................................................................................................. 201 CHAPTER 8: CONCLUSIONS ........................................................................................ 205

ix

CHAPTER 9: FUTURE WORK ....................................................................................... 208 9.1 Introduction............................................................................................................... 209 9.2 Future Work.............................................................................................................. 209 9.2.1. Objective 1 - Initial study.................................................................................. 209 9.2.2. Objective 2 - Fatigue ........................................................................................ 210 9.2.3. Objective 4 - Collagen degradation.................................................................. 211 9.2.4. Objective 5 - Interface ...................................................................................... 212 9.3 References.................................................................................................................. 213 APPENDICES..................................................................................................................... 215 APPENDIX A: Untreated (0-day) versus Control (14-day filled with saline) results .. 215 A.1 Objective 1 - Initial study ........................................................................................ 216 A.2 Objective 2 - Fatigue ................................................................................................ 218 A.3 Objective 3 - Fractography ..................................................................................... 222 A.4 Objective 4 - Collagen degradation ........................................................................ 226 A.5 Objective 5 - Interface ............................................................................................. 228

x

List of Tables Table 2.1 Sample size of female and male emu bones 40 Table 2.2 Summary of techniques 41 Table 3.1 Collagen, mineral and fat content measurements for female and male

emu tibiae 56

Table 3.2 Average geometrical parameters, BMD, BMC and mechanical properties prior to KOH treatment

57

Table 3.3 Average geometrical parameter changes due to KOH treatment for female and male emu tibiae

60

Table 4.1 Sample size of female and male emu tibiae for fatigue testing 80 Table 4.2 BMD and geometrical parameters of emu tibiae after KOH treatment

but prior to fatigue to failure testing 88

Table 4.3 Initial secant modulus, creep strain at fracture and strain at fracture for fatigue to failure emu tibiae

88

Table 4.4 Strain levels during partial fatigue testing of emu tibiae 92 Table 4.5 Average geometrical parameter changes after KOH treatment and

partial fatigue testing of emu tibiae 94

Table 5.1 Percent area roughness and surface roughness (profiler) measurements

for female and male untreated and 14-day KOH treated bones 117

Table 5.2 Porosity parameters from tensile side of female and male emu tibiae fracture surfaces

119

Table 5.3 Correlations between bone fracture surface features and mechanical properties for female and male emu tibiae

122

Table 6.1 Emu bone mineral crystal length (002) and cross section (310)

estimated by XRD and microhardness testing results 147

Table 6.2 Quantitative BSE results for tibiae of female and male KOH treated bone

148

Table 6.3 Average female and male emu tibiae thermal characteristics from DSC 153 Table 7.1 Average female and male emu tibiae Raman spectroscopy parameters 187 Table 7.2 Quantitative grey level distribution and surface roughness results for

female and male emu tibiae 194

xi

List of Figures

Figure 1.1 The hierarchical organization of bone 4 Figure 1.2 The interior structure of a collagen molecule 6 Figure 1.3 The arrangement of collagen molecules within a fibril showing the gap

and overlap regions for mineral deposition 7

Figure 1.4 Organization and interaction of collagen and mineral at different structural levels of hierarchy

12

Figure 1.5 Ideal load-displacement curve 15 Figure 1.6 Ideal stress-strain curve 16 Figure 1.7 Schematic diagram of stress versus number of cycles to failure curve 17 Figure 1.8 Schematic diagram of the stages of fatigue in bone 19 Figure 1.9 Schematic diagram of bone toughening mechanisms: (a)

microcracking, (b) uncracked ligament bridging, (c) crack bridging by collagen fibers and (d) crack deflection by osteons

21

Figure 1.10 MRTA setup for an ulna measurement 24 Figure 2.1 Schematic of sample retrieval for analysis of the different techniques 40 Figure 2.2 Flowchart summary of experimental techniques 42 Figure 2.3 KOH treatment setup 43 Figure 3.1 Percent collagen weight removed versus KOH treatment time for

female and male emu tibiae 58

Figure 3.2 Percent bone weight loss versus KOH treatment time for female and male emu tibiae

58

Figure 3.3 Percent changes of bone quality measurements reported by the different measurements techniques as a function of KOH treatment time

59

Figure 3.4 Representative stress-strain curves for 0-14 day KOH treatment of (a) female and (b) male emu tibiae

61

Figure 3.5 Mechanical properties as a function of KOH treatment time for female and male emu tibiae: (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness

62

Figure 4.1 Schematic diagram of hysteresis curves produced from a typical

fatigue test and resulting defined data 82

Figure 4.2 Schematic diagram of characteristic creep curve observed during fatigue testing showing the three characteristic stages

83

Figure 4.3 Percent bone weight loss versus KOH treatment time for female and male fatigue emu tibiae

89

Figure 4.4 Percent collagen removed versus KOH treatment time for female and male fatigue emu tibiae

89

xii

Figure 4.5 Peak stress versus log(N) curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae

90

Figure 4.6 Peak stress versus damage index rate curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae

91

Figure 4.7 Peak stress versus creep rate curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae

92

Figure 4.8 Percent changes of bone quality measurements after partial fatigue testing reported by the different measurement techniques as a function of KOH treatment time for (a) female and (b) male emu tibiae

93

Figure 4.9 Cumulative damage model showing the transition from creep to crack accumulation behaviour

96

Figure 5.1 Representative digital images of female and male, untreated and 14-

day KOH treated fracture surfaces 110

Figure 5.2 Representative tensile fracture surfaces of male 14-day KOH treated emu tibia

111

Figure 5.3 Representative compressive fracture surfaces of female 14-day KOH treated emu tibia

111

Figure 5.4 Schematic diagram of bone toughening mechanisms: (a) microcracking, (b) uncracked ligament bridging, (c) crack bridging by collagen fibers and (d) crack deflection by osteons

112

Figure 5.5 Representative stress-strain curves for untreated and 14-day KOH treated emu tibiae

114

Figure 5.6 Representative SEM images of untreated and 14-day KOH treated female and male tensile emu tibiae fracture surfaces

118

Figure 5.7 Representative SEM images of male compressive fracture surfaces of untreated and 14-day KOH treated emu tibiae

120

Figure 5.8 Representative SEM images of female compressive fracture surfaces of untreated and 14-day KOH treated emu tibiae

121

Figure 5.9 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the relative roughness area on the tensile fracture surface of bone for female and male emu tibiae

123

Figure 5.10 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the relative roughness area on the compressive fracture surface of bone for female and male emu tibiae

124

Figure 6.1 Schematic diagram of a typical DSC curve and the definitions of the

DSC parameters measured 144

Figure 6.2 BSE images of (a) untreated female, (b) untreated male, (c) 14-day female and (d) 14-day male emu tibiae

149

Figure 6.3 Percent digested collagen as a function of KOH treatment time for female and male emu tibiae

150

xiii

Figure 6.4 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the percent digested collagen for female and male emu tibiae

151

Figure 6.5 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the FWHMH (from DSC curves) for female and male emu tibiae

154

Figure 6.6 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the Height (from DSC curves) for female and male emu tibiae

155

Figure 6.7 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the Tonset (from DSC curves) for female and male emu tibiae

156

Figure 6.8 SDS-PAGE results: (a) analysis of proteins from select samples from each treatment group for female and male emu tibiae, (b) representative densitometric scans of proteins from (a), (c) plot of average relative peak intensity band area versus KOH treatment time for female and male emu tibiae from (a)

157

Figure 6.9 Average percent change in α-chain band intensity area for all samples as a function of KOH treatment time for female and male emu tibiae

159

Figure 6.10 Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the collagen peak area intensity (from SDS-PAGE) for female and male emu tibiae

160

Figure 6.11 Images of transcortical sections of demineralized 14-day KOH treated male and female emu tibiae viewed under non-polarized and polarized light

162

Figure 7.1 Typical Raman spectra of bone showing the calculation of degree of

mineralization, carbonate substitution and crystallinity 183

Figure 7.2 Typical Raman spectra scans taken from male untreated and 14-day KOH treated samples

186

Figure 7.3 Typical Raman spectra scans taken from female untreated and 14-day KOH treated samples

186

Figure 7.4 Changes in Raman peak widths (FWHMH) and Raman peak positions of (a) phosphate band and (b) carbonate band for female and male emu tibiae samples

188

Figure 7.5 Changes in Raman peak widths (FWHMH) and Raman peak positions of (a) amide III band, (b) C-H bending band and (c) amide I band for female and male emu tibiae samples

189

Figure 7.6 A 20 x 20 μm2 AFM tapping mode image of the endocortical surface of male cortical bone, showing (a) densely packed agglomerated spheroidal particles and (b) a higher resolution image from (a), of untreated male sample. Figure (c) represents a 14-day KOH treated image and (d) a higher resolution image from (c).

191

xiv

Figure 7.7 A 20 x 20 μm2 AFM tapping mode image of the endocortical surface of female cortical bone, showing (a) densely packed agglomerated spheroidal particles and (b) a higher resolution image from (a), of untreated female sample. Figure (c) represents a 14-day KOH treated image and (d) a higher resolution image from (c).

192

xv

List of Abbreviations αCT alpha chymotrypsin ΔH enthalpy of denaturation %CS percent area of compressive surfaces %TS percent area of tensile surfaces AFM atomic force microscopy B.Ar. bone cortical area BMC bone mineral content BMD bone mineral density BSP bone sialoprotein Cr.Le. microcrack mean length Cr.Dn. microcrack density Cr.S.Dn. surface microcrack density CT computed tomography CR creep rate DI damage index DSC differential scanning calorimetry DXA dual energy x-ray absorptiometry EDTA ethylenediaminetetraacetic acid EI cross-sectional bending stiffness FWHMH full width at half maximum height GuHCl guanidine hydrochloride HA hydroxyapatite KOH potassium hydroxide MRTA mechanical response tissue analyzer NaOH sodium hydroxide NCP noncollagenous protein OH-Pro hydroxyproline PI protease inhibitors PLM polarized light microscopy PMMA polymethylmethacrylate qBSE quantitative back scattered electron QUS quantitative ultrasound SAXS scanning small-angle x-ray scattering SDS sodium dodecyl sulfate SDS-PAGE sodium dodecyl sulfate polyacrylamide gel electrophoresis SEM scanning electron microscope S/N stress vs. number of cycles SOS speed of sound TEM transmission electron microscopy XRD x-ray diffraction

1

Chapter 1 Introduction 1.1. Rationale and Background

1.1.1. Hierarchical structure of bone 1.1.2. Bone composition 1.1.3. Bone mineral : Hydroxyapatite 1.1.4. Bone matrix : Collagen 1.1.5. Noncollagenous Proteins (NCPs) 1.1.6. Mineral-collagen interface 1.1.7. Bone mineralization

1.2. How to Affect Bone Collagen?

1.3. Bone Mechanics

1.4. Fatigue

1.4.1. Bone fatigue 1.4.2. Bone creep strain

1.5. Bone Toughening Mechanisms

1.6. Tools for Fracture Risk Assessment

1.6.1. Dual Energy X-ray Absorptiometry (DXA) 1.6.2. Quantitative Ultrasound (QUS) 1.6.3. Mechanical Response Tissue Analyzer (MRTA)

1.7. Animal Models – Emu

1.8. Objectives

1.9. Hypothesis

1.10. References

2

1.1. Rationale and Background

Bone is a complex composite material, consisting of mainly two phases: a mineral phase

embedded within a compliant organic matrix. The mineral phase largely contributes to the

overall strength and stiffness of bone [1,2]. It has also been shown that the organic phase

contributes to bone toughness and may also affect bone strength [3]. The mechanical

properties of bone depend on the characteristics of the mineral, collagen and the interaction

between the mineral and collagen [4-6]. Studies have been performed on the effects of

decreased bone quality due to changes in bonding between the mineral and organic phases

[5-8]. The individual contributions of the mineral and organic phases to the mechanical

properties of bone have been studied, with the mineral receiving the majority of research

interests. The mechanisms underlying the effect of alterations in Type I collagen on bone

mechanical properties are not well defined. When these major effects on bone properties are

considered, it is evident that the changes in the collagen network of bone warrant a

quantitative evaluation of varying degrees of collagen loss or degradation on bone

biomechanics.

1.1.1. Hierarchical structure of bone

Bone has a complex hierarchical microstructure [2,9,10] that can be considered at many

dimensional scales (Figure 1.1). From a macroscopic point of view, bone tissue is

nonhomogeneous, porous and anisotropic and can be classified into two types: cortical and

trabecular bone. Trabecular bone has a porosity ranging from 50-95% and is found in the end

of long bones, in vertebrae and flat bones. The pores are interconnected and filled with

marrow while the bone matrix has the form of plates and struts called trabeculae (Figure 1.1

(d)). The individual trabeculae (struts) are connected at nodes. At the microscale, individual

trabeculae have a layered arrangement of lamellae (Figure 1.1 (e)). Therefore, each

trabeculae is itself a plywood-like composite that can resist failure due to bending. Cortical

bone is much denser with a porosity ranging between 5-10% and is found primarily in the

shaft of long bones and surrounding the trabecular bone forming the external shell of flat

bones [2].

3

Bone can also be either woven or lamellar. Woven bone is laid down rapidly during growth

or repair and the fibres are aligned at random and as a result, has low strength. In contrast,

lamellar bone has parallel fibres and is much stronger. Woven bone is replaced by lamellar

bone as growth continues [1].

At the microstructural length-scale, cortical bone is made up of a structure of Haversian

systems or osteons (Figure 1.1 (b)) formed by cylindrical lamellae surrounding a Haversian

canal [2]. Each osteon consists of a blood vessel surrounded by concentric layers of lamellae

and osteons are connected through a lamellar matrix. Osteons run parallel to the long axis of

the bone and act as reinforcing tubes in the lamellar matrix [2]. Therefore, this osteonic

structure assists in the axial support of the cortical bone.

The next hierarchical level focuses on the structure of the individual lamellae. Lamellae are

composed of collagen fibers in a mineral matrix [2]. The inset of Figure 1.1 (e) shows an

individual collagen fiber protruding from this matrix. Figure 1.1 (f) represents collagen-rich

and collagen-poor domains [10].

The single lamella structure is further broken down into collagen fiber assemblies of collagen

fibrils [2]. These fibrils themselves can again be broken down into what are generally

assumed the elementary components of bone: collagen molecules and mineral particles

(Figure 1.1 (g)) [9]. In this manner, bone is a composite of a hard and brittle mineral phase

(E=135 GPa, εf=0.1%) and a soft ductile collagen phase (E=1 GPa, εf=10%), which is

reflected in the compromised mechanical properties of bone (E=10-25 GPa, εf=1-1.5%) [11].

One could consider bone to be an interpenetrating organic-inorganic composite. The

hierarchical structure of bone must be considered when evaluating bone mechanically as the

different hierarchical, structural elements contribute distinct characteristics to mechanical

properties.

4

C)

a)

b) d)

e)

f)

g)

Length

Macro

(mm)10-3

Micro

(μm)10-6

Nano

(nm)10-9

Figure 1.1: The hierarchical organization of bone: (a) Section through femoral head showing cortical shell and trabecular interior [9]; (b) The osteon structure of cortical bone [10]; (c) A closer view of the structure of trabecular bone [10]; (d) A single trabeculae [9]; (e) A lamellar layer with an individual collagen fiber in the inset [9]; (f) The texture of a fibril showing the collagen-rich and collagen-poor regions [10]; (g) The elementary structure of bone [9].

5

1.1.2. Bone composition

In terms of composition, bone has two major phases: an organic and an inorganic phase. The

inorganic or mineral phase constitutes 60 – 70% of bone (by weight) and is poorly crystalline

hydroxyapatite (HA) crystals [Ca10(PO4)6(OH)2] with some carbonate, citrate, magnesium,

strontium and fluoride impurities [1]. The organic matrix of bone accounts for 25 – 35% of

the bone, and 90% of this phase is Type I collagen, with the remainder consisting of various

noncollagenous proteins [1].

The ratio of mineral and organic phases of bone is maintained through a dynamic process

known as bone remodeling or turnover, which is carried out by specialized bone cells. These

include osteoclasts, osteoblasts and osteocytes [1]. Osteoclasts resorb old bone through

production of hydrogen ions, which lowers the pH and consequently increases the solubility

of HA crystals. The organic matrix is then degraded via acidic proteolytic digestion [12].

Osteoblasts are bone forming cells. They line bone surfaces and produce matrix elements

including Type I collagen, osteocalcin, osteopontin, proteoglycans and regulating factors [1].

Osteocytes are osteoblasts that have been trapped within the mineralized matrix. They

maintain the bone by passing nutrients and wastes between the blood and tissue [1].

Together, osteoblasts, osteocytes and osteoclasts can create and reshape bone. If there is an

imbalance of cellular activity in bone, the bone quality will be compromised [1].

1.1.3. Bone mineral: Hydroxyapatite

Bone mineral is composed of poorly crystalline hydroxyapatite (HA). HA is calcium

phosphate [Ca10(PO4)6(OH)2] and recent studies using transmission electron microscopy

(TEM) and scanning small-angle X-ray scattering (SAXS) have shown that HA is plate-like

in shape. Fratzl et al. have determined the most probable size of the particle to be 15-200 nm

long, 10-80 nm wide and 2-5 nm thick [9]. However, their precise shape is unknown and is

still an ongoing debate: needles versus plates [9, 13]. The exact structure of bone mineral is

not well defined due to substitutions occurring into the lattice structure. For example, Na+,

K+, Fe2+, Zn2+, Sr2+and Mg2+ are capable of substituting for Ca2+ in the cationic calcium sites.

Anionic complexes (HPO42-, CO3

2-) can replace PO43- as well as OH- and F- can also

substitute OH- [14].

6

1.1.4. Bone matrix: Collagen

The organic matrix of bone consists mainly (~90%) of Type I collagen, a triple helix

molecule that is specifically arranged in several hierarchical levels to provide elasticity and

toughness to bone. The collagen molecule can be further dissected into three polypeptide

chains as shown in Figure 1.2. Each of the three polypeptide chains consists of approximately

1052 amino acids in length. Two of these chains are identical alpha-1 chains (α1(I)), while

the third is an alpha-2 chain (α2(I)). The 300 nm long and 1.5 nm thick helical structure of

the macromolecule has active sites that allow for inter- and intramolecular crosslinking [15].

The collagen molecule is stabilized by intramolecular hydrogen bonding that occurs between

hydroxyproline and other amino acid residues [15,16]. Hydroxyproline accounts for

approximately 10% of the protein by mass [17]. In fact, the primary sequence of amino acids

in collagen is a repeating Gly-X-Y unit, where X is most commonly proline and Y is most

often hydroxyproline [18,19]. Glycine is the smallest amino acid and is composed of the

smallest side group, a single hydrogen, which allows it to pack into the centre of the helix

[18,19]. Proline and hydroxyproline have very large planar side groups and due to steric

hindrance, must pack in a fairly specific fashion on the outside of the helix [18,19].

Figure 1.2: The interior structure of a collagen molecule [15].

The collagen is post-translationally modified to contain hydroxylysine, hydroxyproline and

glycosylated hydroxylysine [20]. In the extracellular matrix, the hydroxylysine residues are

involved in the formation of stable collagen crosslinks [20]. The initial crosslinks between

collagen molecules are nonstable, divalent, bipolar and reducible forms [1]. As the bone

7

matures, the crosslinks also mature into more stable nonreducible forms connecting more

than two chains [1]. Crosslinking is either enzymatically or non-enzymatically mediated [21].

The enzymatic process, mediated by lysyl oxidase, results in the trivalent collagen crosslinks

pyridinoline (PYD) and deoxypyridinoline (DPD). Non-enzymatic collagen crosslinking

(producing advanced glycation end products such as pentosidine) occurs via spontaneous

condensation of arginine, lysine and free sugars [21]. Enzymatic crosslinks are located at the

collagen overlap position, also known as the telopeptide regions [22]. Conversely, non-

enzymatic crosslinks appear to have no specific spatial arrangement [23]. Crosslinks consist

predominantly of C-C and C-N bonds within a single molecule [20] and therefore provide

local strong interactions between collagen molecules.

Tropocollagen is the subunit of collagen fibrils formed of three polypeptide strands. Each of

the three chains is twisted into a left-handed helix with approximately three amino acids per

turn [15]. The three helical α-chains are then coiled around each other into a right-handed

super-triple helix [15]. The collagen molecules are aligned such that there is a 40 nm

separation between adjacent ends, forming a collagen fibril. The collagen molecules making

up the fiber are aligned in a quarter-staggered pattern. This network of fibers provides the

framework onto which the mineral phase is deposited. The gaps between collagen fibers are

thought to be sites of mineralization initiation, making it vital to the proper formation of bone

[2,15]. Figure 1.3 shows the arrangement of collagen molecules in which the 40 nm space

between molecules is referred to as a gap region, while the 27 nm dimension refers to an

overlap region [15].

Figure 1.3: The arrangement of collagen molecules within a fibril showing the gap and overlap regions for mineral deposition [15].

8

The size of the gap region appears to constrain mineral growth. The commonly accepted

model of higher order aggregation suggests that five tropocollagen units align longitudinally

(overlapping by about one-quarter of the molecular length) into a microfibril [24].

The importance of collagen is clearly defined when considering various bone disorders.

Osteogenesis Imperfecta is a heritable disease resulting from mutations in the COL1A1 and

COL1A2 genes, which encode for the α1 and α2 chains of Type I collagen, respectively

[15,25]. Over 100 different mutations have been identified. A substitution of the glycine

residue in the α1 or α2 chain is an example. Other mutations include deletions, insertions and

duplications. As a result of these mutations, the collagen molecules are unable to form

normally, disrupting the packing of the molecules into fibrils, resulting in excessive

brittleness of bone [15,25]. The importance of the formation of crosslinks in the mechanical

functions of collagen is demonstrated in lathyrism, which inhibits lysyl oxidase activity, the

enzyme needed in crosslink formation [25]. Lathyrism results in severe abnormalities of

bones, joints and blood vessels due to decreased collagen fibril stability [25]. Finally,

dermatosparaxis is a disorder which is caused by a deficient activity of Type I procollagen N-

proteinase, the enzyme that excises the N-terminal propeptides in Type I procollagen. As a

consequence, there is accumulation of collagen that still contains the N- but not the C-

propeptide, resulting in extreme skin fragility [26].

1.1.5. Noncollagenous Proteins (NCPs)

There are over 200 types of noncollagenous proteins (NCPs) in bone, which account for

approximately 10% of the total organic bone matrix content [27]. The noncollagenous

fraction contains both specific bone proteins and serum proteins that are concentrated in bone

[28]. Osteopontin, osteocalcin, osteonectin and bone sialoprotein (BSP) are the most

abundant and widely investigated noncollagenous bone matrix proteins. They are primarily

anionic with negatively charged groups that bind readily to calcium on the surface of

hydroxyapatite [29]. Unfortunately, these acidic macromolecules are difficult to isolate from

the bone matrix and are therefore not well understood.

9

Osteopontin is the major phosphorylated glycoprotein of bone [30]. Its expression is

increased in response to increasing phosphate concentrations in soft tissues, suggesting that it

is a key regulator of mineralization [30]. Osteonectin is another widely distributed

glycosylated phosphoprotein found in mineralized tissues. Similar to the other matrix

proteins, it has been reported to enhance mineral deposition [31] and to be an inhibitor of

mineral crystal growth [32]. Osteocalcin is synthesized only by osteoblasts and it has been

shown to inhibit bone formation but has no effect on the mineralization of bone [33]. BSP

has a high affinity for calcium ions and its expression is generally limited to the later stages

of osteoblast differentiation and early stages of mineralization [34]. BSP also has an affinity

for Type I collagen and involves long-range electrostatic interactions that aid in the

formation of an initial low affinity complex, prior to the formation of a high affinity complex

characterized by specific short-range interactions [35,36].

NCPs have many roles including organization of the collagenous matrix, mediating cell

attachment, attractors of mineral ions and nucleators for their crystallization as well as

‘linkers’, in which they attach extrafibrillar mineral to the collagen fibrils and mineralized

collagen fibrils to each other [37-39]. Furthermore, Raif and Harmand suggested that the

mineral-collagen interface is formed by a complex NCP network [40].

1.1.6. Mineral-collagen interface

The mechanical properties of bone are dependent on the bonding that exists at the interface

between particles of the mineral phase and the complex molecules of the collagen phase, in

which the mineral is dispersed. However, the mineral-collagen interface is still poorly

understood.

The collagen phase is composed of individual collagen helix molecules stabilized by covalent

crosslinks [22], noncollagenous proteins, hydrogen bonds due to structural water [41] and

electrostatic interaction between molecules [42]. Hydrogen bonds are present at every two

out of three turns of the collagen helix [43]. Collagen also contains polar groups capable of

hydrogen bonding with apatite surfaces [44]. The bone apatite surface is capable of bonding

a monolayer of water by strong hydrogen bonding [44]. In fact, it has been shown that bone

10

apatite surfaces will bond polar groups more strongly than non-polar groups [44].

Furthermore, the number and orientation of the polar groups determine the strength of the

interaction adsorption on the apatite surfaces [44].

The strength of the individual bonds depends on the chemical composition of both the

mineral surface and the adjacent organic phase. A large number of bonding is available by

the extensive interface area which exists between the mineral and collagen. This interphase

bonding is due to adsorption forces between the phases [44]. It has been suggested that

interfacial bonding interactions between the mineral and the collagen are due, in part, to the

strong adsorption affinity of hydroxyapatite for organic material [45,46]. This adsorption

involves electrostatic (coulombic) interactions between the positively charged mineral

surface (calcium) and the negatively charged organic domains (carboxyl groups). Hydrogen

bonding networks, van der Walls interactions and hydrophobic bonding may also be present

at the mineral-collagen interface [45,46]. A recent study showed that the cohesion of the

mineral-collagen interface may be due to a layer of structural water [41]. Some evidence also

exists which indicates that direct covalent bonds are formed between HA and collagen [47].

However, the extent to which the adhesion between the phases is due to direct chemical

bonding, covalent or otherwise or to mechanical interlocking links between the phases is

unknown. Furthermore, as mentioned above, the noncollagenous proteins may act as an

interface phase which bonds the hydroxyapatite and collagen together [40].

1.1.7. Bone mineralization

Numerous findings and theories have been published in the literature with regard to the

mechanisms of mineralization. Mineralization has been hypothesized to occur by the

heterogeneous nucleation of calcium-phosphate nanocrystals at specific, highly ordered sites

within the collagen fibrils [48]. Chemical bonds or strong interactions between the collagen

fibrils and certain groups of the collagen matrix may create a special environment within the

fibrils that could also facilitate this nucleation [24]. Another theory involves matrix vesicles,

which are formed and released from the outer membranes of osteoblasts and related cells

[49]. It is believed that hydroxyapatite is first nucleated within the vesicle. As the crystallite

grows bigger, it breaks through the vesicle and is exposed to the extracellular fluid. However,

11

matrix vesicles are not the only site of mineral nucleation [50]. Numerous in vitro studies

have indicated that a wide variety of matrix proteins can nucleate and control growth or

agglomeration of these crystals [50]. Collagen does not seem to stimulate epitaxial growth of

hydroxyapatite and as a result, it has been considered that some of the noncollagenous

proteins might bind to the collagen fibrils to direct the nucleation event [51,52].

With regard to collagen-based mineralization, the process of bone mineral deposition begins

with the nucleation of HA crystals at multiple sites on the collagen fibrils [14,53]. A

schematic diagram illustrating the organization and interaction of collagen and mineral at

different structural levels of hierarchy is shown in Figure 1.4. Unmineralized single collagen

molecules are assembled into complex arrangements of crystal platelets, plates and lamellae

associated with collagen fibrils and fibers at macroscopic and anatomic levels [53]. Initially,

the empty gap zones between the collagen molecules are filled with water but this water is

later replaced with mineral [54]. Crystal platelets nucleate in the collagen gaps created by the

periodic (~67 nm) gap and overlap regions (Figure 1.4 (a)) [53]. This mechanism is based on

the model of collagen assembly proposed by Hodge and Petruska [55] and the manner of its

association with the hydroxyapatite mineral described by Weiner and Traub [13].

The mineral crystal platelets grow in length with a specific crystalline orientation: the c-axes

of the crystals are roughly parallel to the long axes of the collagen fibrils (Figure 1.4 (b))

[38,53,56,57]. As a result, the mineral crystals are aligned along the axis of the collagen

fibrils and reinforce the collagen matrix to provide a very strong and tough composite.

Noncollagenous proteins may bind selectively to different surfaces of the crystal, preventing

further growth and thereby determining its final size and shape [38]. Crystals then coalesce

into larger and thicker plates, maintaining their periodic deposition (~50-70 nm) and parallel

nature, as collagen macromolecules grow into microfibrils and fibrils (~20 nm in diameter)

(Figure 1.4 (c)) [53]. These crystal plates continue to grow at the level of collagen fibers (~80

nm in diameter) (Figure 1.4 (d)) [53]. Next, fibers combine, resulting in a series of parallel

plate aggregates (length of ~500 nm and thickness of ~80 nm), initially separated by ~50 nm

(Figure 1.4 (e)) [53]. However, this space decreases gradually as mineral deposition

continues in the tissue and expands to thicknesses of ~130 nm (Figure 1.4 (f)) [53]. The

12

periodic deposition (~50-70 nm) is still present, indicative of the basic collagen structure

underlying mineral formation [53]. Finally, these plate aggregates grow to be lamellar in

shape and comprise a portion of bone or mineralized tendon (Figure 1.4 (f)) [53].

Independent of the mineralization associated with the gap and overlap regions, there is

surface mineralization of the collagen structures in Figure 1.4 (b) to (f) [58].

Figure 1.4: Organization and interaction of collagen and mineral at different structural levels of hierarchy [53].

Mineralization is a two-stage process that is generally referred to as primary mineralization

and secondary mineralization. During primary mineralization, nucleation occurs in the gap

region after which the mineral grows into the overlap region. This primary mineralization

takes only a few days and accounts for 70% of the mineralization while the remaining 30% is

due to secondary mineralization, which involves further mineral growth and takes years [59].

1.2. How to Affect Bone Collagen?

Several studies have affected collagen and bone collagen using a variety of different

techniques including genetic mutations, heat denaturation, enzymatic digestion and caustic

solutions [60-75]. Naturally occurring mutations involving collagen genes are associated

with skeletal fragility in humans [60]. Specifically, the Mov13 strain carries a provirus that

13

prevents initiation of transcription of the α1(I) collagen gene [60,63] and has been used as a

model of Osteogenesis Imperfecta [60]. The collagen molecular structure has also been

altered in demineralized human cadaveric bone samples by heat induced unwinding and

enzymatic cleavage (pancreas elastase) [62]. Enzymatic digestion to denature the triple-

helical structure with the use of trypsin, cathepsin K and papain have also been used

however, a demineralization step is required to be effective [64].

The use of caustic solutions such as potassium hydroxide (KOH) and sodium hydroxide

(NaOH) as degradation agents have been reported [65-68], although to a lesser extent than

sodium hypochlorite (bleach) [69,70]. Most studies have employed a bleach concentration of

5.25%, although concentrations of 25% were used in some studies. It has been reported that

immersion of bone in commercial bleach solution results in removal of >97.5% of the

collagen from bone matrix [70]. However, most of the studies employing bleach to affect

bone collagen used bone masses on the order of a few hundred milligrams, which may have

caused rapid collagen degradation. The first deproteinization procedure of bone involved the

use of hydrazine and was proposed by Termine et al. [71]. However, this procedure has been

shown to be inefficient in eliminating the organic matrix of bone [72] and the temperature of

55 °C used in the procedure has been shown to alter bone mineral crystallinity [73]. The use

of hydrogen peroxide as a deproteinization agent for bone is not as widespread as in the case

of sodium hypochlorite and is not as effective [74]. Furthermore, it has been shown that both

the use of bleach and hydrogen peroxide as bone deproteinization agents are more effective if

the fat content is removed [74]. Treatment can also be performed using guanidine

hydrochloride, which results in removal of approximately 35% noncollagenous proteins and

small amounts of soluble collagens from bone through dissociative extraction [75]. The

proteins dissociated are either unprotected by or loosely associated with the bone matrix. The

remaining organic components consist of non-soluble collagens, mainly Type I collagen,

bone mineral and noncollagenous proteins bound to collagen and/or bone mineral [75].

1.3. Bone Mechanics

The integrity of collagen molecules is critical to the structural stability of the collagen

network and the collagen matrix plays an important role in bone fragility and fracture risk

14

[62,76]. Given its role in the structural integrity of bone, the paucity of literature devoted to

its effect on bone quality can only be attributed to the assumption that evaluating changes in

the mineral phase of bone is of more use in fracture prediction [77]. It is widely accepted that

bone strength and stiffness can be ascribed primarily to its mineral content. However,

collagen’s role in increasing bone toughness is gaining more prominence as the changes

brought on by osteoporosis and other disorders are known to affect the bone organic matrix

[76-78].

The mechanical testing of bone provides a number of parameters that provide measures of

the structural integrity of bone. Bone strength depends upon both macro- and

microarchitectural characteristics. The primary macroarchitectural components are bone

length, diameter and cortical thickness. Additionally, the cross-sectional shape and

distribution of bone mass within this cross-section are important [79]. At the

microarchitectural level, fiber and crystal alignment along with trabecular spacing,

connectivity, and alignment are the important components that determine bone strength

[80,81].

In normal loading conditions, bones are subjected to either bending, torsion, compression or

tension forces and often a combination [82]. Trabecular and cortical bone will have different

responses to different forces. For example, trabecular bone is stronger in compression than in

bending due to the structural arrangement of struts, which provide higher axial strength [82].

In mechanical testing, a bone is loaded at a specified rate until failure. The deformation of the

bone can be plotted as a function of load, which provides a failure deformation curve (Figure

1.5). From such a curve, we can derive information including failure load, failure

displacement, stiffness and energy to failure. These are subsequently combined with bone-

specific geometric parameters to obtain a stress-strain curve. This provides material

parameters including failure stress (strength), failure strain, normalized energy to failure

(toughness) and the elastic modulus (rigidity) [83] (Figure 1.6).

The ultimate load represents the maximum force applied to the specimen, while failure load

represents the force at which the specimen actually failed or fractured. For a given bone

15

sample, the ultimate and failure loads are often equivalent. The failure displacement

represents the maximum deformation achieved prior to fracture and is related to the ductility

of the specimen. The slope of the linear region of this curve is termed the extrinsic stiffness

or rigidity of the specimen. Stiffness represents elastic behaviour, meaning that the specimen

will return to its original shape if a load is applied and removed within this region. The

transition from the elastic region to the plastic region is noted by the yield point. Finally, the

area under the load-displacement curve up to the point of failure is termed the energy to

failure and represents the amount of energy required to cause failure.

To eliminate effects caused by geometric or size-dependent differences between specimens,

it is necessary to normalize the load-displacement curve. Following normalization, the

ultimate and failure load are represented by the ultimate and failure strength, respectively.

Failure displacement is represented by failure percent strain. The slope of the linear region is

termed the elastic or Young’s modulus and is a measure of the intrinsic stiffness of the

material. The yield point is defined as the intersection point between the stress-strain curve at

a line drawn parallel to the linear portion of the stress-strain curve and offset by 0.2% strain

[83]. Finally, the area under the stress-strain curve is termed “measure of toughness” or

simply “toughness”, which also represents the amount of energy required to cause failure per

unit volume of bone.

Figure 1.5: Ideal load-displacement curve.

Displacement (mm)

Loa

d (N

)

Stiffness (N/mm)

Yield Point

Failure Load (N)

Failure Displacement (mm)

Energy to Failure (mJ)

Ultimate Load (N)

16

Figure 1.6: Ideal stress-strain curve.

1.4. Fatigue

Fatigue is a common cause of failure in metals as well as other materials. It is the progressive

and permanent structural damage caused by repetitive loading at a load under the materials

yield strength. The material strength is decreased, allowing failure to occur at a lower stress

than normal [84]. Initially, there is some plastic deformation that occurs, followed by

microcrack formation, microcrack coalescence and final failure. Crack initiation is highly

dependent on surface defects and stress concentrators. A critical parameter in fatigue testing

is the fatigue life of the material, which is determined by creating a curve of stress vs.

number of cycles or S/N curve (Figure 1.7). This curve can be obtained by performing cyclic

tests until fracture of numerous samples at various stress levels. The stress level at which the

sample fractures is recorded and plotted against the log number of cycles it took to fracture.

For some materials, there exists a fatigue endurance limit, which is a stress level below

which no failure will occur. For these materials, a fatigue endurance limit can be defined as

the stress level at 106 cycles [11,85].

Percent Strain (%)

Stre

ss (M

Pa)

Elastic Modulus (MPa)

Ultimate Stress (MPa)

Failure Strength

(MPa)

Failure Percent Strain (%)

“Toughness” (J/mm3)

Yield Point

17

Number of Cycles to Failure (N)

Stre

ss (M

Pa)

Figure 1.7: Schematic diagram of stress versus number of cycles to failure curve.

1.4.1. Bone fatigue

Although any bone will fracture if exposed to excessive forces, fragility fractures result from

minor or negligible trauma. Cyclic loading, which compromises the mechanical integrity of

the bone, may also induce stress fractures [86]. Fatigue is the accumulated damage that

results from cyclic loading at physiological stresses and strains [87]. Bone fatigue induces a

reduction in stiffness that could lead to fragility and fracture [88]. It is well known that

loading of the bone introduces small amounts of microdamage, which triggers a balanced

remodeling cycle in healthy bone [89-91]. This remodeling allows bone to remove old,

damaged bone and replace it with new, healthy bone. It has been suggested that microdamage

to the bone triggers a positive feedback mechanism: microdamage in bone acts as a stimulus

for bone remodeling and remodeling repairs microdamage [92]. However, during fatigue

damage, the remodeling cycle is stimulated, but cannot repair all of the damage caused by

fatigue. Little is known about the difference between the balanced repair response and the

remodeling response to repair microdamage in bone, which can cause failure.

Bone adapts its mass, architecture and mechanical properties in response to mechanical

loading. While bone is protective against impact, bone is also susceptible to fatigue, a

Fatigue endurance limit

18

process by which repetitive loading damages the bone matrix. This repetitive, smaller loading

of bone leads to microcrack formation and accumulation [88]. The fatigue behaviour of bone

resembles that of composite materials, exhibiting a gradual loss of stiffness and strength

throughout cyclic loading due to fatigue damage accumulation [88]. The stages that bone

undergoes during fatigue loading are shown in Figure 1.8 [93]. First, there is a fast decrease

in stiffness, followed by a stable rate change of stiffness and finally catastrophic failure, as

depicted in Figure 1.8 (a) [93]. Figure 1.8 (b) illustrates the development of damage during

fatigue loading: diffuse damage associated with intralamellar structural discontinuities within

the matrix combined with microcracks formed by interlamellar debonding may coalesce,

forming marcocracks and culminating to catastrophic failure [93]. Finally, the measured

displacement during the entire loading test is shown in Figure 1.8 (c) [93].

19

Figure 1.8: Schematic diagram of the stages of fatigue in bone: (a) showing the stages of fatigue; (b) schematic of the damage to the bone; (c) change in the bone compliance [93].

The fatigue life is a function of crack initiation and crack propagation. Materials where

cracks are easily initiated but difficult to grow often show greater resistance to fatigue failure

as opposed to materials where crack initiation is slow, but propagation is quick [94]. Bone

can be compared to a composite material where discontinuities in the material may provide

stress concentrations. Stress concentrations are prime sights for crack initiation.

Discontinuities in osteonal bone appear in the form of fibers, lamellae and pores and may

provide stress concentration sites for crack initiation [95]. Microcracks seem to develop in

the interstitial regions of bone and stop at the osteonal boundary [96]. It has been suggested

that 80-90% of all microcracks in cortical bone are found in the interstitial bone between

osteons [97].

a)

b)

c)

20

1.4.2. Bone creep strain

Bone accumulates damage over time and eventually fails below its strength. When there is

time-evolving damage, creep rupture may occur. Creep is the gradual increase in material

strain over time at a constant stress [98]. Bones not only undergo cyclic loading but also

constant stress such as standing for a period of time. Therefore, cyclic loading cannot be

studied exclusively since there is an interaction between fatigue and creep. Creep is a known

characteristic of materials such as polymers [98]. In bone, collagen is abundant and displays

polymeric properties [93]. It has been suggested that the collagen component of bone is

responsible for its observed creep behaviour [99,100]. Since collagen plays a major role in

bone biomechanics [61,76,77], it is important to identify the role(s) that collagen may have in

fatigue fractures.

1.5. Bone Toughening Mechanisms

Bone has several crack-stopping (toughening) mechanisms, such as fiber bridging, crack

deflection and microcracks that deflect, slow or stop crack propagation and increase the

resistance to fracture and ultimately failure (Figure 1.9) [101-104]. Collagen fibrils can act as

a toughening mechanism, whereby intact fibers bridge a crack and oppose crack opening

[101,103]. Similarly, uncracked ligaments may serve to bridge opposing sides of a crack,

consequently blunting further crack propagation along the crack path, a phenomenon known

as ‘crack bridging’ [105,106]. Cement lines and interlamellar boundaries are believed to

provide weak interfaces capable of crack deflection, thereby prolonging the crack

propagation path and accordingly increasing bone toughness [101,107,108]. Finally, Nyman

and his group reviewed the crack tip shielding mechanism in which accumulation of damage

in front of the crack acts as a crack-stopping mechanism [15]. They suggest that when

microdamage coalesces, it forms a propagating crack (mother crack). This linear crack

accelerates but as it does this, microdamage (daughter cracks) forms at the tip in order to

absorb the increased energy. As a result, crack propagation decreases but this is only

temporary since the accumulation of microdamage eventually initiates the process again.

Therefore, microdamage from microcracking works through crack-tip shielding [101].

21

Microcracking Uncrackedligament bridging

Crack bridging Crack deflection

(a) (b)

(c) (d) Figure 1.9: Schematic diagram of bone toughening mechanisms: (a) microcracking; (b) uncracked ligament bridging; (c) crack bridging by collagen fibers and (d) crack deflection by osteons [109].

1.6. Tools for Fracture Risk Assessment

The nature of mechanical tests does not permit their use in vivo to assess bone quality, as a

result, surrogate methods of predicting the mechanical properties of bone have been

developed. Two of the most common methods of fracture prediction are Dual Energy X-ray

Absorptiometry (DXA) and Quantitative Ultrasound (QUS). Both DXA and QUS have

gained widespread use in clinical practice however, neither technique has demonstrated the

ability to adequately distinguish between fracture and non-fracture populations [110]. These

indirect measurements of fracture risk are only two of many contributors to bone strength and

fracture risk. Bone strength is derived from bone quantity, which consists of density and size

as well as bone quality, which consists of structure, material properties and bone turnover

[111]. Bone fragility depends not only on mineral content but also on matrix properties,

architecture and geometry [111] and therefore evaluating the mineral content of bone alone is

insufficient to predict changes in bone quality [86]. Consequently, an instrument known as

the Mechanical Response Tissue Analyzer (MRTA), which provides a direct measure of a

mechanical property of bone, is being explored as a new tool that may provide a more

effective means of fracture prediction.

22

1.6.1. Dual Energy X-ray Absorptiometry (DXA)

DXA provides a measure of the areal bone mineral density (BMD), reported in g/cm2 by

dividing the measured bone mineral content (BMC) of a specific region by the area of the

region. This differs from a volumetric density in that the region of interest is the two-

dimensional area projected by the object to be analyzed. The instrument itself consists of a

table or flat platform on which the object to be measured is placed, below which an x-ray

source generates an incident beam that passes through the object and is measured by

detectors above [110]. The beam is attenuated by passage through the object and it is this

degree of attenuation that provides the measure of bone mineral density. As is suggested by

the name, the incident beam consists of two energy levels (140 kVp and 70 kVp). Soft tissue

attenuates both energy levels equally, whereas bone mineral attenuates the lower energy

beam to a much greater extent. Thus, by subtracting the profile of the high energy beam from

the low energy beam, regions containing only soft tissue exhibit a nearly zero value.

However, regions containing bone mineral will yield a non-zero value from the subtraction of

the profiles, indicating the presence of bone [110,112].

DXA is currently the standard in bone quality assessment. While it does not directly measure

a mechanical property of bone, studies have shown that BMD values strongly correlate with

fracture risk [113]. However, fracture and non-fracture populations often have significant

overlap in BMD values and changes in fracture risk resulting from therapy are often not

reflected in DXA results [110]. Also, due to the fact that the BMD determined by DXA is an

areal measure, the thickness of bone is not considered, which can result in misleading values

when particularly large bones are measured [110]. The final major disadvantage to DXA

measurement is the requisite radiation dose needed for measurement. While this dose is

relatively small, it does restrict some patients from measurement [113,114].

1.6.2. Quantitative Ultrasound (QUS)

QUS is a more recent development in bone quality assessment. This instrument measures the

speed of sound (SOS) applied axially or transversely through the bone. SOS in bone is not a

mechanical parameter, but it is related to parameters that contribute to mechanical integrity

23

[110,115,116]. The velocity of sound (v) traveling through solid matter can be related to the

modulus of elasticity using the following equation:

E = ρv2

where E is the modulus of elasticity and ρ is the volumetric material density. This

relationship holds for sound traveling through bone:

E = ρ*SOS2

A hand-held probe containing two piezoelectric transducers is placed on the measurement

site and is put into acoustic contact with the site through ultrasound gel. A high frequency

(1.25 MHz) acoustic signal is generated at one transducer and is received at the opposing

transducer. Since the distance between the transducers is known and fixed, the SOS can be

computed. The signal is known to travel by the fastest possible means, which indicates that

the sound will travel through cortical bone as opposed to the overlying soft tissue or

trabecular bone. Proprietary software from the manufacturers of these instruments claims to

eliminate the effects of the surrounding tissues [110,112]. QUS data has been claimed to be a

predictor of bone strength however, in clinical practice, this device is more commonly used

as a pre-screening tool [110].

1.6.3. Mechanical Response Tissue Analyzer (MRTA)

The MRTA is a radiation-free, non-invasive instrument developed by NASA to investigate

the effects of space travel on astronaut bones [110,117]. Unlike DXA and QUS, it directly

measures a bone mechanical property, the cross-sectional bending stiffness (EI) of long

bones. EI is the product of the elastic modulus, E, and the areal cross sectional moment of

inertia, I [110,117]. This is done by placing the bone in a three-point bending configuration

and applying a low frequency (0 to 1600 Hz) vibration to the skin surface (which is

transmitted to the bone) using an electromagnetic shaker with an impedance head probe

(Figure 1.10). A transducer on the probe measures the force and acceleration response from

the bone, which are used to calculate EI. The cross-sectional bending stiffness of a long bone

with the load (F) applied to the midspan can be calculated using the following equation:

24

where k is the lateral stiffness of a beam in three-point bending, δmax is the maximum bending

displacement and L is the length of the long bone.

Probe Shaker

Impedance Head

Figure 1.10: MRTA setup for an ulna measurement [110].

More specifically, the basic components of the MRTA are the mechanical shaker, the

impedance head, the contact probe, the system control and measurement software and the

analysis software. The mechanical shaker, impedance head and contact probe are connected

in series with the shaker physically driving the impedance head-probe with a randomly

generated frequency from 0 to 1600 Hz, as specified by the controlling software. The

impedance head transmits the dynamic response of the bone (both force and acceleration

data) to the measurement software and the probe (<1 cm2) provides contact with the bone at

the mid-point of either the ulna or tibia. Once contact with the bone is established,

measurements can be taken where each measurement takes approximately five seconds.

Real-time data is converted with Fourier transforms to the frequency domain for analysis. A

software algorithm, developed by Gaitscan and NASA, uses the dynamic response to

calculate the lateral stiffness (k) to solve for EI [110,117].

orLEIFk ,483

max

==δ 48

3kLEI =

25

To evaluate the functional stiffness at the level of the bone and not just the material level,

Young’s modulus is coupled with the cross-sectional moment of inertia (I) to yield the cross-

sectional bending stiffness, EI [83]. The cross-sectional moment of inertia for a

homogeneous material clearly depends upon the simple geometry of the structure, but

heterogeneous materials such as bone are also influenced by the quantity, distribution and

orientation of bone mineral and collagen fibers around the bending axis [80]. The effective

stiffness of bone is also influenced by trabecular connectivity, alignment and number, as well

as the alignment of bone mineral and collagen fibers. These complex factors are difficult, if

even possible to quantify with non-invasive measures, so the ability to directly measure the

cross-sectional bending stiffness (EI) avoids this problem.

Two major issues can complicate MRTA assessment of bone quality. Firstly, the overlying

skin and soft tissue can partially absorb the applied vibration before it is transmitted to the

bone and the muscle also serves to dampen the response of bone to the vibration. These

effects must be accounted for in the software that evaluates the data received by the probe

[110]. Secondly, the mathematical model used in the calculation of EI, models the bone as a

uniform cylinder. Therefore, measurements with the MRTA are limited to the human ulna

and tibia. This assumption is often not valid for long bones that exhibit curvature.

Research has shown that EI of a long bone is predictive of the maximum strength of the bone

[117] thus, measurements of EI can be used to assess bone quality. Furthermore, previous

studies suggest that the determination of EI has the potential to effectively evaluate fracture

risk [117-119]. Studies have demonstrated the ability of the MRTA to detect EI differences

between various populations including exercise [120,121], disease states [122,123] and age

[124]. The effectiveness of the MRTA in predicting fractures has not as of yet, been

evaluated, but the fact that it provides a direct measure of the extrinsic stiffness of bone may

permit it to provide more accurate fracture prediction [110]. Establishing the MRTA as a

clinical tool to replace DXA or QUS is not the goal of this study, but the justification for its

inclusion in this study is its strength as a research tool.

26

1.7. Animal Models – Emu

Due to the difficulty in obtaining human samples at various stages of disease for analysis,

animal models are widespread in biomedical research. The biomechanical tests chosen often

dictate what animal model can be used [82]. In this investigation, the need for a bone sample

of suitable size for analysis by the MRTA restricted the available options for animal models.

Traditional models such as mice, rats, rabbits and dogs are too small to provide such bones

and so the emu (Dromaeus novaehollandie) was selected as a model. The emu is a large,

ostrich-like animal weighing approximately 50-60 kg once mature [125-127]. The tibia of the

emu is of particular interest as their size and approximate cylindrical shape are similar to

human long bones. Emu tibiae are also large enough to be assessed by the MRTA.

Furthermore, sexual dimorphism is apparent in emus where females are notably heavier and

larger compared to males [126]. Finally, even though female emus lay down eggs, it is the

male emus that carry out the 8-week incubation period [128]. These dimorphisms can be

used to evaluate differences between female and male bones. Emu tibiae may also be a

suitable model for human long leg bones as the bipedal nature of emus is analogous to human

locomotion. Lastly, emus are available from slaughterhouses and farms in Ontario due to the

demand for their meat. These factors together, promote the continued characterization of this

animal model for further study in bone research.

1.8. Objectives

The overall goal of this study is to evaluate the consequences of collagen degradation on the

mechanical and fatigue properties of female and male bone, using an emu model. This will

be achieved by endocortically treating emu tibiae with 1 M potassium hydroxide (KOH) for

various time periods. KOH was chosen as the degradation agent as it has been used

previously to affect collagen [68] as well as bone tissue [65-67] and does not require a

demineralization step. In addition, the need for a rapid and cost effective protocol on whole

emu bone samples that would alter bone collagen endocortically and still allow for

mechanical testing thereafter, further supported the use of KOH as a collagen degradation

agent. While this treatment is not physiological, it is a necessary step to understand the

mechanisms by which collagen degradation affects bone mechanical properties. This study

will also test the MRTA as a potential research tool for the assessment of bone quality of

27

both fatigued and treated bone. To accomplish this overall goal, the study can be divided into

five objectives:

1) To investigate the role of collagen on female and male emu bone mechanical

properties and to assess if DXA, QUS and MRTA can detect these changes.

2) To investigate the role of collagen on the fatigue resistance of female and male emu

bone and to assess if DXA, QUS and MRTA can detect these changes.

3) To determine the failure mechanisms of emu bone.

4) To determine the extent of collagen degradation and evaluation of bone mineral.

5) To investigate the mineral-organic matrix interface.

At the conclusion of each objective, journal articles were published or submitted for review

which outlined the experimental work and results of each objective. This thesis is a

culmination of these different journal articles, which highlight the main findings.

1.9. Hypothesis

We anticipate that KOH treated emu bones will have reduced mechanical properties and a

lower resistance to fatigue compared to untreated bones and that collagen degradation and/or

mineral-organic matrix interface alteration will be the underlying mechanisms. It is important

to also note that the endocortical KOH treatment will leave the specimens intact and will

result in an inner ring zone where collagen is significantly affected with less disruption as the

distance from the medullary canal increases. This will result in structurally heterogeneous

cross-sections. Thus, KOH treatment will create a gradient of degradation where the

mechanical properties on the inside will be different than the rest of the bone. Furthermore,

we hypothesize that the changes in mechanical and fatigue properties will be more

pronounced in male emu tibiae compared to female emu tibiae due to sexual dimorphism.

28

1.10. References

[1] Jee WSS. (2001) Integrated Bone Tissue Physiology. In: Cowin SC, editor. Bone Mechanics Handbook. CRC Press. Florida. pp. 1-68.

[2] Rho JY, Kuhn-Spearing L, Zioupos P. (1998) Mechanical properties and the hierarchical

structure of bone. Med.Eng.Phys. 20:92-102. [3] Currey JD, Brear K, Zioupos P. (1996) The effects of ageing and changes in mineral content

in degrading the toughness of human femora. J.Biomech. 29:257-260. [4] Li X, Agrawal M, Wang X. (2003) Age dependence of in situ thermostability of collagen in

human bone. Calcif.Tissue.Int. 72:513-518. [5] Bundy KJ. (1985) Determination of mineral-organic bonding effectiveness in bone--

theoretical considerations. Ann.Biomed.Eng. 13:119-35. [6] Walsh WR, Ohno M, Guzelsu N. (1994) Bone composite behaviour: effects of mineral-

organic bonding. J.Mater.Sci.Mater.Med. 5:72-9. [7] Kotha SP, Guzelsu N. (2000) The effects of interphase and bonding on the elastic modulus of

bone: changes with age-related osteoporosis. Med.Eng.Phys. 22:575-85. [8] Gupta HS, Fratzl P, Kerschnitzki M, Benecke G, Wagermaier W, Kirchner HO. (2007)

Evidence for an elementary process in bone plasticity with an activation enthalpy of 1 eV. J.R.Soc.Interface. 4:277-82.

[9] Fratzl P, Gupta HS, Paschalis EP, Roschger P. (2004) Structure and mechanical quality of the

collagen-mineral nano-composite in bone. J.Mater.Chem. 14:2115-2123. [10] Hellmich C, Barthelemy JF, Dormieux L. (2004) Mineral-collagen interactions in elasticity of

bone ultrastructure – a continuum micromechanics approach Eur.J.Mech.A.Solids. 23:783-810.

[11] Currey JD. (1984) The mechanical adaptation of bones. Princeton University Press. New

Jersey. pp. 80-85. [12] Kaplan FS, Hayes WC, Keaveny TM, Boskey A, Einhorn TA Iannotti JP. (1994) Form and

Function of Bone. In: Simon SR, editor. Orthopaedic Basic Science. American Academy of Orthopaedic Surgeons. pp. 127-184.

[13] Weiner S, Traub W. (1992) Bone structure: from angstroms to microns. Faseb.J. 6:879-885. [14] Boskey A. (2003) Bone mineral crystal size. Osteoporos.Int. 14:S16-S21. [15] Nyman JS, Reyes M, Wang X. (2005) Effect of ultrastructural changes on the toughness of

bone. Micron. 36:566-582. [16] Wang X, TeKoppele JM, Athanasiou KA. (2000) Effect of collagen denaturation on the

toughness of bone. Clin.Orthop.Relat.Res. 371:228-239.

29

[17] Woessner JF Jr. (1976) Determination of hydroxyproline in connective tissues. In: Hall D, editor. The methodology of connective tissue research. Joynson-Bruvvers Ltd. Oxford. pp. 227-233.

[18] Miller EJ. (1984) Chemistry of the collagens and their distribution. In: Piez K, Reddi A,

editors. Extracellular Matrix Biochemistry. Elsevier. New York. pp. 40-81. [19] Burjanadze TV, Veis A. (1997) A thermodynamic analysis of the contribution of

hydroxyproline to the structural stability of the collagen triple helix. Connect.Tissue.Res. 36:347-365.

[20] Knott L, Bailey AJ. (1998) Collagen cross-links in mineralizing tissues: a review of their

chemistry, function, and clinical relevance. Bone. 22:181-187. [21] Bailey AJ, Paul RG, Knott L. (1998) Mechanisms of maturation and ageing of collagen.

Mech.Age.Dev. 106:1-56. [22] Eyre DR, Wu JJ. (2005) Collagen cross-links. Top.Curr.Chem. 247:207-229. [23] Siegmund T, Allen MR, Burr DB. (2008) Failure of mineralized collagen fibrils: Modeling

the role of collagen cross-linking. J.Biomech. 41:1427-1435. [24] Glimcher MJ. (2006) Bone: Nature of the calcium phosphate crystals and cellular, structural,

and physical chemical mechanisms in their formation. Rev.Min.Geo. 64:223-282. [25] Viguet-Carrin S, Garnero P, Delmas PD. (2006) The role of collagen in bone strength.

Osteoporos.Int. 17:319-336. [26] Smith LT, Wertelecki W, Milstone LM, Petty EM, Seashore MR, Braverman IM, Jenkins

TG, Byers PH. (1992) Human dermatosparaxis: a form of Ehlers-Danlos syndrome that results from failure to remove the amino terminal propeptide of type I procollagen. Am.J.Hum.Genet. 51:235-244.

[27] Roach HI. (1994) Why does bone matrix contain non-collagenous proteins? The possible

roles of osteocalcin, osteonectin, osteopontin and bone sialoprotein in bone mineralisation and resorption. Cell.Biol.Int. 18:617-628.

[28] Triffitt JT, Own MO. (1977) Preliminary studies on the binding of plasma albumin to bone

tissue. Calcif.Tissue.Res. 23:303-305. [29] Boskey AL. (1992) Mineral-matrix interactions in bone and cartilage. Clin.Orthop.Relat.Res.

281:244-274. [30] Butler WT. (1989) The nature and significance of osteopontin. Connect.Tissue.Res. 23:123-

136. [31] Termine JD, Kleinman HK, Whitson SW, Conn KM, McGarvey ML, Martin GR. (1981)

Osteonectin, a bone-specific protein linking mineral to collagen. Cell. 26:99-105. [32] Romberg RW, Werness PG, Riggs BL, Mann KG. (1986) Inhibition of hydroxyapatite crystal

growth by bone-specific and other calcium-binding proteins. Biochem. 25:1176-1180.

30

[33] Ducy P, Desbois C, Boyce B, Pinero G, Story B, Dunstan C, Smith E, Bonadio J, Goldstein S, Gundberg C, Bradley A, Karsenty G. (1996) Increased bone formation in osteocalcin-deficient mice. Nature. 382:448-452.

[34] George A, Veis A. (2008) Phosphorylated proteins and control over apatite nucleation, crystal

growth, and inhibition. Chem.Rev. 108:4670-4693. [35] Fujisawa R, Nodasaka Y, Kuboki Y. (1995) Further characterization of interaction between

bone sialoprotein (BSP) and collagen. Calcif.Tissue.Int. 56:140-144. [36] Tye CE, Hunter GK, Goldberg HA. (2005) Identification of the type I collagen-binding

domain of bone sialoprotein and characterization of the mechanism of interaction. J.Biol.Chem. 280:13487-13492.

[37] Gundberg CM. (2003) Matrix proteins. Osteoporos.Int. 14:S37-S42. [38] Hunter GK, Goldberg HA. (1993) Nucleation of hydroxyapatite by bone sialoprotein.

Proc.Natl.Acad.Sci.USA. 90:8562-8565. [39] Goldberg HA, Hunter GK. (1995) The inhibitory activity of osteopontin on hydroxyapatite

formation in vitro. Ann.NY.Acad.Sci. 760:305-308. [40] Raif EM, Harmand MF. (1993) Molecular interface characterization in human bone matrix. I.

Biochemical and IR spectroscopic studies. Biomater. 14:978-984. [41] Wilson EE, Awonusi A, Morris MD, Kohn DH, Tecklenburg M, Beck LW. (2005) Highly

ordered interstitial water observed in bone by nuclear magnetic resonance. J.Bone.Min.Res. 20:625-634.

[42] Suzuki Y, Someki I, Asachi E, Irie S, Hattori S. (1999) Interaction of collagen molecules

from the aspect of fibril formation: acid-soluble, alkali-treated, and MMP1-digested fragments of Type I collagen. J.Biochem. 126:54-67.

[43] Steiner T. (2002) The hydrogen bond in the solid state. Angew.Chem.Int.Ed.Engl. 41:48-76. [44] Posner AS. (1985) The structure of bone apatite surfaces. J.Biomed.Mater.Res. 19:241-250. [45] Pearce EIF. (1981) Ion displacement following the adsorption of anionic macromolecules on

hydroxyapatite. Calcif.Tissue.Int. 33:395-402. [46] Walsh WR, Guzelsu N. (1994) Compressive properties of cortical bone: mineral-organic

interfacial bonding. Biomater. 15:137-145. [47] Posner AS. (1969) Crystal chemistry of bone mineral. Phys.Rev. 49:760-792. [48] Glimcher MJ. (1959) Molecular biology of mineralized tissues with particular references to

bone. Rev.Mod.Phys. 31:359-393. [49] Anderson HC. (1995) Molecular biology of matrix vesicles. Clin.Orthop.Relat.Res. 314:266-

280.0

31

[50] Boskey AL. (1996) Matrix proteins and mineralization: overview. Connect.Tissue.Res. 35:357-363.

[51] Traub W, Jodaikin A, Arad T, Veis A, Sabsay B. (1992) Dentin phosphophoryn binding to

collagen fibrils. Matrix. 12:197-201. [52] Glimcher MJ. (1984) Recent studies of the mineral phase in bone and its possible linkage to

the organic matrix by protein-bound phosphate bonds. Philos.Trans.R.Soc.Lon.(Biol). 304:479-506.

[53] Siperko LM, Landis WJ. (2001) Aspects of mineral structure in normally calcifying avian

tendon. J.Struct.Biol. 135:313-320. [54] Elliott SR, Robinson RA. (1957) The water content of bone. I. The mass of water, inorganic

crystals, organic matrix, and CO2 space components in a unit volume of the dog bone. J.Bone.Joint.Surg.Am. 39:167-188.

[55] Hodge AJ, Petruska JA. (1963) Recent studies with the electron microscope on ordered

aggregates of the tropocollagen molecule. In: Ramachandran GN, editor. Aspects of Protein Structure. Academic Press. London. pp. 289-300.

[56] Landis WJ, Song MF, Leith A, McEwen L, McEwen BF. (1993) Mineral and organic matrix

interaction in normally calcifying tendon visualized in three dimensions by high-voltage electron microscopic tomography and graphic image reconstruction. J.Struct.Biol. 110:39-54.

[57] Arsenault AL. (1988) Crystal-collagen relationships in calcified turkey leg tendons visualized

by selected-area dark field electron microscopy. Calcif.Tissue.Int. 43:202-212. [58] Landis WJ. Hodgens KJ, Song MJ, Arena J, Kiyonaga S, Marko M, Owen C, McEwen BF.

(1996b) Mineralization of collagen may occur on fibril surfaces: Evidence from conventional and high voltage electron microscopy and three-dimensional imaging. J.Struct.Biol. 117:24-35.

[59] Erikson EF, Axelrod DW, Melson F. (1994) Bone Histomorphometry. New York: Raven

Press. [60] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

[61] Oxlund H, Barckman M, Ortoft G, Andreassen TT. (1995) Reduced concentrations of

collagen cross-links are associated with reduced strength of bone. Bone. 17:365S-371S. [62] Wang X, Li X, Bank RA, Agrawal CM. (2002) Effects of collagen unwinding and cleavage

on the mechanical integrity of the collagen network in bone. Calcif.Tissue.Int. 71:186-192. [63] Lohler J, Timpl R, Jaenish R. (1984) Embryonic lethal mutation in mouse collagen I gene

causes rupture of blood vessels and is associated with erythropoietic and mesenchymal cell death. Cell. 38:597-607.

32

[64] Brady JD, Robins SP. (2001) Structural characterization of pyrrolic cross-links in collagen using a biotinylated ehrlich’s reagent. J.Biol.Chem. 276:18812-18818.

[65] Arsenault AL. (1990) Vascular canals in bovine cortical bone studied by corrosion casting.

Calcif.Tissue.Int. 47:320-325. [66] Lin TC, Su CY, Chang CS. (1997) Stereomorphologic observation of bone tissue response to

hydroxyapatite using SEM with the EDTA-KOH method. J.Bone.Miner.Res. 36:91-97. [67] Abe K, Hashizume H, Ushiki T. (1992) An EDTA-KOH method to expose bone cells for

scanning electron microscopy. J.Electron.Microsc.(Tokyo). 41:113-115. [68] Privalova LG, Konstantinova ML, Kulagin VN, Zaikov GY, Sorokin YY, Dreizenshtok GS.

(1980) Kinetic regularities of the degradation of collagen in dilute solutions of sulphuric acid and potassium hydroxide. Pol.Sci. 22:2583-2590.

[69] Carter DH, Scully AJ, Heaton DA, Young MP, Aaron JE. (2002) Effect of deproteination on

bone mineral morphology: implications for biomaterials and aging. Bone. 31:389-395. [70] Broz JJ, Simske SJ, Corley WD, Greenberg AR. (1997) Effects of deproteinization and

ashing on site-specific properties of cortical bone. J.Mater.Sci.Mater.Med. 8:395-401. [71] Termine JD, Eanes ED, Greenfie DJ, Nylen MU, Harper RA. (1973) Hydrazine-

deproteinated bone mineral: Physical and chemical properties. Calcif.Tissue.Res. 12:73-90. [72] Bertazzo S, Bertran CA. (2008) Effect of hydrazine deproteination on bone mineral phase: A

critical review. J.Inorg.Biochem. 102:137-145. [73] Kim HM, Rey C, Glimcher MJ. (1995) Isolation of calcium-phosphate crystals of bone by

non-aqueous methods at low temperature. J.Bone.Miner.Res. 10:1589-1601. [74] Johnson GS, Mucalo MR, Lorier MA. (2000) The processing and characterization of animal-

derived bone to yield materials with biomedical applications Part 1: Modifiable porous implants from bovine condyle cancellous bone and characterization of bone materials as a function of processing. J.Mat.Sci.Mat.Med. 11:427-441.

[75] Guzelsu N, Walsh WR. (1990) Streaming potential of intact wet bone. J.Biomech. 23673-

685. [76] Wang X, Shen X, Li X, Agrawal CM. (2002) Age-related changes in the collagen network

and toughness of bone. Bone. 31:1-7. [77] Zioupos P, Currey JD, Hamer AJ. (1999) The role of collagen in the declining mechanical

properties of aging human cortical bone. J.Biomed.Mater.Res. 45:108-116. [78] Turner CH. (2006) Bone strength: current concepts. Ann.NY.Acad.Sci. 1068:429-446. [79] Martin RB. (1991) Determinants of the mechanical properties of bones. J.Biomech. 24:79-88. [80] Martin RB, Ishida J. (1989) The relative effects of collagen fiber orientation, porosity,

density, and mineralization on bone strength J.Biomech. 22:419-26.

33

[81] Turner CH., Forwood MR, Rho JY, Yoshikawa T. (1994) Mechanical loading thresholds for lamellar and woven bone formation. J.Bone.Miner.Res. 9:87.

[82] Turner CH, Burr DB. (2001) Experimental Techniques for Bone Mechanics. In: Cowin SC,

editor. Bone Mechanics Handbook. CRC Press. Florida. pp. 1-35. [83] Turner CH, Burr DB. (1993) Basic biomechanical measurements of bone: a tutorial. Bone.

14:595-608. [84] Fine M, Chung Y. Fatigue Fracture in Metals. ASTM Handbook Vol. 19 Fatigue and

Fracture. [85] Hastings A, Gibson LJ, Moore TLA, Cheng DW, Guo XE. (2004) Endurance limit for bovine

trabecular bone. In: Proceedings of the 50th annual meeting of the Orthopaedic Research Society. 29:34. San Francisco. CA. March 7-10.

[86] Burr DB. (2002) The contribution of the organic matrix to bone's material properties. Bone.

31:8-11. [87] Burr DB, Forwood MR, Fyhrie DP, Martin RB, Schaffler MB, Turner CH. (1997) Bone

microdamage and skeletal fragility in osteoporotic and stress fractures. J.Bone.Miner.Res. 12:6-15.

[88] Lee T.C., O'Brien F.J., Taylor D. (2000) The nature of fatigue damage in bone. Int.J.Fat.

22:847-853. [89] Pattin CA, Caler WE, Carter DR. (1996) Cyclic mechanical property degradation during

fatigue loading of cortical bone. J.Biomech. 29:69-79. [90] Hsieh YF, Silva MJ. (2002) In vivo fatigue loading of the rat ulna induces both bone

formation and resorption and leads to time-related changes in bone mechanical properties and density. J.Orthop.Res. 20:764-771.

[91] Mori S, Burr DB. (1993) Increased intracortical remodeling following fatigue damage. Bone.

14:103-109. [92] Lee TC, Staines A, Taylor D. (2002) Bone adaptation to load: microdamage as a stimulus for

bone remodelling. J.Anat. 201:437-446. [93] Tami AE, Nasser P, Schaffler MB, Knothe Tate ML. (2003) Noninvasive fatigue fracture

model of the rat ulna. J.Orthop.Res. 21:1018-1024. [94] Schaffler MB, Choi K, Milgrom C. (1995) Aging and matrix microdamage accumulation in

human compact bone. Bone. 17:521-525. [95] Jepsen KJ, Davy DT, Krzypow DJ. (1999) The role of the lamellar interface during torsional

yielding of human cortical bone. J.Biomech. 32:303-310. [96] Schaffler MB, Choi K, Milgrom C. (1994a) Microcracks and aging in human femoral

compact bone. Proc.Orthopae.Res.Soc. 19:190.

34

[97] Park HC, Lakes RS. (1986) Cosserat micromechanics of human bone – strain redistribution by a hydration sensitive constituent. J.Biomech. 19:386-397.

[98] Callister WD. (2003) Materials Science and Engineering: An Introduction. John Wiley &

Sons, New York. [99] Bowman SM, Gibson LJ, Hayes WC, McMahon TA. (1999) Results from demineralized

bone creep tests suggest that collagen is responsible for the creep behavior of bone. J.Biomech.Eng. 121:253-258.

[100] Fois M, Lamure R, Fauran MJ, Lacabanne C. (1998) Viscoelastic properties of human bone.

J.Biomech. 31:2. [101] Nalla RK, Kinney JH, Ritchie RO. (2003) Mechanistic fracture criteria for the failure of

human cortical bone. Nat.Mater. 2:164-168. [102] Zioupos P, Currey JD, Mirza MS, Barton DC. (1995) Experimentally determined

microcracking around a circular hole in a flat plat of bone: comparison with predicted stresses. Philos.Trans.R.Soc.Lond.B.Biol.Sci. 347:383-396.

[103] Yeni YN, Norman TL. (2000) Fracture toughness of human femoral neck effect of

microstructure, composition and age. Bone. 26:499-504. [104] Yan , Mecholsky Jr. JJ, Clifton KB. (2007) How tough is bone? Application of elastic-plastic

fracture mechanisms to bone. Bone. 40:479-484. [105] Vashishth D, Behiri JC, Bonfield W. (1997) Crack growth resistance in cortical bone:

concept of microcrack toughening. J.Biomech. 30:763-769. [106] Vashishth D, Tanner KE, Bonfield W. (2000) Contribution, development and morphology of

microcracking in cortical bone during crack propagation. J.Biomech. 33:1169-1174. [107] Yeni YN, Norman TL. (2000) Calculation of porosity and osteonal cement line effects on the

effective fracture toughness of cortical bone in longitudinal crack growth. J.Biomed.Mater.Res. 51:504-509.

[108] Feng QL, Cui FZ, Pu G, Wang RZ, Li HD. (2000) Crystal orientation toughening

mechanisms and a mimic of nacre. Mat.Sci.Eng.C. 11:19-25. [109] Ritchie RO, Kinney JH, Kruzic JJ, Nalla RK. (2005) A fracture mechanics and mechanistic

approach to failure of cortical bone. Fat.Fract.Engng.Mater.Struct. 28:345-371. [110] Djokoto C, Tomlinson G, Waldman S, Grynpas M, Cheung AM. (2004) Relationship among

MRTA, DXA, and QUS. J.Clin.Densitom. 7:448-456. [111] Cefalu CA. (2004) Is bone mineral density predictive of fracture risk reduction?

Curr.Med.Res.Opin. 20:341-349. [112] Fogelman I, Blake GM. (2000) Different approaches to bone densitometry. J.Nucl.Med.

41:2015-25.

35

[113] Genton L, Hans D, Kyle UG, Pichard C. (2002) Dual energy x-ray absorptiometry and body composition: Differences between devices and comparison with reference methods. Nutrition. 18:66-70.

[114] Bonnick SL, Lewis LA. (2002) Bone densitometry for technologists. Human Press. New

Jersey. [115] Greenfield MA, Craven JD, Huddleston A. Kehrer ML, Wishko D, Stern R. (1981)

Measurement of the velocity of ultrasound in human cortical bone in vivo. Estimation of its potential value in the diagnosis of osteoporosis and metabolic bone disease. Radiology. 138:701-710.

[116] Hodgskinson R, Njeh CF, Currey JD Langton CM. (1997) The ability of ultrasound velocity

to predict the stiffness of cancellous bone in vitro. Bone. 21:183-190. [117] Roberts SG, Hutchinson TM, Arnaud SB, Kiratli BJ, Martin RB, Steele CR. (1996)

Noninvasive determination of bone mechanical properties using vibration response: a refined model and validation in vivo. J.Biomech. 29:91-98.

[118] Hutchinson TM, Bakulin AV, Rakhmanov AS, Martin RB, Steele CR, Arnaud SB. (2001)

Effects of chair restraint on the strength of the tibia in rhesus monkeys. J.Med.Primatol. 30:313-321.

[119] Norrdin RW, Simske SJ, Gaarde S, Schwardt JD, Thrall MA. (1995) Bone changes in

mucopolysaccharidosis VI in cats and the effects of bone marrow transplantation: mechanical testing of long bones. Bone. 17:485-489.

[120] Hutchinson TM, Steele CR, Snow-Harter R, Whalen RT, Marcus R, Arnaud SB. (1994)

Bending stiffnesss in the tibia of healthy men aged 26-51 years. Med.Sci.Sports. 26:S21. [121] Myburgh KH, Charette S, Zhou L, Steele CR, Arnaud S, Marcus R. (1993) Influence of

recreational activity and muscle strength on ulnar bending stiffness in men. Med.Sci.Sports.Exerc. 25:592-596.

[122] Kiebzak GM, Box JH, Box P. (1999) Decreased ulnar bending stiffness in osteoporotic

Caucasian women. J.Clin.Densitom. 2:143-152. [123] Smith SR, Burshell J, Lindberg J, Bober M, Davies MJ. (1994) Adaptation of bone in a

kindred with osteogenesis imperfect. J.Bone.Miner.Res. 9:S424. [124] McCabe F, Zhou LJ, Steele CR, Marcus R. (1991) Noninvasive assessment of ulnar bending

stiffness in women. J.Bone.Miner.Res. 6:53-59. [125] Conzemius MG, Brown TD, Zhang Y, Robinson RA. (2002) A new animal model of femoral

head osteonecrosis: one that progresses to human-like mechanical failure. J.Orthop.Res. 20:303-309.

[126] Maloney SK, Dawson TJ. (1993) Sexual dimorphism in basal metabolism and body

temperature of a large bird, the emu. Condor. 95:1034-1037.

36

[127] Reed KL, Brown TD. (2001) Elastic modulus and strength of emu cortical bone. Iowa.Orthop.J. 21:53-57.

[128] Davies SJJF. (2002) Ratites and Tinamous. Oxford University Press. New York.

37

Chapter 2 Experimental Approach 2.1. Bone Samples

2.2. Potassium Hydroxide (KOH) Treatment

2.3. Statistical Analysis

2.4. References

38

The following chapter outlines the experimental approach and various techniques used to

determine the consequences of collagen degradation on bone mechanical properties.

2.1. Bone Samples

Female and male emu leg bones were obtained from slaughterhouses and farms in Southern

Ontario, Canada. The animals were approximately 3-5 years of age and therefore skeletally

mature [1,2]. The tibiae were carefully separated from the femora and tarsometatarsi with a

scalpel. The shaft of each tibia was isolated by using a circular saw to remove the ends (15%

of the total bone length from the proximal end and 10% from the distal end), resulting in

bone samples 26 to 32 cm in length. The marrow and trabecular bone from the diaphysis of

the tibiae were removed by drilling longitudinally through the bone shaft, after which the

medullary canal was flushed with tap water. Finally, the skin and overlying tissue was

carefully removed with a scalpel.

The prepared bones were then individually wrapped in saline-soaked gauze and frozen at

-20°C for 6-10 months. Research has shown that freezer storage of well-hydrated bone

specimens does not adversely affect mechanical properties [3-5]. In addition, because certain

bones underwent more techniques than others, the number of freeze-thaw cycles may have

slightly varied among samples. Most samples underwent a maximum of three freeze-thaw

cycles. Furthermore, research has shown that the number of freeze-thaw cycles that a bone

experiences does not significantly affect its mechanical properties (after five freeze-thaw

cycles) [6]. Nevertheless, care was taken to minimize these freeze-thaw cycles and thaw the

samples only when necessary. Samples were allowed to thaw at room temperature for three

hours prior to analysis. Computed Tomography (CT) scans were taken at the mid-point of

each emu tibia with an Aquilion 64 CT scanner (Toshiba, Canada). External anterior-

posterior and medial-lateral diameters (mm), cortical thickness (mm), cross-sectional area

(mm2) and second moment of area in bending orientation (mm4) were then measured from

these binarized images using image analysis software (ImageJ 1.28u, National Institutes of

Health).

39

Bones were divided into four groups. The first group (female and male left tibiae) were used

for bone composition analysis and the second group (female and male left tibiae) were KOH

treated for 0, 1, 3, 7 or 14 days and used for Objective 1 (Initial Study). The third and fourth

groups (female and male right tibiae) were KOH treated at the same time points and then

either fatigued to failure or to 100,000 cycles to induce stiffness loss but not failure

(Objective 2 - Fatigue). Histological assessment was performed on basic fuchsin-stained

microdamage samples in the immediate area of the fatigue to failure fractured mid-span

regions. In all four main groups, two additional 14-day groups of ten female and ten male

bones were filled with saline instead of KOH to act as controls. Herein, the 0-day treated

groups will be referred to as untreated groups and the 14-day filled with saline groups will be

noted as control groups. CT scans were taken at the mid-point of each emu tibia before and

after any treatment (KOH, fatigue to induce stiffness loss) to determine any changes in

geometrical parameters.

Table 2.1 shows the sample sizes (n) for each group. One-centimeter wide sections were cut

from various sections in these bone samples for the techniques used in Objectives 3, 4 and 5.

The fracture surfaces were analyzed using digital images and scanning electron microscopy

(SEM) to study failure mechanisms (Objective 3 - Fractography). Bone samples were

evaluated using quantitative backscattered electron (qBSE) imaging, powder X-ray

diffraction (XRD) and microhardness testing in an attempt to help explain the effect of KOH

treatment on bone mineral (Objective 4 – Collagen Degradation). A selective digestion

technique (α-chymotrypsin), differential scanning calorimetry (DSC), sodium dodecyl sulfate

polyacrylamide gel electrophoresis (SDS-PAGE) and polarized light microscopy (PLM)

were utilized to determine the extent of collagen degradation due to KOH treatment

(Objective 4 – Collagen Degradation). Finally, Raman spectroscopy and atomic force

microscopy (AFM) were used in an attempt to characterize the effects of KOH on the bone

mineral-collagen interface (Objective 5 - Interface). Figure 2.1 represents a schematic of

sample retrieval for analysis of the different techniques (blue boxes represent techniques

focusing on bone mineral and red boxes are highlighting techniques used for bone collagen

evaluation). A flow chart of the experimental techniques performed are summarized in Table

2.2 and shown in Figure 2.2.

40

Table 2.1: Sample sizes of female and male emu tibiae. n

Sex KOH

Treatment Time (days)

Composition Analysis

Mechanical testing

Fatigue to failure

Fatigue to 100,000 cycles

0 (untreated) 20 14 14 10 1 - 10 16 10 3 - 10 17 10 7 - 11 17 10

14 - 10 21 10 14 (saline) - 10 10 10

Female

TOTAL 20 65 95 60 0 (untreated) 20 11 17 10

1 - 10 13 10 3 - 11 20 10 7 - 11 16 10

14 - 11 28 10 14 (saline) - 10 10 10

Male

TOTAL 20 64 104 60

Figure 2.1: Schematic of sample retrieval for analysis for the different techniques. Each box represents a ten mm thick section. Five mm spacing was left between each box.

Fracture

DSC/Powder XRD

Raman/AFM/PLM

SDS-PAGE

OH-Pro/α-CT Assay

μdamage (Fatigue only)

Fractography

μhardness/BSE

41

Table 2.2: Summary of techniques. Hierarchical level Characteristic Technique Macrostructure Bone density

Bone size and shape DXA, QUS, MRTA CT 3-point bending Fatigue testing

Microstructure Cortical microarchitecture BSE Microhardness Microdamage Fractography

Nanostructure Mineral type and crystal alignment Collagen structure and cross-linking Collagen-mineral interface

XRD Gel electrophoresis αCT, DSC, PLM AFM, Raman

Time Points: 1-day, 3-day, 7-day, 14-day Control (0-day, 14-day filled with saline)

Figure 2.2: Flowchart summary of experimental techniques.

Left Tibia

Right Tibia

MASS MRTA QUS DXA CT SCAN 3-point BEND

KOH Treatment

KOH Treatment FATIGUE

MASS MRTA QUS DXA CT SCAN 3-point BEND

MASS MRTA QUS DXA CT SCAN 3-point BEND Microdamage

Microhardness Fractography αCT DSC SDS-PAGE BSE XRD AFM Raman PLM

Left Tibia

Right Tibia

MASS MRTA QUS DXA CT SCAN 3-point BEND

Bone composition analysis

FATIGUE

MASS MRTA QUS DXA CT SCAN 3-point BEND Microdamage

Microhardness Fractography αCT DSC SDS-PAGE BSE XRD AFM Raman PLM

42

43

2.2. Potassium Hydroxide (KOH) Treatment

Mature female and male emu tibiae were endocortically treated with 1 M potassium

hydroxide (KOH) for various time points. The use of KOH, which is known to denature and

digest proteins [7-10], does not require demineralization and consequently, it permits

treatment while retaining the bone mineral [9]. Therefore, this degradation method retains the

mineral phase and the geometric integrity of the bone. While the mechanism of action of

KOH is not well understood, it is likely to be similar to sodium hypochlorite. Sodium

hypochlorite extraction leaves the mineral phase largely unaffected [11]. The high pH of 14

of the 1 M KOH solution is likely to result in conformational changes to the collagen and

may also affect mineral-organic interactions [12].

The experimental setup for the KOH treatment involves first sealing the ends of the bone

segment with polymethylmethacrylate (PMMA). Two segments of clear tubing were placed

in one end of the bone prior to hardening of the PMMA, to allow for filling of the medullary

shaft with a known volume of KOH and to allow any trapped air to escape. Afterwards, the

clear tubing was removed and the ends sealed with PMMA. The bones were placed

horizontally and the periosteal surface wrapped in gauze that was kept moist with 0.9%

saline solution drip, as shown in Figure 2.3. The bones were rotated 180 ° around the axial

axis every 12 hours during KOH treatment, at room temperature. After the desired treatment

time, the KOH solution was reclaimed, its volume measured and the bones were rinsed in

running tap water for one hour.

Figure 2.3: KOH treatment setup.

44

2.3. Statistical Analysis

Statistical analysis was performed using SPSS (SPSS 16.0 for Windows, Chicago, IL) or

SigmaStat (SigmaStat 3.0; San Jose, CA) statistical analysis software packages. All data

(apart from regressions) are presented as mean ± standard error of the mean. A confidence

level of 95% (p=0.05) was considered statistically significant and a confidence level of 90%

(p=0.1) indicated a statistical trend.

Tests for normality and equality of variances were initially performed to determine whether

parametric or non-parametric t-tests should be used. Two-way analysis of variances

(ANOVA, general linear model) was performed to examine the effects of sex and KOH

treatment on all measured parameters. Two-way ANOVA was also used to determine

whether the two factors of measurement technique and KOH treatment time interact on the

respective technique output for both female and male emu tibiae samples. Post hoc pairwise

testing utilized the Fisher’s Least Significant Difference (LSD) test to detect significant

differences between the groups. To compare tibiae fatigue resistance, generalized linear

models were used when comparing regression lines fit to variables that varied by applied

stress or strain (analysis of covariance and multiple linear regressions). This included number

of cycles to failure (N), damage index (DI) rate and creep rate versus applied stress/strain and

specimen group (KOH treatment time, sex). Regressions were tested for homogeneity of

residual variances, differences in slope and differences in height. In addition, multiple

regression analyses were performed to evaluate correlations between certain parameters.

45

2.4. References

[1] Reed KL, Brown TD. (2001) Elastic modulus and strength of emu cortical bone. Iowa.Orthop.J. 21:53-57.

[2] Davies SJJF. (2002) Ratites and Tinamous. Oxford University Press. New York. [3] Turner CH, Burr DB. (1993) Basic biomechanical measurements of bone: a tutorial. Bone.

14:595-608. [4] Pelker R, Friedlaender G, Markham T, Panjabi M, Moen C. (1984) Effects of freezing and

freeze-drying on the biomechanical properties of rat bone. J.Orthop.Res. 1:405-411. [5] Borchers R, Gibson L, Burchardt H, Hayes W. (1995) Effects of selected thermal variables

on the mechanical properties of trabecular bone. Biomaterials.16:545-551. [6] Kang Q, An YH, Friedman RJ. (1997). Effects of multiple freezing-thawing cycles on

ultimate indentation load and stiffness of bovine cancellous bone. Am.J.Vet.Res. 58:1171-1173.

[7] Abe K, Hashizume H, Ushiki T. (1992) An EDTA-KOH method to expose bone cells for

scanning electron microscopy. J.Electron.Microsc.(Tokyo). 41:113-115. [8] Lin TC, Su CY, Chang CS. (1997) Stereomorphologic observation of bone tissue response to

hydroxyapatite using SEM with the EDTA-KOH method. J.Bone.Miner.Res. 36:91-97. [9] Privalova LG, Konstantinova ML, Kulagin VN, Zaikov GY, Sorokin YY, Dreizenshtok GS.

(1980) Kinetic regularities of the degradation of collagen in dilute solutions of sulphuric acid and potassium hydroxide. Pol.Sci. 22:2583-2590.

[10] Arsenault AL. (1990) Vascular canals in bovine cortical bone studied by corrosion casting.

Calcif.Tissue.Int. 47:320-325. [11] Broz JJ, Simske SJ, Corley WD, Greenberg AR. (1997) Effects of deproteinization and

ashing on site-specific properties of cortical bone. J.Mater.Sci.Mater.Med. 8:395-401. [12] Walsh WR, Labrador DP, Kim HD, Guzelsu N. (1994) The effect of in vitro fluoride ion

treatment on the ultrasonic properties of cortical bone. Ann.Biomed.Eng. 22:404-15.

46

Chapter 3 Initial Study

3.1. Introduction

3.2. Experimental Details

3.2.1. Emu bone samples 3.2.2. Bone composition 3.2.3. KOH treatment 3.2.4. CT 3.2.5. DXA 3.2.6. QUS 3.2.7. MRTA 3.2.8. Mechanical testing 3.2.9. Statistical analysis

3.3. Results

3.3.1. Emu bone composition 3.3.2. Sex differences in untreated and control bones 3.3.3. Effect of KOH treatment on bone composition 3.3.4. Effect of KOH treatment on BMD, SOS and EI

3.3.5. Effect of KOH treatment on structural and mechanical properties

3.4. Discussion

3.4.1. Emu bone composition 3.4.2. KOH treatment and collagen degradation 3.4.3. Effect of KOH treatment on bone ductility and toughness 3.4.4. Composite material behaviour 3.4.5. Sex differences 3.4.6. Clinical tools

3.5. Conclusions

3.6. Chapter Summary

3.7. References

47

The following chapter has been reprinted with the kind permission of Elsevier and was

published in Bone under, “A new tool to assess the mechanical properties of bone due to

collagen degradation,” 44 (2009) 840-848. This chapter outlines the initial work that was

done in testing the MRTA as a potential tool for the assessment of bone quality in

endocortically KOH treated emu tibiae to be used as a model for determining the

mechanisms by which bone collagen degradation affects bone mechanical properties.

3.1. Introduction

Bone is a composite material composed of highly substituted, poorly crystalline mineral

(apatite) and a hydrated organic matrix, consisting mainly (~90-95%) of Type I collagen. The

remaining 5-10% of the organic matrix components consists of various noncollagenous

proteins, proteoglycans and small molecules [1]. For the purpose of this study, the organic

matrix will be considered as collagen only. Historically, many investigators have sought to

understand the individual contributions of the mineral and organic phases to the mechanical

properties of bone. Research has commonly been focused on the mineral and it is accepted

that the non-organic component of bone is the primary contributor to its strength and stiffness

[1]. In contrast, the collagen of bone is generally considered to contribute to the toughness

(energy to fracture) of the tissue, mitigating the brittleness of the mineral. However, recent

work suggests that collagen also contributes to bone strength [2,3]. Moreover, denaturing the

collagen or ‘debonding’ the collagen from the mineral phase compromises the composite

structure and results in correspondingly significant decreases in the modulus of elasticity,

ultimate stress and toughness [4,5].

Clinically, bone mineral density (BMD) measurement is a widely used, non-invasive means

of identifying individuals considered to have a high risk of fracture. However, BMD

measures the bone mineral areal density only; this is only one of a number of measurable

contributors to bone strength and fracture risk. Bone strength is derived from both the

amount of bone tissue present (quantity), which is related to bone density and size, as well as

bone quality, which comprises the bone structure and material properties, both of which are

affected by bone remodeling [6]. From a clinical perspective, the BMD may not accurately

predict fracture risk, as it only assays one of the relevant parameters. For example, BMD

48

could only account for approximately 16% of the reduction in fracture risk in an alendronate

treatment study [7], suggesting that other factors were significant contributors to the efficacy

of the treatment. BMD alone is therefore insufficient to predict skeletal fragility. A clinical

tool that can incorporate a broader assessment of bone quality and therefore more accurately

predict fracture risk, would be extremely valuable in identifying at-risk patients.

The most common method of measuring BMD is with dual energy x-ray absorptiometry

(DXA), which uses differential attenuation of x-rays to quantify the mineral component of

bone. Quantitative ultrasound (QUS) is another clinical measure of bone density. As with

DXA, QUS provides an indirect measurement of fracture risk; in this case, using the speed of

sound (SOS) through bone to characterize the tissue [8]. Both DXA and QUS are in

widespread clinical use. However, neither technique has been demonstrated to adequately

distinguish between fracture and non-fracture populations [9]. This is unsurprising since, as

discussed above, bone fragility depends not only on mineral content but also on matrix

properties, architecture and geometry [6] and therefore evaluating the mineral content of

bone alone may be insufficient to predict changes in bone quality [3].

An alternative technology to assess bone quality is the Mechanical Response Tissue Analyzer

(MRTA). The MRTA is a radiation-free, non-invasive instrument developed by NASA to

investigate the effects of space travel on the mechanical properties of astronaut bones [9].

The MRTA measures the cross-sectional bending stiffness (EI) of long bones on the basis of

their response to low-frequency vibration. EI is the product of the elastic modulus, E, and the

areal cross-sectional moment of inertia, I [9]. Research has shown that the measured EI of a

long bone is predictive (R2>0.9) of the maximum strength of the bone [10] and thus in vivo

measurements of EI can be used to assess bone strength. Furthermore, previous studies

suggest that the measurement of EI has the potential to effectively evaluate fracture risk [10-

12]. Finally, studies have demonstrated the ability of the MRTA to measure EI differences

between various populations resulting from exercise [13,14], disease state [15,16] and age

[17].

49

MRTA measurements in humans are limited to the ulna and the tibia because the shafts of

these bones are relatively close to the skin. The tibia of the emu, Dromaius novaehollandiae,

was selected as a model for this study as its size and approximately cylindrical shape is

similar to human long bone and they are therefore suitable for assessment by the MRTA,

which was designed for humans. Furthermore, sexual dimorphism is apparent in emus, as

females are notably heavier and larger compared to males [18], providing a model system to

evaluate differences between female and male bones. Finally, emu tibiae may be a suitable

model for long human leg bones as the bipedal nature of emus is analogous to human

locomotion [19,20]; the emu model has previously been used as a model for femoral head

osteonecrosis [19].

The objectives of this study were two-fold: to investigate the ability of the MRTA to detect

changes in bone mechanical properties induced by changes in the collagen matrix of emu

bone and to evaluate any differences between female and male emu bone. This was achieved

through bone composition analysis, KOH treatment for various durations and finally,

validation of the MRTA and comparison with DXA and QUS in predicting these induced

changes on the mechanical properties of emu tibial bone.

3.2. Experimental Details

3.2.1. Emu bone samples

Leg bones from female and male emus were obtained from slaughterhouses and farms in

Southern Ontario, Canada. The animals were approximately 3-5 years of age and therefore

skeletally mature [20,21]. The tibiae were carefully separated from the femora and

tarsometatarsi with a scalpel. The shaft of each tibia was isolated by using a circular saw to

remove the ends (15% of the total bone length from the proximal end and 10% from the

distal end), resulting in bone samples 26 to 32 cm in length. The marrow and trabecular bone

from the diaphysis of the tibiae were removed by drilling longitudinally through the bone

shaft, after which the medullary canal was flushed with tap water. Finally, the skin and

overlying tissue were carefully removed with a scalpel. The prepared bones were then

individually wrapped in saline-soaked gauze and frozen at -20 °C until use.

50

All bone samples were allowed to thaw at room temperature for three hours prior to analysis.

Bones were divided into two groups: the first group (n=40) was used for bone composition

analysis and the second group (n=120) was used to evaluate the effects of different

degradation times on the measurements made by various bone analysis techniques. Within

these two groups, the bones were further subdivided into female and male groups.

3.2.2. Bone composition

The four major constituents of bone - water, fat, mineral and collagen - were quantified to

provide the overall composition of emu bone. One centimeter wide sections were cut from

the midshaft of each tibial shaft from female (n=20) and male (n=20) emu tibiae, using a

circular saw. One section from each shaft was used to quantify each component.

The water content of emu bone was determined by weighing the samples before and after

drying overnight in an oven at 105 °C. The percentage weight change was determined and

ascribed to the water content. All other measurements were expressed with respect to the ‘dry

weight’ of bone.

Fat was removed from the dry bone samples using 99.5% acetone (Fisher Scientific,

Pittsburgh, PA). Individual bone samples were placed in plastic cassettes and immersed in

acetone, while being agitated on a Vibromax Shaker (IKA Process Equipment, Wilmington,

NC). After 30 minutes, the acetone was exchanged for fresh acetone. This was repeated five

times, resulting in a three-hour extraction. The samples were then oven-dried for 30 minutes

at 105 °C, reweighed and the percentage weight change recorded and attributed to fat

content.

The mineral content in the emu bone was determined by weighing the dry samples before

and after ashing overnight in a furnace (Model F46120CM Barnstead, Thermolyne

Corporation, Dubuque, IA) at 800 °C. The percentage of weight remaining after ashing was

attributed to mineral content.

51

The average initial collagen content in the emu bone was then calculated from the weight

loss induced by ashing. Initial collagen content was also measured independently using the

hydroxyproline assay. This colourimetric assay is an established method of determining

collagen content in cartilage and soft tissue [22-24] as well as in bone [25]. Hydroxyproline

was quantified according to the assay described by Woessner [22], which involves four

stages: digestion of the collagen, hydrolysis of the peptides, colour development and

microplate absorbance measurement. The hydroxyproline assay was performed on dried,

defatted and demineralized bone specimens that were digested in papain (Sigma-Aldrich,

Milwaukee, WI) at 65 °C for 96 hours. After digestion, the bone samples were hydrolyzed in

6 N hydrochloric acid (Fisher Scientific, Pittsburgh, PA), incubated in a block heater at 110

°C for 18 hours and then neutralized using sodium hydroxide (Fisher Scientific, Pittsburgh,

PA). Next, the samples were prepared for colourimetric analysis with the addition of 0.05 N

chloramine-T (Sigma-Aldrich, Milwaukee, WI), 3.15 N perchloric acid (Fisher Scientific,

Pittsburgh, PA) and Ehrlich’s Reagent (Sigma-Aldrich, Milwaukee, WI). The colourimetric

reaction was quantified with a μ-Quant Microplate Spectrophotometer (BioTek Instruments,

Winooski, VT) at 560 nm. It was assumed that 10% of the protein in emu bone collagen is

hydroxyproline [22,26]. Absorbance values were plotted against the concentration of

standard hydroxyproline (0-5 μg) and the quantity of hydroxyproline in the bone samples

was determined from the standard curve.

3.2.3. KOH treatment

The endocortical lumens of whole emu tibiae were filled with 1 M potassium hydroxide

(KOH) (Fisher Scientific, Pittsburgh, PA) to degrade the protein [27]. The use of KOH as a

degradation agent does not require a demineralization step. Consequently, this degradation

method retains the mineral phase and the geometric integrity of the bone. In addition, this

endocortical treatment (from the inside of the bone) models the degradation of bone with age.

The effect of KOH treatment on the mechanical properties of bone was determined in 50

female and 50 male tibiae, with ten samples allocated to each of 0, 1, 3, 7 or 14-day

treatments. Two additional 14-day groups of ten female and ten male emu tibiae were filled

with saline instead of KOH to act as controls.

52

The experimental setup for KOH treatment began with sealing the ends of the bone segment

with polymethylmethacrylate (PMMA, SR Ivolen kit, Ivoclar Vivadent, Mississauga, ON).

Two segments of clear tubing were placed in one end of the bone prior to hardening of the

PMMA seal, to allow for filling of the medullary cavity with a known volume of 1 M KOH

and to allow air to escape. Afterwards, the clear tubing was removed and the ends sealed with

PMMA. The bones were held horizontally over a collection basin with the periosteal surface

wrapped in four layers of gauze that was periodically moistened with saline. The bones were

rotated 180 ° around the axial axis every 12 hours during KOH treatment at room

temperature. After the desired treatment period, the KOH solution was reclaimed and its

volume measured (to ensure that no leakage occurred during KOH treatment). The

endocortical surfaces of the bones were rinsed in running tap water for one hour.

The hydroxyproline content in the KOH solution was determined using the hydroxyproline

assay, as described above, to estimate the amount of collagen removed during KOH

treatment (presumed to migrate from the bone to the KOH solution). However, during the

hydrolysis stage of the hydroxyproline assay, the addition of acid was not required as the

solutions were sufficiently basic to permit alkali hydrolysis of the peptides by incubation for

18 hours at 110 °C. After hydrolysis, the hydrolyzate was neutralized using 37%

hydrochloric acid (Fisher Scientific, Pittsburgh, PA). The amount of hydroxyproline was

determined using a linear standard curve (0-5 μg of hydroxyproline). The calculated

concentrations (μg/mL) were then multiplied by the amount of KOH used (mL) and

normalized to specimen dry mass, assuming 30% collagen is present in dry bone [2].

Collagen content in the KOH treated bone specimens (two one cm thick sections of bone

from both sides of the fracture surface in the 24 cm region of interest) was also determined

with the hydroxyproline assay assuming that collagen is 10% hydroxyproline [22,26] by

mass and normalizing to specimen dry mass.

The emu tibiae properties were assessed using CT, DXA, QUS, MRTA and three-point

bending before and after KOH treatment.

53

3.2.4. CT

Computed Tomography (CT) scans were taken at the mid-point of each emu tibia with an

Aquilion 64 CT scanner (Toshiba, Canada). External anterior-posterior and medial-lateral

diameters (mm), cortical thickness (mm), cross-sectional area (mm2) and second moment of

area in bending orientation (mm4) were then measured from these binarized images using

image analysis software (ImageJ 1.28u, National Institutes of Health).

3.2.5. DXA

The bone mineral density (BMD) was measured by DXA on a Lunar Prodigy Advance

system (General Electric, Madison, WI) using the ‘L-Spine’ mode. The bones were scanned

in a clear acrylic container filled with water to mimic the effects of soft tissue [9]. The region

of interest for this scan was 24 cm and thus, a 24 cm region at the centre of each tibial

diaphysis was demarcated with a marker in order that the scanned section of the bone

specimens was consistent and reproducible.

3.2.6. QUS

Speed of sound (SOS) measurements were obtained with a Sunlight Omnisense QUS device

(Sunlight Medical Ltd., Tel-Aviv, Israel). As the apparatus is designed for in vivo testing and

these bone samples had been excised from the surrounding tissue, slices of extra-firm tofu

(6.5 cm x 4.5 cm x 0.5 cm; Sunrise Soya Foods) were used to mimic soft tissue [9].

Measurements were taken at the centre of the 24 cm region of interest, defined by the DXA

analysis, using a hand-held probe with ultrasound gel to acoustically couple the probe to the

bone. During the test, the probe was moved perpendicular to the bone axis in the manner

recommended by the manufacturer.

3.2.7. MRTA

The MRTA measures the cross-sectional bending stiffness (EI) of long bones by applying a

low frequency (0 to 1600 Hz) vibration to the skin surface, which is transmitted to the bone

using an electromagnetic shaker with an impedance head probe and tip. A transducer

connected to the probe measures the force and acceleration response from the bone. The bone

is modeled as a beam in three-point bending and the force and acceleration values are used to

54

calculate EI [9,10]. MRTA measurements were taken at the centre of the 24 cm region of

interest defined by the DXA analysis. As with QUS testing, the soft tissue was mimicked, in

this case by placing a 3 cm x 3 cm x 0.5 cm foam segment (Foam Craft Sheets) between the

probe and the bone [9]. The contact points used to support the bone in the three-point

bending configuration were aligned with the demarcation lines, 24 cm apart. Five replicates

of the EI measurements were taken for each specimen.

3.2.8. Mechanical testing

Female and male control, untreated and KOH treated emu tibiae were tested in three-point

bending using a servo-hydraulic materials testing machine (Model 8511, Instron, Canton,

MA). Each specimen was placed on the two lower support bars (24 cm apart) with the

anterior side of the bone facing up. The bone was secured in the three-point bending jig with

a 100 N preload and was then loaded to failure at a displacement rate of 0.04 mm/s. Force

and displacement data were collected at a rate of 0.1 data points per second using a

computerized system interfaced with the Instron (Fast-track 2 Software, Instron Corp.,

Canton, MA) until failure. The displacement was then calculated from the crosshead speed

and time data. A load-displacement curve was created for every specimen and was used to

determine the ultimate load (N), failure displacement (mm), post-yield displacement (mm),

energy to failure (mJ), plastic energy (mJ) and stiffness (N/mm).

Load-displacement data was normalized to eliminate differences caused by geometrical

variation and to evaluate intrinsic bone material properties. Computed tomography (CT)

scans were taken at the mid-point of each emu tibia with an Aquilion 64 CT scanner

(Toshiba, Canada). Stress (σ; MPa) and strain (ε; %) were then calculated using the following

equations [28]:

where F is the measured load (N), L is the span of the lower supports (24 cm), APφ is the

external diameter in the anterior-posterior direction (mm), Ixx is the moment of inertia about

the mediolateral axis (mm4) and D is the measured displacement (mm). Stress-strain curves

xx

AP

ILF

σ⋅⋅

= 10062 ∗=

LD APφ

ε

55

were created from the normalized data. From these curves, the elastic modulus (E; GPa),

ultimate stress (σuts; MPa), yield stress (σy; MPa), failure stress (σf; MPa), failure strain (εf,

%), post-yield strain (εpy, %), post-yield toughness (plastic energy) (Upy; mJ/mm3) and

toughness (U; mJ/mm3) were determined.

Prior to KOH treatment, all bones were tested non-destructively in the elastic region (to a

maximum of 1000 N) to determine the initial stiffness and elastic modulus. These values

were used to determine the percent change in modulus for the MRTA and three-point

bending techniques.

3.2.9. Statistical analysis

For comparisons between measurement techniques within the same time groups, Student’s t-

tests were used. Tests for normality and equality of variances were initially performed to

determine whether parametric or non-parametric t-tests should be used. Two-way analysis of

variance (ANOVA) was used to determine whether the two factors of measurement

technique and time interact on the respective technique output. Post hoc pairwise testing

utilized the Fisher’s Least Significant Difference (LSD) test. The effect of KOH treatment

time and sex on mechanical data was also analyzed using a two-way ANOVA. Two-way

ANOVA tests were conducted using SigmaStat statistical analysis software (SigmaStat 3.0,

San Jose, CA), while all other tests were conducted using SPSS (SPSS 16.0 for Windows,

Chicago, IL) statistical analysis software. All data are presented as mean ± standard error of

the mean. A confidence level of 95% (p=0.05) was considered statistically significant.

3.3. Results

All parameters measured for the control groups (14-day filled with saline) were similar to

those of the untreated groups (0-day) for both female and male emu tibiae (Appendix A).

3.3.1. Emu bone composition

There were no significant composition differences detected for water, fat, mineral and

collagen content measures between female and male emu tibiae. The water content was

found to be 14.4% ± 0.3% and 14.4% ± 0.5% for the female and male emu bone samples,

56

respectively. The overall dry weight composition of the emu tibiae samples is summarized in

Table 3.1.

Table 3.1: Collagen, mineral and fat content measurements for female and male emu tibiae. Female Male

Collagen Mineral Fat Collagen Mineral Fat

Average % dry mass 31.3 ± 1.8 68.1 ± 0.1 0.6 ± 0.1 29.8 ± 2.1 68.4 ± 0.3 0.5 ± 0.1

Total % dry mass 100.0 ± 1.8 98.7 ± 2.1

3.3.2. Sex differences in untreated and control bones

Table 3.2 summarizes the significant differences observed between untreated female and

male emu bones. The control bones were not different from untreated samples, allowing

pooling of data. CT analysis of all samples before KOH treatment showed significant

differences in geometrical properties between sexes. Females had significantly greater

moment of inertia (p=0.001), cortical area (p=0.001), thickness (p=0.05) and mediolateral

(ML) diameter (p=0.04) compared to male bones. Female emu bones also had significantly

higher mass (p=0.001), BMD (p=0.001) and BMC (p=0.001) values compared to male emu

bones. In terms of mechanical properties, female control (14-day filled with saline) and

untreated samples (0-day) had significantly higher ultimate stress (p=0.01) and failure stress

(p=0.001) compared to male emu tibiae. However, male bones (both untreated and control)

had significantly higher failure strain (p=0.03) and post-yield strain compared to female

(p=0.02) emu tibiae.

57

Table 3.2: Average geometrical parameters, BMD and BMC values for all samples prior to KOH treatment. Average normalized mechanical properties of untreated (0-day) and control emu bones.

Parameter Female Male

Medial-Lateral diameter (mm) 28.6 ± 0.3a 27.8 ± 0.3a

Anterior-Posterior diameter (mm) 23.4 ± 0.2A 22.8 ± 0.3A

Cortical Thickness (mm) 4.2 ± 0.1b 3.9 ± 0.1b

Cortical area (mm2) 274.9 ± 4.7c 240 ± 5.2c

Moment of Inertia (mm4) 14132 ± 416d 12068 ± 455d

BMD (g/cm2) 1.3 ± 0.02e 1.2 ± 0.02e

BMC (g) 73.7 ± 1.1f 65.4 ± 1.6f

Ultimate Strength (MPa) 141.9 ± 5.1g 124.7 ± 6.8g

Failure Strength (MPa) 111.3 ± 6.0h 70.8 ± 13.7h

Failure strain (%) 1.6 ± 0.2i 2.2 ± 0.3i

Post Yield strain (%) 1.1 ± 0.2j 1.8 ± 0.3j

Toughness (mJ/mm3) 1.2 ± 0.2 1.6 ± 0.2

‘Post Yield’ Toughness (mJ/mm3) 1.0 ± 0.2 1.4 ± 0.2

Elastic Modulus (GPa) 16.9 ± 0.4 17.8 ± 0.6

Lowercase letters (a,b,c) denote significance (p ≤ 0.05) between the marked groups. Uppercase letters (A,B,C) denote a trend (0.1 < p ≤ 0.05) between the marked groups.

3.3.3. Effect of KOH treatment on bone composition

The results for the percent collagen removed versus KOH treatment time are presented in

Figure 3.1. Two-way ANOVA indicates that the percentage of removed collagen varied with

KOH treatment time but not sex; post hoc testing indicated a statistically significant

difference between 1 day and 14 days. Percent collagen loss was less than 0.05% after 14

days of KOH treatment. Total bone weight loss over this time period was negligible (0.5%)

and not a function of KOH treatment time or sex (Figure 3.2).

58

KOH Treatment Time1 day 3 day 7 day 14 day

Perc

ent

colla

gen

loss

(%

)

0.00

0.02

0.04

0.06

0.08

0.10

MaleFemale

Figure 3.1: Percent collagen weight removed versus KOH treatment time for female and male emu tibiae. Significant differences were observed in the collagen weight removed with KOH treatment time but sex was not a factor in collagen removal. Maximum percent collagen removed was negligible (0.05%).

KOH Treatment Time

1 day 3 day 7 day 14 day

Perc

ent

bone

mas

s ch

ange

(%

)

0.0

0.2

0.4

0.6

0.8

1.0 MaleFemale

Figure 3.2: Percent bone weight loss versus KOH treatment time for female and male emu tibiae. No differences were observed in the bone weight loss with KOH treatment time or sex.

59

3.3.4. Effect of KOH treatment on BMD, SOS and EI

The effect of KOH treatment on bone quality was assessed by DXA (BMD), QUS (SOS),

MRTA (EI) and three-point bending before and after KOH treatment at all time points for

both female (Figure 3.3 (a)) and male (Figure 3.3 (b)) emu tibiae. No differences in BMD or

SOS, as measured by DXA and QUS, respectively, resulted from KOH treatment; the

measured values varied by less than 2% between KOH treatment times. However, there were

significant changes in the modulus of elasticity between all time points, except between 3

and 7 days, as determined by three-point bending and by MRTA for both female and male

emu tibiae. No differences were seen between percent changes in modulus of elasticity as

measured by three-point bending and the MRTA for both sexes. There were no differences

between sexes.

Figure 3.3: Percent changes of bone quality measurements reported by the different measurement techniques as a function of KOH treatment time for (a) female and (b) male emu tibiae. No differences were seen between sexes. Significant changes in modulus of elasticity were observed between all time points (except between 3 and 7 days), as determined by three-point bending and MRTA for both female and male emu tibiae. DXA and QUS measurements did not detect any changes caused by KOH treatment.

3.3.5. Effect of KOH treatment on structural and mechanical properties

CT analysis showed no changes in geometrical properties such as moment of inertia, cortical

area, thickness and anterior-posterior (AP) or medial-lateral (ML) diameter with KOH

treatment time or between sexes (Table 3.3). On the other hand, three-point bending tests

demonstrated significant variations between the different time points for both sexes. All

a) Females

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

b) Males

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

0 1 3 7 14-30

-25

-20

-15

-10

-5

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

0 1 3 7 14-30

-25

-20

-15

-10

-5

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

a) Females

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

b) Males

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

0 1 3 7 14-30

-25

-20

-15

-10

-5

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

0 1 3 7 14-30

-25

-20

-15

-10

-5

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

60

normalized parameters followed the same trends as un-normalized parameters, confirming

that there are significant differences in mechanical integrity with KOH treatment at the bone

material level. Representative stress-strain curves for different KOH treatment times for

female (Figure 3.4 (a)) and male (Figure 3.4 (b)) emu tibiae reveal changes with KOH

treatment time and significant differences between sexes.

Table 3.3: Average geometrical parameter changes after KOH treatment for female and male emu tibiae.

KOH treatment time in days Parameter

Sex

1 3 7 14

Female 0.9 ± 0.5 0.3 ± 0.3 0.5 ± 0.3 0.3 ± 0.4 Change in medial-lateral diameter (%) Male 0.4 ± 0.4 0.4 ± 0.4 0.3 ± 0.3 0.2 ± 0.3

Female 0.3 ± 0.2 0.2 ± 0.2 0.8 ± 0.6 0.8 ± 0.6 Change in anterior-posterior diameter (%) Male 0.2 ± 0.3 0.5 ± 0.5 0.4 ± 0.3 0.8 ± 0.7

Female 0.2 ± 0.5 0.4 ± 0.6 0.6 ± 0.3 0.4 ± 0.6 Change in cortical thickness (%) Male 0.7 ± 0.8 0.6 ± 0.7 0.8 ± 0.2 0.2 ± 0.4

Female 0.2 ± 0.2 0.2 ± 0.4 0.2 ± 0.6 0.8 ± 0.6 Change in cortical area (%) Male 0.7 ± 0.6 0.5 ± 0.9 0.9 ± 0.6 0.9 ± 0.3

Female 0.2 ± 0.2 0.2 ± 0.3 0.3 ± 0.7 0.6 ± 0.2 Change in moment of inertia (%) Male 0.6 ± 0.4 0.7 ± 0.4 0.4 ± 0.6 0.6 ± 0.4

61

Figure 3.4: Representative stress-strain curves for 0-14 day KOH treatment of (a) female and (b) male emu tibiae. Significant differences were seen with KOH treatment time for both sexes. Significant differences were seen in the plastic region between female and male emu tibiae with KOH treatment time.

The average ultimate stress and yield stress decreased significantly with KOH treatment for

female and male emu tibiae. Moreover, the elastic modulus significantly decreased at the 14-

day KOH treatment time point compared to the untreated (0-day) samples for female and

male bones. A significant increase in failure strain was seen with KOH treatment time due to

a significant increase in the post-yield strain for both female and male bones. This increased

post-yield strain contributed to a significant increase in toughness over the time period of the

experiment. No relationship between the mechanical properties in the elastic region and

KOH treatment time was observed (Figure 3.5 (a)). Male bones failed at lower stresses

(Figure 3.5 (b)), exhibited larger failure strains (Figure 3.5 (c)) and absorbed more energy

(Figure 3.5 (d)) compared to female bones.

a) Females

Strain (%)

Stre

ss (

MP

a)b) Males

Strain (%)

Stre

ss (

MP

a)

0 1 2 3 40

20

40

60

80

100

120

140

160

180 0 day1 day3 day7 day14 day

0 1 2 3 40

20

40

60

80

100

120

140

160

1800 day1 day3 day7 day14 day

a) Females

Strain (%)

Stre

ss (

MP

a)b) Males

Strain (%)

Stre

ss (

MP

a)

0 1 2 3 40

20

40

60

80

100

120

140

160

180 0 day1 day3 day7 day14 day

0 1 2 3 40

20

40

60

80

100

120

140

160

1800 day1 day3 day7 day14 day

62

Figure 3.5: Mechanical properties as a function of KOH treatment time for female and male emu tibiae: (a) Elastic Modulus: significant decrease when comparing treatment time 0-day and 14-day for both female and male emu tibiae. (b) Failure Stress: significant differences between sexes at all time points except the 1-day and 14-day treatment time point. (c) Failure Strain: significant increase after 14-day KOH treatment time and significant differences between sexes at all time points. (d) Toughness: significant increase after 14-day KOH treatment and significant differences between male and female emu tibiae at all time points except at 0-day and 14-day KOH treatment.

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

0 day 1 day 3 day 7 day 14 day0

20

40

60

80

100

120

140

160

180 MaleFemale

KOH Treatment (days) KOH Treatment (days)

0 day 1 day 3 day 7 day 14 day0

5

10

15

20

25 MaleFemale

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

0 day 1 day 3 day 7 day 14 day0

20

40

60

80

100

120

140

160

180 MaleFemale

KOH Treatment (days) KOH Treatment (days)

0 day 1 day 3 day 7 day 14 day0

5

10

15

20

25 MaleFemale

c) Failure Strain (%) d) Toughness (mJ/mm3)

KOH Treatment (days) KOH Treatment (days)

0 day 1 day 3 day 7 day 14 day0

1

2

3

4

5

6 MaleFemale

0 day 1 day 3 day 7 day 14 day0

1

2

3

4 MaleFemale

c) Failure Strain (%) d) Toughness (mJ/mm3)

KOH Treatment (days) KOH Treatment (days)

0 day 1 day 3 day 7 day 14 day0

1

2

3

4

5

6 MaleFemale

0 day 1 day 3 day 7 day 14 day0

1

2

3

4 MaleFemale

63

3.4. Discussion

In this experiment, female and male emu tibiae were treated with 1 M potassium hydroxide

(KOH) solutions for various lengths of time. DXA, QUS, MRTA and three-point bending

measurements were performed on the tibiae before and after KOH treatment. KOH treatment

did not change the geometry or mineral content of the bone samples and only a minimal

amount of collagen was extracted. However, the mechanical properties of the bone were

significantly compromised, which suggests that the mechanical contribution of the collagen

was impaired. These mechanical changes were observed with MRTA but not with either of

the other two non-invasive assessment techniques.

3.4.1. Emu bone composition

The mean weight fractions obtained for each of the main constituents of emu bone (mineral,

collagen and fat) accounted for 100.0% ± 1.8 (female) and 98.7% ± 2.1 (male) of the

measured dry weight. The 68:31 ratio of mineral to organic matrix obtained for the emu

tibiae is similar to the 65:35 ratio reported for human bone [3,29]. Moreover, previous

studies indicate that avian (chick) bone at two years of age contains approximately 67% of

the non-defatted dry weight as mineral (determined by ash weight measurement) [29].

3.4.2. KOH treatment and collagen degradation

KOH solution has been used previously as a bone degradation agent [30,31], although it is

used less frequently than sodium hypochlorite (bleach) [32,33]. Several authors have used an

EDTA-KOH method to remove the bone matrix to expose bone cells for scanning electron

microscopy [27,31,34-35], decalcifying the tissue with EDTA followed by digestion of

collagen fibers with KOH. Ushiki and Ide determined that the length of KOH treatment was

critical and that a short treatment time resulted in incomplete digestion of the collagen [34].

Immersion in commercial (5.25%) sodium hypochlorite solution results in removal of

>97.5% of the collagen from small blocks of bone (3.6 mm x 3.6 mm x 40 mm) over 21 days

[33]. Note that in our study, only about 0.05% of the collagen was actually extracted, based

on the hydroxyproline assay of the KOH solution. This may be due to either a difference in

the mechanism of action or simply because the bleach-treated samples had a much higher

64

surface-area-to-volume ratio. Sodium hypochlorite extraction leaves the mineral phase

largely unaffected [32-33,36]; similarly, we did not observe changes to the bone mineral as a

result of KOH treatment.

While the KOH treatment did not appear to remove a significant amount of collagen, the

collagen may have been degraded in situ, as an increase in pH may alter secondary and

tertiary structures of the proteins [37]. Degradation of the collagen would explain the

negative effects on the mechanical properties of bone in the absence of substantial changes to

the geometry or the composition of the samples. The endocortical degradation treatment may

have resulted in an inner zone where collagen was significantly degraded, with less

disruption as the distance from the medullary canal increased. This would result in

structurally heterogeneous cross-sections. The increasingly compromised mechanical

properties observed with increased KOH treatment time may therefore be related to an

increasing volume of the degradation zone, rather than a difference in the type or magnitude

of the damage to the collagen. Broz et al. observed an analogous heterogeneous structure

after immersion of bovine bone blocks in sodium hypochlorite, with the outer anorganic

layers displaying severely brittle behaviour when compared to the block interior [33]. The

authors determined that the bleach deproteinization occurred primarily along the bone-fluid

interface and to a much lesser extent within the specimen cross-section.

The exact changes that occur in the collagen as a result of KOH treatment are not well

understood. A minimal amount of collagen was observed to be solubilized. However, the

presence of crosslinks and the intimate apposition of collagen and mineral in bone may

prevent solubilization, even if the polypeptide chains are significantly damaged by alkaline

hydrolysis. In addition, the high pH is likely to result in conformational changes to the

collagen. Possible mechanisms for the altered mechanical properties are discussed below.

3.4.3. Effect of KOH treatment on bone ductility and toughness

Collagen is the component of bone that is associated with ductility and toughness. It has been

shown that a decrease in collagen content result in decreases in toughness [2,3]. In addition,

if the collagen is altered, the toughness may also be affected. Wang et al. [5] denatured the

65

collagen in human cadaver bone using heat and observed substantial decreases in ultimate

strength, elastic modulus, failure strain and toughness. In a transgenic mouse model of

Osteogenesis Imperfecta, a collagen mutation was shown to be associated with reduced post-

yield deformation of bone and consequently reduced toughness in cortical bone [38]. In

addition, a reduction in crosslink density (in rats treated with an inhibitor of collagen cross-

linking) was associated with decreased bone strength and modulus [39]. Finally, the collagen

network tends to become weaker with age, leading to decreased bone toughness [4,39]. This

weaker organic network is influenced by the amount and type of collagen crosslinks, which

vary in maturity and valency [4].

However, in this study, the KOH treated bones behaved in a more ductile manner with

increased KOH treatment time, with increasing strains at fracture and correspondingly

increased toughness. This suggests that the mechanism of action of KOH may differ from

those discussed above. We hypothesize that the observed changes in mechanical properties

(reduced modulus and strength, increased ductility and toughness) may result from partial

debonding between the mineral and organic matrix and/or in situ collagen degradation.

3.4.4. Composite material behaviour

Bone can be considered as a polymeric matrix/mineral-filled composite where the mineral

component affects the stiffness and strength of the bone while the quality of the collagen

phase influences bone toughness [40]. The mechanical properties of any composite material

are influenced by the shape, size, orientation, mechanical properties and volume fraction of

the constituents [41,42]. Interfacial bonding interactions between the constituents of a

composite also play an important role in its mechanical properties [43]. Bundy was the first

to suggest that the quality of bone decreases due to changes in bonding between the bone

mineral and the organic phase [44,45]. Since then, several authors have tested and published

data that support this hypothesis [37,46-59].

Changes to bone mechanical properties that are similar to those observed here (lower elastic

modulus, yield and ultimate stress and higher ultimate strain and toughness) were shown to

result from in vitro immersion of bone specimens in concentrated fluoride or phosphate ion

66

solutions for several days [46-50]. The authors attributed these alterations to compromised

interfacial bonding between the bone mineral and collagen as a result of the ability of the free

F- and PO43- ions to compete with the negative domains of organic constituents for the

binding sites on the bone mineral surface. This competition may break or partially debond

the links between the mineral and matrix.

Kotha and Guzelsu developed a simple shear lag model to analyze the stress transfer between

the mineral and organic components of bone in order to investigate the mineral-collagen

interface [51]. They focused on the bone mineral-organic tissue interactions by investigating

the effect of interphase mechanical properties and bonding on the mechanical properties of

bone. Their model showed that a composite with lower bonding would have decreased elastic

modulus, yield and ultimate stresses while the ultimate strain is increased [51]. They

concluded that bone with less bonding between the mineral and the organic phase would

become more ductile. This is very similar to our observations and suggests a mechanism for

the significant alteration of mechanical properties in the absence of geometric or

compositional changes.

Adsorption of the organic matrix to bone mineral may also depend upon electrostatic bonds

[52]. Gupta et al. recently showed that the plastic deformation in bone is characterized by a

very small activation enthalpy that is lower than the energy required to break typical covalent

bonds but much higher than the energy of hydrogen bonds [53]. Based on these values, the

authors postulated that the plastic (post-yield) deformation in bone might be associated with

the disruption of electrostatic or ‘sacrificial’ bonds within or between molecules in the

extrafibrillar matrix of bone [53]. These ‘sacrificial’ bonds have also been postulated to be

partially responsible for the toughness of bone [54,55]. These sacrificial bonds may be the

reason for the observed increase in toughness.

The pH of the bone may also affect mineral-organic interactions [56,57]. At high pH, the

surface charges of the mineral and organic components become negative, which establishes a

different and electrostatically unfavourable condition for mineral-collagen interactions,

leading to reduction in interfacial bonding [58]. This change in mineral-collagen interactions

67

caused by alkalinity may have taken place within the emu bones given that the 1 M KOH

solution had a pH of 14 for the duration of the treatment.

Therefore, we hypothesize that the interface between the mineral and collagen matrix may be

partially compromised by the KOH solution. It is well-known that the interface in composite

materials plays a major role in governing specific properties such as compressive and shear

behaviour, fracture modes and toughness, as well as stress transfer from externally applied

loads to the reinforcement phase [41,59]. The observed changes in the mechanical properties

of endocortically treated whole emu bone with increased KOH treatment time is consistent

with composite material theory, which predicts material failure primarily at the interface

between the matrix (collagen) and reinforcing phase (mineral) [41].

3.4.5. Sex differences

The mechanical properties of the untreated (0-day) and control specimens (14-day filled with

saline) are similar to results reported in the literature. Reed and Brown [20] measured the

elastic modulus and strength of emu cortical bone (femur, unknown gender) in four-point

bending. They reported that the elastic modulus was 13.1 ± 3.9 GPa, the yield stress was

113.1 ± 29.2 MPa and the ultimate strength was 146.9 ± 32.2 MPa. In this study, the elastic

modulus, yield stress and ultimate strength of the female untreated (0-day) samples were

determined to be 16.9 ± 0.4 GPa, 78.8 ± 4.1 MPa and 141.9 ± 5.1 MPa, respectively

(calculated from three-point bending tests). For the male emu bones, the elastic modulus was

17.8 ± 0.6 GPa, the yield stress was 77.9 ± 4.4 MPa and the ultimate strength was 124.7 ± 6.8

MPa. While both female and male results are comparable with the data obtained by Reed and

Brown [20], significant differences were observed between female and male untreated (0-

day) samples. Female emu tibiae had significantly higher cross-sectional geometry, mass,

BMD and BMC values as well as ultimate stress and failure stress compared to male emu

tibiae. However, male emu tibiae had significantly higher failure strain and post-yield strain

compared to female emu tibiae. As such, male emu tibiae behave in a more ductile manner

than female emu tibiae. This is consistent with the larger size of female emus relative to

males [18]. Furthermore, the brittle behaviour of female emu bones is supported by the

significantly higher BMD and BMC values.

68

There were significant differences observed in the plastic deformation regions between KOH

treated female and male bones. More specifically, male bones had significantly decreased

failure stress and increased failure strain and toughness with increasing KOH treatment times

compared to female bones. In other words, the male emu bones were able to resist

deformation longer and absorb more energy before fracturing. This was an unexpected result,

given that female emu bones are larger and heavier than their male counterparts [18].

However, structural differences may be responsible for the increased post-yield behaviour in

male emu tibiae compared to female emu tibiae where the collagen and/or interface between

the mineral and collagen in female bones may be easier to alter.

Fracture prevalence and incidence are known to differ between the sexes. The resistance of

bone to fracture depends on both its structural characteristics (connectivity and architecture)

and its material characteristics (mineralization and composition) [60]. Seeman has shown that

at any adult age, a lower proportion of men have fractures than women [61]. Panagiotopoulos

et al. recently conducted a cadaveric biomechanical study investigating the impact load

application on strips of bone in relation to age and sex [62]. They found significant

differences in fracture energy absorption (toughness) between the young male and female

groups, whereas the values observed for the older male groups did not significantly differ

from the female groups [62]. Sex affected the fracture energy absorption in the young, with

young males having a higher value. This is in agreement with our results, where the male

emu tibiae had increased failure strain and toughness with increased KOH treatment time

compared to female bones. Panagiotopoulos et al. attributed their results to the fact that bone

strength is not directly dependent on bone density and mass but also on bone microstructure

[62]. Structural differences between the female and male emu tibiae may be responsible for

the significant differences observed in mechanical properties in the plastic deformation

region however, the quality of neither the organic phase nor the bonding between the mineral

and organic phases were addressed in this study and may also be factors.

3.4.6. Clinical tools

Ex vivo mechanical testing of bone samples provides direct measures of the mechanical

properties of bone, which are normally inferred from non-invasive clinical measurements. In

69

this study, KOH treatment of bone altered the properties of the collagen, leading to impaired

mechanical properties as measured by three-point bending tests. This is consistent with other

studies [3-5,63]. These changes were not observed using DXA or QUS. As the composition

of the bone was largely unaltered, the BMD (as measured by DXA) would be unaffected.

The relative insensitivity of DXA to bone quality, rather than just mineral content, suggests

that it may be limited as a clinical tool to predict fracture risk. Similarly, no changes were

observed in the SOS, as measured by QUS. This may be due to a similar insensitivity to bone

quality. With this particular technique however, it may also be an artifact of the specific

treatment protocol used here. It has been shown that QUS measurements depend on the

thickness and density of cortical bone but is more strongly influenced by the density of the

cortex near the periosteal surface than by the endocortex [64]. As the bones were degraded

from the medullary canal towards the periosteum, the acoustic signal may not have

penetrated deeply enough to detect the changes in degraded bone properties [9]. In contrast to

both DXA and QUS, the change in modulus measured by MRTA closely paralleled that

measured by mechanical testing. This is likely due to the ability of the MRTA to provide a

direct measure of the cross-sectional bending stiffness (EI), which is a mechanical property

of long bones [9,10]. This suggests that the MRTA should be further investigated for its use

as a clinical tool. It is a non-invasive test that may more effectively integrate factors that

contribute to bone quality – geometry, mineral content and the amount and characteristics of

the organic component of bone.

3.5. Conclusions

The deleterious effect of KOH treatment on the mechanical properties of bone is apparent

from our findings and those of preceding investigations [27,30-31]. As only a negligible

amount of collagen was removed during this treatment, it is proposed that the observed

impairment of mechanical properties results from in situ collagen degradation and/or

interfacial debonding. Unfortunately, conventional fracture risk screening tools (DXA and

QUS) could not detect this reduction in mechanical integrity. However, this study

demonstrated that the MRTA is capable of detecting the changes in bone mechanical

properties induced by changes in collagen quality and could therefore be a more effective

clinical tool for predicting fracture risk.

70

3.6. Chapter Summary

Through the use of a unique animal model, female and male emu whole tibiae were

endocortically treated with 1 M KOH solution for 1-14 days, resulting in negligible collagen

loss (0.05%), bone mass loss (0.5%), no differences in geometrical parameters and bone

mineral content, but significant changes in mechanical properties. Specifically, KOH treated

samples showed significant decreases in modulus and failure stress and increases in failure

strain and toughness. These changes were more significant in male bones compared to female

bones. The MRTA detected these changes whereas DXA and QUS did not.

The significant changes in bone mechanical properties in this unique emu model resulted in

KOH treated bones behaving in a more ductile manner. This result is a contradiction to

previous studies. Previous studies that have affected collagen have shown a more brittle-like

behaviour [2-5]. As a result, it is hypothesized that the significant changes in bone

mechanical properties may be due to in situ collagen degradation rather than collagen

removal and/or partial debonding between the mineral and organic matrix.

71

3.7. References

[1] Boskey A. (2001) Bone Mineralization. In: Cowin SC, editor. Bone Mechanics Handbook. CRC Press. Boca Raton. Florida. pp. 1-33.

[2] Kaplan FS, Hayes WC, Keaveny TM, Boskey A, Einhorn TA, Iannotti JP. (1994) Form and

Function of Bone. In: Simon SR, editor. Orthopaedic Basic Science. American Academy of Orthopaedic Surgeons. Rosemont. Illinois. pp. 129-138.

[3] Burr DB. (2002) The contribution of the organic matrix to bone's material properties. Bone.

31:8-11. [4] Wang X, Shen X, Li X, Agrawal CM. (2002) Age-related changes in the collagen network

and toughness of bone. Bone. 31:1-7. [5] Wang X, Li X, Bank RA, Agrawal CM. (2002) Effects of collagen unwinding and cleavage

on the mechanical integrity of the collagen network in bone. Calcif.Tissue.Int. 71:186-192. [6] Cefalu CA. (2004) Is bone mineral density predictive of fracture risk reduction?

Curr.Med.Res.Opin. 20:341-349. [7] Cummings SR, Karpf DB, Harris F, Genant HK, Ensrud K, LaCroix AZ, Black DM. (2002)

Improvement in spine bone density and reduction in risk of vertebral fractures during treatment with antiresorptive drugs. Am.J.Med. 112:281-289.

[8] Genant HK, Engelke K, Fuerst T Gluer CC, Grampp S, Harris ST, Jerqas M, Lang T, Lu Y,

Majumdar S, Mathur A, Takada M. (1996) Noninvasive assessment of bone mineral and structure: state of the art. J.Bone.Miner.Res. 11:707-730.

[9] Djokoto C, Tomlinson G, Waldman S, Grynpas M, Cheung AM. (2004) Relationship among

MRTA, DXA, and QUS. J.Clin.Densitom. 7:448-456. [10] Roberts SG, Hutchinson TM, Arnaud SB, Kiratli BJ, Martin RB, Steele CR. (1996)

Noninvasive determination of bone mechanical properties using vibration response: a refined model and validation in vivo. J.Biomech. 29:91-98.

[11] Norrdin RW, Simske SJ, Gaarde S, Schwardt JD, Thrall MA. (1995) Bone changes in

mucopolysaccharidosis VI in cats and the effects of bone marrow transplantation: mechanical testing of long bones. Bone. 17:485-489.

[12] Hutchinson TM, Bakulin AV, Rakhmanov AS, Martin RB, Steele CR, Arnaud SB. (2001)

Effects of chair restraint on the strength of the tibia in rhesus monkeys. J.Med.Primatol. 30:313-321.

[13] Myburgh KH, Charette S, Zhou L, Steele CR, Arnaud S, Marcus R. (1993) Influence of

recreational activity and muscle strength on ulnar bending stiffness in men. Med.Sci.Sports.Exerc. 25:592-596.

[14] Hutchinson TM, Steele CR, Snow-Harter R, Whalen RT, Marcus R, Arnaud SB. (1994)

Bending stiffnesss in the tibia of healthy men aged 26-51 years. Med.Sci.Sports.Exerc. 26:S21.

72

[15] Smith SR, Burshell J, Lindberg J, Bober M, Davies MJ. (1994) Adaptation of bone in a kindred with osteogenesis imperfecta. J.Bone.Miner.Res. 9:S424.

[16] Kiebzak GM, Box JH, Box P. (1999) Decreased ulnar bending stiffness in osteoporotic

Caucasian women. J.Clin.Densitom. 2:143-152. [17] McCabe F, Zhou LJ, Steele CR, Marcus R. (1991) Noninvasive assessment of ulnar bending

stiffness in women. J.Bone.Miner.Res. 6:53-59. [18] Maloney SK, Dawson TJ. (1993) Sexual dimorphism in basal metabolism and body

temperature of a large bird, the emu. Condor. 95:1034-1037. [19] Conzemius MG, Brown TD, Zhang Y, Robinson RA. (2002) A new animal model of femoral

head osteonecrosis: one that progresses to human-like mechanical failure. J.Orthop.Res. 20:303-309.

[20] Reed KL, Brown TD. (2001) Elastic modulus and strength of emu cortical bone.

Iowa.Orthop.J. 21:53-57. [21] Davies SJJF. (2002) Ratites and Tinamous. Oxford University Press. New York. [22] Woessner JF Jr. (1976) Determination of hydroxyproline in connective tissues. In: Hall D,

editor. The methodology of connective tissue research. Joynson-Bruvvers Ltd. Oxford. pp. 227-233.

[23] Walsh BJ, Thornton SC, Penny R, Breit SN. (1992) Microplate reader-based quantitation of

collagens. Anal.Biochem. 203:187-190. [24] Vachon AM, Keeley FW, McIlwraith CW, Chapman P. (1990) Biochemical analysis of

normal articular cartilage in horses. Am.J.Vet.Res. 51:1905-1911. [25] Sinha RK, Shah SA, Hume EL, Tuan RS. (2002) The effect of a 5-day space flight on the

immature rat spine. Spine.J. 2:239-243. [26] Nyman JS, Reyes M, Wang X. (2005) Effect of ultrastructural changes on the toughness of

bone. Micron. 36:566-582. [27] Abe K, Hashizume H, Ushiki T. (1992) An EDTA-KOH method to expose bone cells for

scanning electron microscopy. J.Electron.Microsc.(Tokyo). 41:113-115. [28] Turner CH, Burr DB. (1993) Basic biomechanical measurements of bone: a tutorial.

Bone.14:595-608. [29] Jee WSS. (2001) Integrated Bone Tissue Physiology. In: Cowin SC, editor. Bone Mechanics

Handbook. CRC Press. Boca Raton. Florida. pp. 1-68. [30] Arsenault AL. (1990) Vascular canals in bovine cortical bone studied by corrosion casting.

Calcif.Tissue Int. 47:320-325. [31] Lin TC, Su CY, Chang CS. (1997) Stereomorphologic observation of bone tissue response to

hydroxyapatite using SEM with the EDTA-KOH method. J.Bone.Miner.Res. 36:91-97.

73

[32] Carter DH, Scully AJ, Heaton DA, Young MP, Aaron JE. (2002) Effect of deproteination on bone mineral morphology: implications for biomaterials and aging. Bone. 31:389-395.

[33] Broz JJ, Simske SJ, Corley WD, Greenberg AR. (1997) Effects of deproteinization and

ashing on site-specific properties of cortical bone. J.Mater.Sci.Mater.Med. 8:395-401. [34] Ushiki T, Chizuka I. (1987) Scanning electron microscopic studies of the myelinated nerve

fibres of the mouse sciatic nerve with special reference to the Schwann cell cytoplasmic network external to the myelin sheath. J.Neurocytol. 16:737-747.

[35] Miller BG, Woods RI, Bohlen HG, Evan AP. (1982) A new morphological procedure for

viewing microvessels: a scanning electron microscopic study of the vasculature of small intestine. Anat.Rec. 203:493-503.

[36] Johnson GS, Mucalo MR, Lorier MA. (2000) The processing and characterization of animal-

derived bone to yield materials with biomedical applications: Part 1: modifiable porous implants from bovine condyle cancellous bone and characterization of bone materials as a function of processing. J.Mater.Sci.Mater.Med. 11:427-441.

[37] Moreno EC, Kresak M, Hay DI. (1984) Adsorption of molecules of biological interest onto

hydroxyapatite. Calcif.Tissue.Int. 36:48-59. [38] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

[39] Oxlund H, Barckman M, Ortoft G, Andreassen TT. (1995) Reduced concentrations of

collagen cross-links are associated with reduced strength of bone. Bone. 17:365S-371S. [40] Turner CH. (2006) Bone strength: current concepts. Ann.N.Y.Acad.Sci. 1068:429-446. [41] Jones RM. (1999) Mechanics of composite materials. Taylor & Francis. Philadelphia. [42] Piekarski K. (1973) Analysis of bone as a composite material. Int.J.Eng.Sci. 11:557-565. [43] Monette L, Anderson MP, Grest GS. (1993) Effect of interphase modulus and cohesive

energy on the critical aspect ratio in short-fibre composites. J.Mater.Sci. 28:79-99. [44] Bundy KJ. (1985) Determination of mineral-organic bonding effectiveness in bone--

theoretical considerations. Ann.Biomed.Eng. 13:119-135. [45] Bundy KJ. (1989) Composite Material Models for Bone. In: Cowin SC, editor. Bone

Mechanics. CRC Press. Boca Raton. Florida. pp. 197-208. [46] Walsh WR, Guzelsu N. (1993) The role of ions and mineral-organic interfacial bonding on

the compressive properties of cortical bone. Biomed.Mater.Eng. 3:75-84. [47] Walsh WR, Guzelsu N. (1994) Compressive properties of cortical bone: mineral-organic

interfacial bonding. Biomater. 15:137-145.

74

[48] Walsh WR, Ohno M, Guzelsu N. (1994) Bone composite behaviour: effects of mineral-organic bonding. J.Mater.Sci.Mater.Med. 5:72-79.

[49] Kotha SP, Walsh WR, Pan Y, Guzelsu N. (1998) Varying the mechanical properties of bone

tissue by changing the amount of its structurally effective bone mineral content. Biomed.Mater.Eng. 8:321-334.

[50] DePaula CA, Abjornson C, Pan Y, Kotha SP, Koike K, Guzelsu N. (2002) Changing the

structurally effective mineral content of bone with in vitro fluoride treatment. J.Biomech. 35:355-361.

[51] Kotha SP, Guzelsu N. (2000) The effects of interphase and bonding on the elastic modulus of

bone: changes with age-related osteoporosis. Med.Eng.Phys. 22:575-585. [52] Pearce EI. (1981) Ion displacement following the adsorption of anionic macromolecules on

hydroxyapatite. Calcif.Tissue.Int. 33:395-402. [53] Gupta HS, Fratzl P, Kerschnitzki M, Benecke G, Wagermaier W, Kirchner HO. (2007)

Evidence for an elementary process in bone plasticity with an activation enthalpy of 1 eV. J.R.Soc.Interface. 4:277-282.

[54] Thompson JB, Kindt JH, Drake B, Hansma HG, Morse DE, Hansma PK. (2001) Bone

indentation recovery time correlates with bond reforming time. Nature. 414:773-776. [55] Fantner GE, Hassenkam T, Kindt JH, Weaver JC, Birkedal H, Pechenik L, Cutroni JA,

Cidade GA, Stucky GD, Morse DE, Hansma PK. (2005) Sacrificial bonds and hidden length dissipate energy as mineralized fibrils separate during bone fracture. Nat.Mater. 4:612-616.

[56] Boskey AL. (1992) Mineral-matrix interactions in bone and cartilage. Clin.Orthop.Relat.Res.

244-274. [57] Tye CE, Hunter GK, Goldberg HA. (2005) Identification of the type I collagen-binding

domain of bone sialoprotein and characterization of the mechanism of interaction. J.Biol.Chem. 280:13487-13492.

[58] Walsh WR, Labrador DP, Kim HD, Guzelsu N. (1994) The effect of in vitro fluoride ion

treatment on the ultrasonic properties of cortical bone. Ann.Biomed.Eng. 22:404-415. [59] Weiner S, Wagner D. (1998) The material bone: structure-mechanical function relations.

Annu.Rev.Mater.Sci. 28:271-298. [60] Seeman E, Delmas PD. (2006) Bone quality--the material and structural basis of bone

strength and fragility. N.Engl.J.Med. 354:2250-2261. [61] Seeman E. (2001) Unresolved issues in osteoporosis in men. Rev.Endocr.Metab.Disord. 2:45-

64. [62] Panagiotopoulos E, Kostopoulos V, Tsantzalis S, Fortis AP, Doulalas A. (2005) Impact

energy absorption by specimens from the upper end of the human femur. Injury. 36:613-617.

75

[63] Zioupos P, Currey JD, Hamer AJ. (1999) The role of collagen in the declining mechanical properties of aging human cortical bone. J.Biomed.Mater.Res. 45:108-116.

[64] Prevrhal S, Fuerst T, Fan B, Njeh C, Hans D, Uffmann M, Srivastav S, Genant HK. (2001)

Quantitative ultrasound of the tibia depends on both cortical density and thickness. Osteoporos.Int. 12:28-34.

76

Chapter 4 Fatigue

4.1. Introduction

4.2. Experimental Details

4.2.1. Emu bone samples and KOH treatment 4.2.2. Fatigue testing to failure 4.2.3. Partial fatigue testing 4.2.4. CT 4.2.5. DXA 4.2.6. QUS 4.2.7. MRTA 4.2.8. Three-point bend testing 4.2.9. Microdamage 4.2.10. Statistical analysis

4.3. Results

4.3.1. Effect of KOH treatment on fatigue properties 4.3.2. Microdamage 4.3.3. Effect of partial fatigue testing on BMD, SOS and EI

4.4. Discussion

4.4.1. Fatigue behaviour 4.4.2. Microdamage 4.4.3. Sex differences 4.4.4. Clinical tools

4.5. Conclusions

4.6. Chapter Summary

4.7. References

77

The following chapter has been reprinted with the kind permission of Wiley InterScience and

is currently in press in the Journal of Orthopaedic Research under, “Changes in bone fatigue

resistance due to collagen degradation”. This chapter is a continuation of the work done in

Chapter 3. In this chapter, we further tested the MRTA as a potential tool for the assessment

of bone quality in endocortically treated emu tibiae to be used as a model for determining the

mechanisms by which KOH treatment affects bone fatigue properties.

4.1. Introduction

Bone adapts its mass, architecture and mechanical properties in response to mechanical

loading. While bone is protective against impact, bone is also susceptible to fatigue, a

process by which repetitive loading damages the bone matrix. This repetitive, smaller

loading of bone leads to microcrack formation and accumulation [1]. The fatigue behaviour

of bone resembles that of composite materials, exhibiting a gradual loss of stiffness and

strength throughout cyclic loading due to fatigue damage accumulation [1]. It is well known

that loading of bone introduces small amounts of microdamage which triggers an adaptive

balanced remodeling cycle in healthy bone [2-4]. It has been suggested that bone

microdamage triggers a positive feedback mechanism: microdamage in bone acts as a

stimulus for bone remodeling and remodeling repairs microdamage [5]. However, during

fatigue damage, the remodeling cycle is stimulated, but cannot repair all of the damage

caused by fatigue. Bone accumulates damage over time and eventually fails below its

theoretical strength. When there is time-evolving damage, bone may also experience creep

rupture [1]. Creep is the gradual increase in material strain developed over time when a

material is exposed to a constant stress. Cyclic creep involves the application of a static

force, resulting in permanent (plastic) damage. Creep is a known characteristic of materials

such as polymers [6]. In bone, collagen is abundant and displays polymeric properties [7]. It

has been suggested that the collagen component of bone is responsible for its observed creep

behaviour [8,9]. Since collagen plays a major role in bone biomechanics [10-12], it is

important to identify the role(s) that collagen may have in fatigue fractures.

There is a great deal of interest in developing technologies capable of measuring bone

mechanical properties, such as stiffness, that predict the risk of sustaining a fragility fracture.

78

Current technologies such as dual energy x-ray absorptiometry (DXA) and quantitative

ultrasound (QUS) only measure bone mineral density, a surrogate measure of fracture risk.

Bone fragility depends not only on its mineral content, but also on its matrix properties,

architecture and geometry [13]. Therefore, evaluating the mineral content of bone alone is

insufficient to predict changes in bone quality [14]. An alternative technology for predicting

changes in bone quality is the mechanical response tissue analyzer (MRTA).

The MRTA is a radiation-free, non-invasive instrument that directly measures a bone

mechanical property, the cross-sectional bending stiffness (EI), of long bones. EI is the

product of the elastic modulus, E, and the areal cross sectional moment of inertia, I. Research

has shown that EI of a long bone is predictive of the maximum strength of the bone [15]

thus, measurements of EI can be used to assess bone quality. Furthermore, previous studies

suggest that the determination of EI has the potential to effectively evaluate fracture risk [15-

18].

To study the contribution of collagen to the mechanical properties of bone and the ability of

the MRTA to detect changes in the collagen matrix, we developed a model using the emu

tibiae. In this model, bone collagen was chemically modified with 1 M potassium hydroxide

(KOH) while maintaining the mineral content of bone unaltered [19]. While the KOH

treatment is not physiological, it helps to understand the mechanisms by which collagen

degradation affects bone mechanical properties. The emu tibia was chosen due to its size and

approximate cylindrical shape, making it ideal for devices designed to accommodate human

long bones.

To further determine the contribution of collagen on the mechanical properties of bone, we

investigated the effects of KOH treatment on the fatigue resistance of female and male emu

bone, as controlled cyclic loading of bone is known to significantly reduce bone stiffness and

cause damage [4]. The objectives of this study were two-fold: to investigate changes in

fatigue properties of bone due to collagen modification by endocortical KOH treatment and

to investigate the ability of the MRTA, DXA and QUS to detect changes in bone fatigue

resistance of untreated and KOH treated bone.

79

4.2. Experimental Details

4.2.1. Emu bone samples and KOH treatment

Emu bone tibiae samples were prepared and treated as described previously [19]. Briefly,

skeletally mature (3-5 years of age) whole female and male emu tibiae were carefully

separated from the femora and tarsometatarsi with a scalpel. A circular saw was then used to

remove the ends (15% of the total bone length from the proximal end and 10% from the

distal end) of each tibiae. The marrow and trabecular bone from the diaphysis of the tibiae

were removed by drilling longitudinally through the bone shaft, after which the medullary

canal was flushed with tap water [19]. Finally, the skin and overlying tissue were carefully

removed with a scalpel. The bones were individually wrapped in saline-soaked gauze and

frozen at -20 °C until use. All bone samples were allowed to thaw at room temperature for

three hours prior to analysis.

Bones were divided into two groups: the first group (female and male right tibiae) was KOH

treated for 0, 1, 3, 7 or 14 days and then fatigued to failure. The second group (female and

male right tibiae) was KOH treated at the same time points and then partially fatigued. Partial

fatigue testing refers to specimens fatigued to induce stiffness loss without fracture

(described below). Partial fatigue was undertaken to evaluate the effects of stiffness loss on

the measurements made by the MRTA, DXA and QUS bone analysis techniques. In each

group, two additional 14-day groups of female and male bones were filled with saline instead

of KOH to serve as controls. Table 4.1 shows the sample size (n) for each group.

80

Table 4.1: Sample size of female and male emu tibiae for fatigue testing. n

Sex KOH

treatment time (days)

Fatigue to failure

Fatigue tests that did not fail prior to

106 cycles

Partial fatigue testing (2000 με to 100,000 cycles)

0 14 2 10 1 16 1 10 3 17 - 10 7 17 - 10

14 21 - 10

Female

14 (saline) 10 - 10 0 17 1 10 1 13 2 10 3 20 2 10 7 16 1 10

14 28 1 10

Male

14 (saline) 10 - 10

The emu tibiae were filled with 1 M KOH where the ends were sealed with

polymethylmethacrylate (PMMA, SR Ivolen Liquid and Powder, Ivoclar Vivadent,

Mississauga, ON). The bones were held horizontally over a collection basin and kept moist

with a 0.9% saline solution drip. The bones were rotated 180 ° around the axial axis every 12

hours during the KOH treatment time. After the desired treatment period, the KOH solution

was reclaimed, its volume measured and the endocortical surfaces of the bones rinsed in

running tap water for one hour. The hydroxyproline content in the KOH solution was

determined using the hydroxyproline assay to estimate the amount of collagen removed

during KOH treatment (presumed to migrate from the bone to the KOH solution) [19].

4.2.2. Fatigue testing to failure

Fatigue to failure tests were performed on the first group of female and male emu tibiae. All

fatigue tests were performed in three-point bending using a 40 kN load cell on an Instron

8511 servohydraulic mechanical testing machine (Instron, Norwood, MA) with the anterior

side of the bone facing up and kept moist during the entire test with 0.9% saline solution

drip. Strain was measured using a fatigue-rated, encapsulated single element strain gauge

(WK-06-125BT-350, Vishay Micromeasurements Inc., Malvern, PA) located on the

posterior side directly opposite the loading site oriented along the long axis of the bone [20].

This location is the site of maximum tensile strain in three-point bending. The bones were

81

secured in the jig with a 100 N pre-load and fatigued at a frequency of 2 Hz (physiological

frequency) [21] under load control, corresponding to a set of initial strains ranging from 1500

microstrain to 6000 microstrain peak-to-peak [22]. Time, load, displacement and strain were

acquired using an HBM Spider 8 data acquisition system (HBM, Darmstadt, Germany) at

100 Hz. Data files were exported to ASCII files for subsequent analysis using custom written

LabVIEW code (version 5.0.1, National Instruments, Austin, TX) and Excel 2000

(Microsoft, WA) spreadsheets [20].

The custom-written LabVIEW code sampled the fatigue data files in the time domain every

ith cycle based on the length of the test (i=1 for <5,000 cycles, i=18 for 5,000-10,000 cycles,

i=36 for 10,000-100,000 cycles, i=72 for >100,000 cycles) and provided the following data:

the cycle number (i), time (t), minimum strain (εmin), maximum strain (εmax), minimum stress

(σmin), maximum stress (σmax), secant modulus (E), damage index (DI) and creep strain

(εcreep). The number of cycles to failure (N) was recorded at the end of each test [20].

Data were normalized to eliminate differences caused by geometrical variation. Computed

Tomography (CT) scans were taken at the mid-point of each emu tibia with an Aquilion 64

CT scanner (Toshiba, Canada). Cortical thickness (mm), cross-sectional area (mm2), distance

from the neutral axis to the furthest point in tension (mm) and second moment of area in

bending orientation (mm4) were then measured from these binarized images using image

analysis software (ImageJ 1.28u, National Institutes of Health). Stress (σ; MPa) was

calculated using the following equation [23]:

where F is the measured load (N), L is the span of the lower supports (24 cm), c is the

distance from the neutral axis to the furthest point in tension adjacent to the strain gauge

(mm) and Ixx is the second moment of area about the neutral axis in the bending orientation

(mm4).

xxIcLFσ

4⋅⋅

=

82

From the hysteresis curves, the secant modulus was determined from the slope of the

regression line that passes through the minimum and maximum points of the curve [24].

After each cycle, at the lowest stress in the cycle, the residual strain in the sample was

recorded [24]. The accumulation of residual strains from each cycle until failure was defined

as the accumulated strain, which is often defined as cyclic creep strain (see Figure 4.1).

These strains were plotted as a function of the number of cycles to failure and fitted to a

characteristic creep curve. The creep rate was determined as the slope of the steady-state

region of the creep curve [24] (Figure 4.2). The damage index (DI) for each cycle was

defined as the fractional loss in modulus from the initial value of the secant modulus (Eo) to

the current secant modulus (Ei) in that cycle and related to the following equation: DI = 1-

Ei/Eo. These damage indices were plotted as a function of the number of cycles and the

material damage index rate was determined as the slope of the steady-state region [25,26].

From analyzed cycle-by-cycle data, initial secant modulus (Eo), maximum strain at fracture

(εmax_fracture) and creep strain at fracture (εcreep_fracture) were recorded.

Figure 4.1: Schematic diagram of hysteresis curves produced from a typical fatigue test and resulting defined data (as described above).

83

Figure 4.2: Schematic diagram of characteristic creep curve observed during fatigue testing showing the three characteristic stages: an initial fast strain rise (Stage I), a steady-state region (Stage II) and a final fast increase to failure (Stage III). The creep rate data reported in this study is the steady-state creep rate in Stage II.

4.2.3. Partial fatigue testing

Based on the data collected from fatigue to failure testing, the second group of emu tibiae

(control, untreated and KOH treated) was fatigued at a lower strain (2000 με) to a set number

of cycles (100,000) to induce a reduction in stiffness but not fracture. This strain level was

selected to avoid permanent crushing at the loading and support points on the bones.

Therefore, after the desired treatment period (control (14-day filled with saline), untreated (0-

day), 1, 3, 7 or 14 day KOH treatment), this second group of emu tibiae were then assessed

using CT, DXA, QUS, MRTA and three-point bending. Next, the emu tibiae were subjected

to partial fatigue testing and subsequently re-assessed using CT, DXA, QUS, MRTA and

three-point bending.

4.2.4. CT

Computed Tomography (CT) scans were taken at the mid-point of each emu tibia with an

Aquilion 64 CT scanner (Toshiba, Canada). External anterior-posterior and medial-lateral

diameters (mm), cortical thickness (mm), cross-sectional area (mm2), distance from the

neutral axis to the furthest point in tension (mm) and second moment of area in bending

Slope = creep rateC

reep

Str

ain

Time or Number of Cycles

εcreep_fracture

Stage II Stage III

Stage I

Slope = creep rateC

reep

Str

ain

Time or Number of Cycles

Slope = creep rateC

reep

Str

ain

Time or Number of Cycles

εcreep_fracture

Stage II Stage III

Stage I

84

orientation (mm4) were then measured from these binarized images using image analysis

software (ImageJ 1.28u, National Institutes of Health).

4.2.5. DXA

The bone mineral density (BMD) was measured by DXA on a Lunar Prodigy Advance

system (General Electric, Madison, WI) using the ‘L-Spine’ mode. The region of interest for

this scan was 24 cm and thus, a 24 cm region at the centre of each tibial diaphysis was

demarcated with a permanent marker in order to ensure that the scanned section of the bone

specimens was consistent and reproducible.

4.2.6. QUS

Speed of sound (SOS) measurements were obtained with a Sunlight Omnisense QUS device

(Sunlight Medical Ltd., Tel-Aviv, Israel). Measurements were taken at the centre of the 24

cm region of interest, defined by the DXA analysis, using a hand-held probe with ultrasound

gel to acoustically couple the probe to the bone.

4.2.7. MRTA

The MRTA measures the cross-sectional bending stiffness (EI) of long bones by applying a

low frequency (0 to 1600 Hz) vibration to the skin surface, which is transmitted to the bone

using an electromagnetic shaker with an impedance head probe and tip. A transducer

connected to the probe measures the force and acceleration response from the bone. The

bone is modeled as a beam in three-point bending and the force and acceleration values are

used to calculate EI [15,19]. MRTA measurements were taken at the centre of the 24 cm

region of interest defined by the DXA analysis. Five replicates of the EI measurements were

taken for each specimen.

4.2.8. Three-point bend testing

All bones assigned to partial fatigue testing were tested non-destructively in the elastic

region to determine the elastic modulus. The bones were secured in the three-point bending

jig with a 100 N pre-load and then loaded at a displacement rate of 0.04 mm/s to a maximum

of 1000 N, avoiding permanent damage [19]. After static testing, the specimens were tested

85

in cyclic three-point bending at a frequency of 2 Hz under load control at a strain of 2000 με

to 100,000 cycles to induce stiffness loss. After partial fatigue testing, static pre-yield three-

point bending tests were conducted again, as described above. These values were used to

determine the percent change in modulus as measured by the MRTA and three-point bending

techniques.

4.2.9. Microdamage

For each bone fatigued to failure at 2000 microstrain and 4000 microstrain in each group, ten

mm thick cross-sections of bone were cut and subsequently divided into four sections:

anterior, posterior, medial and lateral. Sections were taken as close as possible to the fracture

surface without including any part of the fractured material. Samples were stained following

a modified version of Burr and Stafford’s [27] amendment of the basic fuchsin method of

Frost [28]. This technique stains pre-existing microcracks in bone prior to histological

embedding and sectioning, allowing them to be differentiated from any damage introduced

during subsequent tissue processing. Initial experiments attempting to infiltrate samples in

1% basic fuchsin (J.T. Baker, Phillipsburg, NJ, USA, Cat. #B660-03) in 70% to 100%

ethanol proved unsuccessful. As such, bone samples were dehydrated in ascending

concentrations of acetone that was incorporated with 1% basic fuchsin and subsequently

infiltrated in ascending ratios of unpolymerized Spurr resin and acetone. The bones were

then embedded in blocks of Spurr resin that was polymerized in a 60 °C oven for 48 hours.

From these blocks, two 100 μm thick cross-sections were cut using a low-speed diamond-

wire saw (DDK Diamond Wire Histo-Saw Model 3241, Wilmington, DE). Sections were

mounted on glass slides for microscopic analyses using a Zeiss microscope attached to a

video camera (Retiga 1300) on a Bioquant image analysis system (Bioquant Nova Prime,

version 6.50.10). Magnification was set to 125X. Microcracks were defined as linear

structures with basic fuchsin staining around the cracks [29]. The following

histomorphometric variables were quantified [26]: microcrack mean length (Cr.Le, μm),

bone cortical area (B.Ar, mm2), microcrack density (Cr.Dn, #/mm2) and surface microcrack

density (Cr.S.Dn, μm/mm2).

86

4.2.10. Statistical analysis

Statistical analysis was performed using SPSS (SPSS 16.0 for Windows, Chicago, IL)

statistical analysis software. All data (apart from regressions) are presented as mean ±

standard error of the mean. A confidence level of 95% (p=0.05) was considered statistically

significant for differences and regressions.

Tests for normality and equality of variances were initially performed to determine whether

parametric or non-parametric t-tests should be used. Two-way analysis of variance

(ANOVA, general linear model) was performed to determine whether the two factors of

measurement technique and KOH treatment time interact. The effect of KOH treatment and

sex on geometrical parameters, bone mineral density and microdamage parameters was also

analyzed using two-way ANOVA. Multiple comparisons were made using appropriate post-

hoc tests to detect significant differences between groups. Post hoc pairwise testing utilized

the Fisher’s Least Significant Difference (LSD) test. Generalized linear models were used

when comparing regression lines fit to variables that varied by applied stress or strain

(analysis of covariance and multiple linear regressions). This includes number of cycles to

failure (N), damage index (DI) rate and creep rate versus applied stress/strain and specimen

group (KOH treatment time, sex). Regressions were tested for homogeneity of residual

variances, differences in slope and differences in height.

4.3. Results

Significant differences were found in the fatigue parameters measured in this study between

the untreated groups and the KOH treated groups, but not between sexes. All parameters

measured for the control group (14-day filled with saline) were similar to those of the

untreated (0-day) time point (Appendix A) for both female and male emu bones.

4.3.1. Effect of KOH treatment on fatigue properties

A certain number of bone samples did not fail before the arbitrary yet practical 106 cycle cut-

off value (see Table 4.1). Furthermore, certain fatigue tests were not successful due to strain

gauge debonding or short-circuiting due to saline penetration at some point during the tests.

87

For these tests, the number of cycles to failure versus applied stress data was reliable,

whereas the other parameters were eliminated from further analysis.

Table 4.2 summarizes the significant differences observed between female and male emu

bones prior to fatigue testing. Female emu tibiae had significantly higher BMD (p<0.001)

values compared to male emu tibiae for all groups. CT analysis of all samples at the mid-

diaphysial loading site before fatigue testing showed significant differences in geometrical

properties between sexes. Female emu tibiae had significantly greater thickness (p<0.001),

cortical area (p<0.001) and moment of inertia (p<0.001) compared to male emu tibiae. There

were no differences in these values between KOH treatment groups. Contrary to the BMD

data, there were no differences in initial secant modulus detected between the different

groups and between sexes (Table 4.3). Creep strain and maximum strain at failure also did

not vary between the groups (Table 4.3). KOH treatment resulted in negligible bone mass

loss (Figure 4.3) and collagen loss (Figure 4.4), similar results seen in Chapter 3 – Initial

Study.

Table 4.2: BMD and geometrical parameters of emu tibiae after KOH treatment but prior to fatigue to failure testing. KOH treatment time in days

Parameter

Sex 0 1 3 7 14

Female 1.32 ± 0.03 1.34 ± 0.03 1.30 ± 0.03 1.32 ± 0.03 1.31 ± 0.02 BMD (g/cm2) Male 1.20 ± 0.04a 1.20 ± 0.03a 1.20 ± 0.02a 1.22 ± 0.03a 1.20 ± 0.02a

Female 4.55 ± 0.14 4.72 ± 0.17 4.67 ± 0.10 4.62 ± 0.11 4.57 ± 0.10 Thickness (mm) Male 3.91 ± 0.14a 3.84 ± 0.15a 4.10 ± 0.12a 4.16 ± 0.11a 4.05 ± 0.10a

Female 292 ± 11 301 ± 13 302 ± 8 288 ± 8 288 ± 7 Cortical Area (mm2) Male 251 ± 10a 240 ± 11a 248 ± 9a 245 ± 8a 246 ± 8a

Female 15234 ± 729 14770 ± 851 13732 ± 524 14504 ± 534 14988 ± 524 Moment of Inertia (mm4) Male 11805 ± 685a 11699 ± 755a 11393 ± 631a 12255 ± 554a 12249 ± 524a

Female 11.4 ± 0.3 10.9 ± 0.3 11.0 ± 0.2 10.9 ± 0.2 11.3 ± 0.2 Distance from neutral axis (mm) Male 11.0 ± 0.2 11.0 ± 0.3 10.9 ± 0.2 10.8 ± 0.2 10.6 ± 0.2

ap ≤ 0.05 versus female

Table 4.3: Initial secant modulus, creep strain at fracture and maximum strain at fracture for fatigue to failure emu tibiae. KOH treatment time in days

Parameter

Sex 0 1 3 7 14

Female 24947 ± 1110 27687 ± 946 23460 ± 728 24674 ± 713 24115 ± 1002 Eo (MPa) Male 27821 ± 1139 25311 ± 1155 25708 ± 1221 25116 ± 1159 26412 ± 683

Female 3340 ± 793 3349 ± 562 3470 ± 1183 3927 ± 1592 4109 ± 1115 εcreep_fracture (με) Male 3394 ± 800 3181 ± 868 3736 ± 1114 3518 ± 1024 3690 ± 856

Female 7983 ± 1059 6200 ± 1221 6697 ± 1467 7407 ± 1877 7442 ± 1504 εmax_fracture

(με) Male 6457 ± 814 6378 ± 893 6791 ± 1454 6003 ± 1199 6275 ± 1140

88

89

KOH Treatment Time

1 day 3 day 7 day 14 day

Perc

ent

bone

mas

s ch

ange

(%

)

0.0

0.2

0.4

0.6

0.8

1.0 MaleFemale

Figure 4.3: Percent bone weight loss versus KOH treatment time for female and male fatigue emu tibiae. No differences were observed in the bone weight loss with KOH treatment time or between sexes.

KOH Treatment Time1 day 3 day 7 day 14 day

Perc

ent

colla

gen

loss

(%

)

0.00

0.02

0.04

0.06

0.08

0.10

MaleFemale

Figure 4.4: Percent collagen removed versus KOH treatment time for female and male fatigue emu tibiae. Significant differences were observed in the collagen removed with KOH treatment time but sex was not a factor in collagen removal. Maximum percent collagen removed was negligible (0.05%).

90

In terms of fatigue properties, there was an inverse relationship between stress and number of

cycles to failure for all KOH treatment groups. The number of fatigue cycles to failure decreased

with increasing initial stress for both sexes. The 14-day KOH treated specimens needed

significantly fewer cycles to fail at high stresses (>60 MPa) for both sexes. However, the number

of cycles to failure at low stresses (<60 MPa) was not affected by KOH treatment (Figure 4.5).

The regression slopes were only significantly different between untreated (0-day) and 14-day

KOH treatment groups for female (p=0.024) and male (p=0.019) emu tibiae. No differences were

found in slopes between sexes at all KOH treatment time points. The covariate, stress, was

significantly related to the number of cycles to failure (p<0.001) for both female and male emu

tibiae. There was a significant effect of KOH treatment on number of cycles to failure after

controlling for the effect of stress for both female (p<0.001) and male (p<0.001) emu tibiae. The

regression curves obtained from damage data were found to be statistically significant (p<0.001)

when comparing the untreated (0-day) group to the 14-day KOH treatment group for female and

male bones. No differences were found in fatigue life between sexes at all KOH treatment time

points.

Figure 4.5: Peak stress versus log(N) curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae. Both female and male 14-day KOH treated samples have a lower resistance to fatigue compared to the other groups at high stresses only (>60 MPa). The number of cycles to failure at lower stresses is not affected by KOH treatment. No differences were seen between sexes. All tests were subjected to a constant minimum stress of 5 MPa.

1 10 100 1000 10000 100000 1000000

50

100

150

2000-day1-day3-day7-day14-day

R2=0.80, p<0.001R2=0.97, p<0.001R2=0.74, p<0.001R2=0.77, p<0.001R2=0.72, p<0.001

R2=0.86, p<0.001R2=0.92, p<0.001R2=0.85, p<0.001R2=0.90, p<0.001R2=0.76, p<0.001

a) Females

Cycles to failure (N)

Pea

k S

tres

s (M

Pa)

b) Males

Cycles to failure (N)

Pea

k S

tres

s (M

Pa)

1 10 100 1000 10000 100000 10000000

50

100

150

2000-day1-day3-day7-day14-day

1 10 100 1000 10000 100000 1000000

50

100

150

2000-day1-day3-day7-day14-day

R2=0.80, p<0.001R2=0.97, p<0.001R2=0.74, p<0.001R2=0.77, p<0.001R2=0.72, p<0.001

R2=0.86, p<0.001R2=0.92, p<0.001R2=0.85, p<0.001R2=0.90, p<0.001R2=0.76, p<0.001

a) Females

Cycles to failure (N)

Pea

k S

tres

s (M

Pa)

b) Males

Cycles to failure (N)

Pea

k S

tres

s (M

Pa)

1 10 100 1000 10000 100000 10000000

50

100

150

2000-day1-day3-day7-day14-day

91

The creep strain and damage index due to fatigue loading exhibited the three characteristic

phases: an initial rapid strain increase, a steady state region and a final rapid increase to failure

(as shown in Figure 4.2). For both sexes, the damage and creep rates showed a positive

correlation with the stress range: high initial stresses resulted in high damage and creep rates

while low initial stresses resulted in low damage index and creep rates. Figure 4.6 and Figure 4.7

show the changes of damage index rate and creep rate as a function of stress, respectively. The

covariate, stress, was significantly related to the damage index rates (p<0.001 for both sexes) as

well as creep rates (p=0.002 for female and p=0.001 for male emu tibiae). Female and male 14-

day KOH treated samples exhibited accelerated damage index and creep rates at high stresses

only (>60 MPa). At low stresses (<60 MPa), female and male emu tibiae damage index and

creep rates were similar for both sexes. Comparison between groups revealed that 14-day KOH

treatment significantly accelerated the damage index rate and creep rate compared to untreated

(0-day) groups for both sexes at high stresses. There was a significant effect of KOH treatment

on damage rate after controlling for the effect of stress for female (p<0.001) and male (p=0.003)

emu tibiae. Similarly, KOH treatment significantly affected the creep rate for female (p<0.001)

and male (p=0.001) bones. No differences were found in damage index and creep rates between

sexes at all KOH treatment time points.

Figure 4.6: Peak stress versus damage index rate curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae. KOH treatment caused accelerated damage index rates at high stresses only (>60 MPa). The damage index rate at lower stresses is not affected by KOH treatment. No differences were seen between sexes. All tests were subjected to a constant minimum stress of 5 MPa.

R2=0.78, p<0.001R2=0.71, p<0.001R2=0.98, p<0.001R2=0.83, p<0.001R2=0.86, p<0.001

a) Females

Damage index rate (1/sec)

Pea

k S

tres

s (M

Pa)

b) Males

Damage index rate (1/sec)

Pea

k S

tres

s (M

Pa)

R2=0.85, p<0.001R2=0.90, p<0.001R2=0.83, p<0.001R2=0.75, p<0.001R2=0.65, p<0.001

10-910-810-710-610-510-410-310-210-11000

50

100

150

2000-day1-day3-day7-day14-day

10-910-810-710-610-510-410-310-210-11000

50

100

150

2000-day1-day3-day7-day14-day

R2=0.78, p<0.001R2=0.71, p<0.001R2=0.98, p<0.001R2=0.83, p<0.001R2=0.86, p<0.001

a) Females

Damage index rate (1/sec)

Pea

k S

tres

s (M

Pa)

b) Males

Damage index rate (1/sec)

Pea

k S

tres

s (M

Pa)

R2=0.85, p<0.001R2=0.90, p<0.001R2=0.83, p<0.001R2=0.75, p<0.001R2=0.65, p<0.001

10-910-810-710-610-510-410-310-210-11000

50

100

150

2000-day1-day3-day7-day14-day

10-910-810-710-610-510-410-310-210-11000

50

100

150

2000-day1-day3-day7-day14-day

92

Figure 4.7: Peak stress versus creep rate curves for 0-14 day KOH treatment for (a) female and (b) male emu tibiae. KOH treatment caused accelerated creep rates at high stresses only (>60 MPa). The creep rate at lower stresses is not affected by KOH treatment. No differences were seen between sexes. All tests were subjected to a constant minimum stress of 5 MPa.

4.3.2. Microdamage

Histological examination of cross sections fatigued at 2000 microstrain and 4000 microstrain

revealed no evidence of microdamage by light microscopy.

4.3.3. Effect of partial fatigue testing on BMD, SOS and EI

Partial fatigue testing to 100,000 cycles resulted in similar strain levels across all groups, with no

differences detected between sexes, nor for different KOH treatment times (Table 4.4).

Table 4.4: Strain levels during partial fatigue testing of emu tibiae. KOH treatment time in days

Parameter

Sex 0 1 3 7 14

Female 1980 ± 34 1975 ± 58 1953 ± 67 1960 ± 60 1997 ± 39

Strain (με) Male 2003 ± 20 2002 ± 57 1960 ± 48 1943 ± 56 2002 ± 52

R2=0.92, p<0.001R2=0.78, p<0.001R2=0.83, p<0.001R2=0.80, p<0.001

R2=0.87, p<0.001R2=0.73, p<0.001R2=0.61, p<0.001R2=0.86, p<0.001R2=0.60, p<0.001

R2=0.86, p<0.001

a) Females

Creep rate (1/sec)

Pea

k St

ress

(M

Pa)

b) Males

Creep rate (1/sec)

Pea

k S

tres

s (M

Pa)

10-910-810-710-610-510-410-310-20

50

100

150

200

0-day1-day3-day7-day14-day

10-910-810-710-610-510-410-310-20

50

100

150

200

0-day1-day3-day7-day14-day

R2=0.92, p<0.001R2=0.78, p<0.001R2=0.83, p<0.001R2=0.80, p<0.001

R2=0.87, p<0.001R2=0.73, p<0.001R2=0.61, p<0.001R2=0.86, p<0.001R2=0.60, p<0.001

R2=0.86, p<0.001

a) Females

Creep rate (1/sec)

Pea

k St

ress

(M

Pa)

b) Males

Creep rate (1/sec)

Pea

k S

tres

s (M

Pa)

10-910-810-710-610-510-410-310-20

50

100

150

200

0-day1-day3-day7-day14-day

10-910-810-710-610-510-410-310-20

50

100

150

200

0-day1-day3-day7-day14-day

93

The effect of KOH treatment and bone fatigue was assessed by DXA (BMD), QUS (SOS),

MRTA (EI) and three-point bending before and after partial fatigue testing (2000 με; 100,000

cycles) for all treatment groups for both female (Figure 4.8 (a)) and male (Figure 4.8 (b)) emu

tibiae. There were no differences in BMD or SOS, as measured by DXA and QUS, respectively.

In fact, the measured values after partial fatigue testing varied by less than 1% and 2% between

KOH treatment times for DXA and QUS, respectively, for both sexes. However, there were

significant changes in the modulus of elasticity in all groups, as determined by three-point

bending and by MRTA for both female and male emu tibiae. The average percent modulus

change after partial fatigue testing was similar for all groups for both sexes. In each group, tibiae

lost on average 20% of their stiffness for both sexes. No differences were seen between percent

changes in modulus of elasticity as measured by three-point bending and the MRTA for both

sexes. There were no differences detected in partial fatigue testing between sexes. CT analysis

showed no changes in geometrical properties such as moment of inertia, cortical area and

thickness after partial fatigue testing for untreated and KOH treated groups (Table 4.5) for both

sexes.

Figure 4.8: Percent changes of bone quality measurements after partial fatigue testing (2000 με; 100,000 cycles) reported by the different measurement techniques as a function of KOH treatment time for (a) female and (b) male emu tibiae. No differences were seen between sexes. Significant changes in modulus of elasticity were observed as determined by three-point bending and MRTA for both female and male emu tibiae. DXA and QUS measurements did not detect any changes caused by partial fatigue testing.

0 1 3 7 14-50

-40

-30

-20

-10

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA) % Change SOS (QUS)

0 1 3 7 14-50

-40

-30

-20

-10

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

a) Females

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

b) Males

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

0 1 3 7 14-50

-40

-30

-20

-10

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA) % Change SOS (QUS)

0 1 3 7 14-50

-40

-30

-20

-10

0

% Change Modulus (3-point)% Change Modulus (MRTA)% Change BMD (DXA)% Change SOS (QUS)

a) Females

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

b) Males

KOH Treatment (days)

Per

cen

t C

han

ge (

%)

94

Table 4.5: Average female and male emu tibiae geometrical parameter changes after KOH treatment and partial fatigue testing.

KOH treatment time in days Parameter

Sex

0 1 3 7 14

Female 0.4 ± 0.9 0.7 ± 0.9 0.3 ± 0.8 0.6 ± 0.8 0.5 ± 0.2 Change in medial-lateral diameter (%) Male 0.6 ± 0.8 0.3 ± 0.4 0.6 ± 0.9 0.5 ± 0.5 0.9 ± 0.6

Female 0.4 ± 0.7 0.3 ± 0.5 0.3 ± 0.3 0.2 ± 0.7 0.7 ± 0.4 Change in anterior-posterior diameter (%) Male 0.7 ± 0.7 0.2 ± 0.4 0.3 ± 0.2 0.3 ± 0.4 0.1 ± 0.9

Female 0.5 ± 0.8 0.8 ± 0.8 0.7 ± 0.5 0.8 ± 0.4 0.8 ± 0.7 Change in cortical thickness (%) Male 1.0 ± 0.9 0.7 ± 0.6 0.5 ± 0.8 0.6 ± 0.5 0.3 ± 0.4

Female 0.4 ± 0.6 0.5 ± 0.4 0.3 ± 0.7 0.8 ± 0.6 0.6 ± 0.5 Change in cortical area (%) Male 0.4 ± 0.4 0.3 ± 0.9 0.3 ± 0.2 0.3 ± 0.9 0.9 ± 0.6

Female 0.4 ± 0.6 0.5 ± 0.4 0.3 ± 0.2 1.0 ± 0.7 0.6 ± 0.3 Change in moment of inertia (%) Male 0.7 ± 0.6 0.6 ± 0.9 0.2 ± 0.9 0.4 ± 0.5 0.5 ± 0.6

4.4. Discussion

The objectives of this study were to determine how emu bone fatigue properties are affected by

KOH treatment, a treatment targeted to affect only the organic component of bone and to

determine if current clinical tools could detect the effects of partial fatigue testing in untreated

and KOH treated bones. Similarly to our previous experiments [19], KOH treatment did not

change the geometry or mineral content of the bone samples and a negligible amount of collagen

was extracted. However, the fatigue properties of the bone were significantly compromised for

bones of both sexes, which suggests that the mechanical contribution of the collagen was

impaired. The mechanical changes induced by partial fatigue testing were observed by the

MRTA, but were not detected with DXA or QUS.

4.4.1. Fatigue behaviour

In both sexes, KOH treated samples need significantly fewer cycles to fail at high stresses (>60

MPa). Analysis of the damage index and creep rates showed that the deterioration rate was

similar for bones of both sexes at all KOH treatment time points and KOH treatment appears to

cause an acceleration in damage index and creep rates at high stresses only (>60 MPa).

95

Cyclic creep may be an important mechanism in the fast deterioration rate of the KOH treated

bones at high stresses and strains, which leads to earlier fatigue failure of KOH treated groups.

Carter and Caler have shown that creep is the major cause of fatigue failure if the bone is loaded

in tension or at high stress (>60 MPa) while crack accumulation causes fatigue failure if the

specimen is loaded in compression or at low stress (<60 MPa) (Figure 4.9) [30]. At high stresses,

cracks will continue to propagate under a sustained load indicating a time-dependent mechanism

[31]. At lower stress intensities, cracks will arrest and cyclic loading is needed to further

propagate the crack [32]. Several studies have suggested that collagen is the phase responsible

for creep in bone [8]. Creep is thought to be a result of debonding of mineral crystals from the

organic matrix resulting in load transfer from the linearly elastic hydroxyapatite to the

viscoelastic collagen matrix, or simply the result of damage to collagen fibrils [8,33-35].

Collagen degradation and disorganization caused by KOH treatment may be the reason for the

observed altered fatigue behaviour at high stresses, since collagen is responsible for the creep

behaviour in bone.

96

0 20 40 60 80 100 120CYCLIC STRESS (MPa)

DA

MA

GE

FR

AC

TIO

N

1.0

0.8

0.6

0.4

0.2

0

INITIAL MAXIMUM STRAIN (mm/mm)0.0 0.001 0.002 0.003 0.004 0.005 0.006

109 108 107 106 105 104 103 102 10 1 0.5

CYCLES TO FAILURE

CREEP

CRACK ACCUMULATION

NORMALLOADING

Figure 4.9: Cumulative damage model showing the transition from creep to crack accumulation behaviour (adapted from [30]).

Conversely, at low stresses, the degraded and disorganized collagen due to KOH treatment may

act as additional microstructural barriers, leading to an increased amount of crack deflection and

thus, resistance to crack propagation. This may be the reason why no significant differences in

fatigue behaviour were observed for bones fatigued at low stresses. In addition, at low stresses,

sacrificial bonds may play a role in bone material properties. Sacrificial bonds are additional,

weak, but reformable bonds found within or between collagen molecules. These bonds break at

lower energies than the stronger collagen backbone bonds, serving to protect the collagen

molecule backbone by dissipating the applied energy [36,37]. The energy absorbed by breaking

the weak sacrificial bonds and the subsequent stretching of the otherwise tortuous collagen

molecules, increases the total energy required to fracture the bone. This increase in total energy

required to fracture the material increases the “toughness” of the material [36]. After the force is

reduced, these sacrificial bonds reform and the collagen molecules re-crimp, partially restoring

the original bone strength [36]. This sacrificial bond mechanism would increase the amount of

energy necessary for a crack to propagate through the bone at low stresses [14,36,37].

A study on the fatigue behaviour of Mov13 (Type I collagen mutation) cortical bone showed a

significant reduction in fatigue life under cyclic loading [38]. The authors postulated that the

97

reduced collagen content, the altered collagen-matrix organization, the increased proportion of

woven bone and the two-fold increase in bone porosity interfered with the ability of the Mov13

microstructure to deter and isolate cracks during the tissue-level fatigue tests [38]. In this study,

the fatigue behaviour of KOH treated bones was significantly affected at high stresses only (>60

MPa). This suggests a different mechanism of action of the KOH. We hypothesize that a

combination of in situ collagen degradation, which results in a modified collagen organization

and/or partial debonding between the mineral and organic matrix may be the reasons for the

varying fatigue behaviour observed in KOH treated bones.

4.4.2. Microdamage

Other effects such as microdamage accumulation may also play an important role in the altered

damage behaviour of KOH treated bones. Although the tibiae tested at 2000 microstrain and

4000 microstrain were subjected to cycles varying from 10,000 to 1,000,000 cycles and showed

reductions in stiffness values, histological examination of sections failed to reveal any evidence

of microdamage. This may be due to inadequate resolution capabilities of the light microscope

[27]. Even though microdamage was not observed at the level of light microscopy, it is possible

that microdamage may have been initiated at the ultrastructural level, as damage initiates

differently in different directions [39]. Results obtained by Forwood and Parker showed that

intensive exercise, producing approximately 30,000 loading cycles, was not sufficient to produce

evidence of microdamage in the cortex of tibiae from rats [40]. No evidence of microdamage

was observed in the tibiae of marathon runners following a similar duration of loading [41].

Based on these studies, it was assumed that microdamage analysis of the samples fatigued at low

stresses to 100,000 cycles may not have revealed any additional information and was not

performed. It has been shown that bone accumulates a considerable amount of microdamage

before failure due to cyclic fatigue loading [30]. However, evidence of microfracture in bone that

has yielded is difficult to observe [42,43]. As a result, there is no clear correlation between

microdamage and impaired mechanical properties [39]. Additional factors such as staining

protocol and intensity of the basic fuchsin accumulation may also affect microdamage evaluation

[44]. The three-point bending setup should also be reconsidered as a loading method, as this

testing evaluates the fatigue behaviour at the mid-diaphysis, forcing the crack to initiate and

propagate in the small region under the crosshead. As such, a four-point bending setup may be

98

more practical for future microdamage analysis. Finally, Burr and colleagues have shown that

bone can undergo a significant amount of modulus degradation before microcracks appear and

that the absence of any microdamage is not an indication of bone mechanical integrity [39].

4.4.3. Sex Differences

Female emu tibiae had significantly higher BMD compared to male tibiae for all groups. A

difference in BMD has been previously shown to affect fatigue properties [45]. According to the

result of that study, our samples with higher BMD would have been predicted to exhibit an

increased fatigue life than the bones with lower BMD (male bones). In terms of sexual

dimorphisms, male emus also differ from female emus in that they undergo 8-weeks of

immobilization when performing egg incubation [46]. This immobilization can result in

differences in bone behaviour. Both the lower BMD and possible immobilization factors suggest

that male emu tibiae should have decreased fatigue resistance compared to female emu tibiae.

However, sex differences in fatigue resistance were not observed in this study. This may be due

to the scatter inherent in fatigue testing of bone, or this data may indicate that male emu tibiae

have developed a mechanism to equally withstand fatigue testing as effectively as their female

counterparts. There may be species-specific differences in emu skeletal development that could

affect the fatigue behaviour of their bones. There may also be differences in the organization of

collagen of the female and male emu bones. Therefore, structural/organizational alterations due

to KOH treatment may be compromising normal fatigue damage processes.

Fatigue strength of bone has also been reported to be a function of microstructure [47,48].

Therefore, it may be hypothesized that a unique microstructure in female and male emu tibiae

may result in different fatigue behaviour. Choi et al. [47] demonstrated similar contradictory

results when fatigue testing single trabeculae and similarly sized cortical bone specimens. The

authors found that trabecular specimens had significantly lower fatigue strength compared to

cortical specimens even though trabecular bone exhibited higher mineral density. The authors

concluded that mineral density may be an insufficient estimator for modulus or fatigue strength

unless microstructural variations are accounted for when structurally different bones are being

compared [47]. This may also be the case in our study. Furthermore, another study has shown

that the effects of microstructural changes on fatigue strength were approximately five times

99

greater than those due to mineral density differences [48]. Finally, it is possible that no

difference in fatigue life was observed between the sexes because of excessive data scatter due to

other independent factors such as surface condition or flaw distribution [32].

4.4.4. Clinical tools

The measured changes in fatigue properties by three-point bending and MRTA were not detected

using DXA or QUS. Partial fatigue testing to induce stiffness loss without fracture did not alter

bone BMD and neither technique is designed to evaluate changes in bone mechanical properties.

In contrast, the MRTA closely mirrored the stiffness loss measured by three-point bending. This

is due to the ability of the MRTA to provide a direct measure of the cross-sectional bending

stiffness (EI), which reflects the elastic modulus, E, a material property, and the cross-sectional

moment of inertia, I, a geometric property determined by the distribution of material around the

bone central axis [15,19]. The MRTA has been shown to detect changes due to KOH treatment

[19] and we similarly detected differences in the MRTA measures between untreated and KOH

treated groups before partial fatigue testing.

Partial fatigue testing at 2000 microstrain revealed no differences between untreated (0-day) and

KOH treated groups for both sexes. This is in agreement with the fatigue to failure data where no

differences were observed at low stresses/strains. Differences between untreated (0-day) and

KOH treated groups for both sexes may have been observed if partial fatigue testing was

performed at higher stresses/strains, based on the fatigue to failure data. Partial fatigue testing

was not performed at higher strains/stresses as it is difficult to ensure stiffness loss and not

fracture.

The stiffness loss due to partial fatigue testing of untreated bones that were not exposed to KOH

treatment, which could be considered a model analogue for fragility fractures, was detected by

the MRTA and three-point bending. However, the loss of stiffness measured by mechanical

testing and detected by MRTA, was not detected by DXA or QUS. Therefore, the MRTA is

capable of detecting changes in bone mechanical properties due to fatigue. EI may be a better

method for clinical evaluation of fracture risk compared with a measurement of bone mineral

density alone.

100

4.5. Conclusions

These findings provide new insights into reductions in the cyclic behaviour of bone, in which

collagen has been degraded in situ. Noncollagenous proteins located at the mineral-collagen

interface may have also been adversely affected by KOH treatment. The present study also

emphasizes the ability of the MRTA to detect changes in bone mechanical properties induced by

fatigue. The MRTA may therefore be a more effective tool for predicting fracture risk than DXA

or QUS.

4.6. Chapter Summary

The effect of collagen alteration (induced by KOH treatment) was investigated on the fatigue

resistance of emu tibiae. KOH treated samples exhibited lower fatigue resistance compared to

untreated (0-day) bones at high stresses (>60 MPa) only for both sexes. There were no

differences in fatigue behaviour at low stresses (<60 MPa) for all groups and no evidence of

microdamage. Partial fatigue testing caused a decrease in modulus, on average 20%, for all

groups (including untreated groups) and for both sexes. MRTA detected this change however,

DXA and QUS did not.

The significant decrease in bone fatigue properties at high stresses only in this unique emu model

may be due to a change in the collagen structure caused by KOH treatment. As a result, it is

hypothesized that the significant changes in bone mechanical properties may be due to in situ

collagen degradation and/or partial debonding between the mineral and organic matrix.

101

4.7. References

[1] Lee TC, O'Brien FJ, Taylor D. (2000) The nature of fatigue damage in bone. Int.J.Fat. 22:847-853.

[2] Hsieh YF, Silva MJ. (2002) In vivo fatigue loading of the rat ulna induces both bone formation

and resorption and leads to time-related changes in bone mechanical properties and density. J.Orthop.Res. 20:764-771.

[3] Mori S, Burr DB. (1993) Increased intracortical remodeling following fatigue damage. Bone.

14:103-109. [4] Pattin CA, Caler WE, Carter DR. (1996) Cyclic mechanical property degradation during fatigue

loading of cortical bone. J.Biomech. 29:69-79. [5] Lee TC, Staines A, Taylor D. (2002) Bone adaptation to load: microdamage as a stimulus for

bone remodelling. J.Anat. 201:437-446. [6] Callister WD. (2003) Materials Science and Engineering: An Introduction. John Wiley & Sons.

New York. [7] Tami AE, Nasser P, Schaffler MB, Knothe Tate ML. (2003) Noninvasive fatigue fracture model

of the rat ulna. J.Orthop.Res. 21:1018-1024. [8] Bowman SM, Gibson LJ, Hayes WC, McMahon TA. (1999) Results from demineralized bone

creep tests suggest that collagen is responsible for the creep behavior of bone. J.Biomech.Eng. 121:253-258.

[9] Fois M, Lamure R, Fauran MJ, Lacabanne C. (1998) Viscoelastic properties of human bone.

J.Biomech. 31:2. [10] Wang X, Shen X, Li X, Agrawal CM. (2002) Age-related changes in the collagen network and

toughness of bone. Bone. 31:1-7. [11] Oxlund H, Barckman M, Ortoft G, Andreassen TT. (1995) Reduced concentrations of collagen

cross-links are associated with reduced strength of bone. Bone. 17:365S-371S. [12] Zioupos P, Currey JD, Hamer AJ. (1999) The role of collagen in the declining mechanical

properties of aging human cortical bone. J.Biomed.Mater.Res. 45:108-116. [13] Cefalu CA. (2004) Is bone mineral density predictive of fracture risk reduction?

Curr.Med.Res.Opin. 20:341-349. [14] Burr DB. (2002) The contribution of the organic matrix to bone's material properties. Bone. 31:8-

11. [15] Roberts SG, Hutchinson TM, Arnaud SB, Kiratli BJ, Martin RB, Steele CR. (1996) Noninvasive

determination of bone mechanical properties using vibration response: a refined model and validation in vivo. J.Biomech. 29:91-98.

102

[16] Hutchinson TM, Bakulin AV, Rakhmanov AS, Martin RB, Steele CR, Arnaud SB. (2001) Effects of chair restraint on the strength of the tibia in rhesus monkeys. J.Med.Primatol. 30:313-321.

[17] Norrdin RW, Simske SJ, Gaarde S, Schwardt JD, Thrall MA. (1995) Bone changes in

mucopolysaccharidosis VI in cats and the effects of bone marrow transplantation: mechanical testing of long bones. Bone. 17:485-489.

[18] Djokoto C, Tomlinson G, Waldman S, Grynpas M, Cheung AM. (2004) Relationship among

MRTA, DXA, and QUS. J.Clin.Densitom. 7:448-456. [19] Wynnyckyj C, Omelon S, Savage K, Damani M, Chachra D, Grynpas MD. (2009) A new tool to

assess the mechanical properties of bone due to collagen degradation. Bone. 44:840-848. [20] Willett TL, Wynnyckyj C, Wang J, Grynpas MD. The fatigue resistance of rabbit tibiae varies

with age from youth to middle age. Osteoporos.Int. May 22. Epub ahead of print. [21] Taylor D. (2000) Scaling effects in the fatigue strength of bones from different animals.

J.Theor.Biol. 206:299-306. [22] Boyce TM, Fyhrie DP, Glotkowski MC, Radin EL, Schaffler MB. (1998) Damage type and strain

mode associations in human compact bone bending fatigue. J.Orthop.Res. 16:322-329. [23] Turner CH, Burr DB. (1993) Basic biomechanical measurements of bone: a tutorial. Bone.

14:595-608. [24] Moreno LD, Waldman SD, Grynpas MD. (2006) Sex differences in long bone fatigue using a rat

model. J.Orthop.Res. 24:1926-1932. [25] Cotton JR, Winwood K, Zioupos P, Taylor M. (2005) Damage rate is a predictor of fatigue life

and creep strain rate in tensile fatigue of human cortical bone samples J.Biomech.Eng. 127: 213-219.

[26] Zioupos P, TW X, Currey JD. (1996) The accumulation of fatigue microdamage in human

cortical bone of two different ages in vitro. Clin.Biomech.(Bristol.Avon). 11:365-375. [27] Burr DB, Stafford T. (1990) Validity of the bulk-staining technique to separate artifactual from in

vivo bone microdamage. Clin.Orthop.Relat.Res. 305-308. [28] Frost HM. (1960) Presence of microscopic cracks in vivo in bone. Henry Ford Hospital Medical

Bulletin. 8:27-35. [29] Frank JD, Ryan M, Kalscheur VL, Ruaux-Mason CP, Hozak RR, Muir P. (2002) Aging and

accumulation of microdamage in canine bone. Bone. 30:201-206. [30] Carter DR, Caler WE. (1985) A cumulative damage model for bone fracture. J.Orthop.Res. 3:84-

90. [31] Nalla RK, Kruzic JJ, Kinney JH, Ritchie RO. (2005) Aspects of in vitro fatigue in human cortical

bone: time and cycle dependent crack growth. Biomater. 26:2183-2195.

103

[32] Kruzic JJ, Ritchie RO. (2008) Fatigue of mineralized tissues: cortical bone and dentin. J.Mech.Beh.Biomed.Mat. 3-17.

[33] Cotton JR, Zioupos P, Winwood K, Taylor M. (2003) Analysis of creep strain during tensile

fatigue of cortical bone. J.Biomech. 36:943-949. [34] Fondrk MT, Bahniuk EH, Davy DT. (1999) A damage model for nonlinear tensile behavior of

cortical bone. J.Biomech.Eng. 121:533-541. [35] Rimnac CM, Petko AA, Santner TJ, Wright TM. (1993) The effect of temperature, stress, and

microstructure on the creep of compact bovine bone. J.Biomech. 26:219-228. [36] Fantner GE, Hassenkam T, Kindt JH, Weaver JC, Birkedal H, Pechenik L, Cutroni JA, Cidade

GA, Stucky GD, Morse DE, Hansma PK. (2005) Sacrificial bonds and hidden length dissipate energy as mineralized fibrils separate during bone fracture. Nat.Mater. 4:612-616.

[37] Thompson JB, Kindt JH, Drake B, Hansma HG, Morse DE, Hansma PK. (2001) Bone

indentation recovery time correlates with bond reforming time. Nature. 414:773-776. [38] Jepsen KJ, Schaffler MB, Kuhn JL, Goulet RW, Bonadios J, Goldstein SA. (1997) Type I

collagen mutation alters the strength and fatigue behavior of MOV13 cortical tissue. J.Biomech. 11/12:1141-1147.

[39] Burr DB, Turner CH, Naick P, Forwood MR, Ambrosius W, Hasan MS, Pidaparti R. (1998) Does

microdamage accumulation affect the mechanical properties of bone? J.Biomech. 31:337-345. [40] Forwood, MR, Parker AW. (1989) Microdamage in response to repetitive torsional loading in the

rat tibia. Calcif.Tissue.Int. 45:47-53. [41] Rubin CT, Pratt GW, Porter AL, Lanyon LE, Poss R. (1987) The use of ultrasound in-vivo to

determine acute change in the mechanical properties of bone following intense physical activity. J.Biomech. 20:723-727.

[42] Currey JD. (1984) The mechanical adaptations of bones. Princeton University Press. Princeton. [43] Forwood, MR, Parker AW. (1991) Repetitive loading, in vivo, of the tibiae and femora of rats:

effects of repeated bouts of treadmill-running. Bone.Min. 13:35-46. [44] Huja SS, Sayeed Hasan M, Pidaparti R, Turner CH, Garetto LP, Burr DB. (1999) Development of

a fluorescent light technique for evaluating microdamage in bone subjected to fatigue loading. J.Biomech. 32:1243-1249.

[45] Jarvinen TL, Kannus P, Pajamaki I, Vuohelainen T, Tuukkanen J, Jarvinen M, Sievanen H.

(2003) Estrogen deposits extra mineral into bones of female rats in puberty, but simultaneously seems to suppress the responsiveness of female skeleton to mechanical loading. Bone. 32:642-651.

[46] Davies SJJF. (2002) Ratites and Tinamous. Oxford University Press. New York. [47] Choi K, Goldstein SA. (1992) A comparison of the fatigue behavior of human trabecular and

cortical bone tissue. J.Biomech. 25:1371-1381.

104

[48] Carter DR, Hayes WC. (1976) Fatigue life of compact bone-I. Effects of stress amplitude, temperature and density. J.Biomech. 9:27-37.

105

Chapter 5 Fractography 5.1. Introduction

5.2. Experimental Details

5.2.1. Emu bone samples and KOH treatment 5.2.2. Fractography analysis 5.2.3. Surface roughness measurements 5.2.4. Statistical analysis

5.3. Results

5.3.1. Tensile versus compressive areas 5.3.2. Degree of roughness 5.3.3. Regions of interest 5.3.4. Correlations

5.4. Discussion

5.4.1. Tensile versus compressive areas 5.4.2. Degree of roughness 5.4.2. Failure mechanisms

5.5. Conclusions

5.6. Chapter Summary

5.7. References

106

The following manuscript has been submitted to the Journal of Bone and Mineral

Metabolism for review in May 2010 under, “Fracture surface analysis in order to investigate

failure mechanisms of collagen degraded bone”. The complete list of authors includes: C.

Wynnyckyj, L. Wise-Milestone, S. Omelon, Z. Wang and M.D. Grynpas. In this chapter, we

use a fractography protocol previously established in our laboratory to explain the observed

increase in toughness from Chapter 3.

5.1. Introduction

Fracture surface analysis, or fractography, is the microscopic examination of failure surfaces

in an effort to gain insight about the events implicated in fracture. This technique can be

applied to the investigation of the mechanisms of bone failure [1]. Several studies have used

scanning electron microscopy (SEM) to study the fracture behaviour of bone [1-7]. These

investigators concluded that fracture surface morphology is essential in understanding how

bone fails [1-7]. Fractography has been used to help explain altered mechanical properties.

Piekarski attempted to correlate the energy required to propagate a crack by observing the

mechanism of fracture in bone on a microscopic scale [4]. Similarly, Saha and Hayes

correlated the tensile impact strength to fracture surface morphologies of human and bovine

compact bone [8]. Recently, Wise et al. showed the analytical value in quantifying the rough

and smooth areas of the fracture surfaces of mouse bone [9]. Regarding the collagen

component specifically, Jepsen et al. performed fracture surface analysis to help explain

decreased toughness in a mouse model of Osteogenesis Imperfecta [10]. Similarly, Fantner et

al. related fracture surface analysis of trabecular bone in which the organic matrix had been

degraded to the resulting modified mechanical properties [11]. Finally, George et al.

identified failure modes in bone using fractographic analysis [12].

Bone has several toughening mechanisms, such as fiber bridging, crack deflection and

microcracks that deflect, slow or stop crack propagation and increase the resistance to

fracture and ultimately failure [7,13,14]. Collagen fibrils can act as a toughening mechanism,

whereby intact fibers bridge a crack and oppose crack opening [7,13]. Similarly, uncracked

ligaments may serve to bridge opposing sides of a crack, consequently blunting further crack

propagation along the crack path, a phenomenon known as ‘crack bridging’ [15,16]. Cement

107

lines and interlamellar boundaries are believed to provide weak interfaces capable of crack

deflection, thereby prolonging the crack propagation path and accordingly increasing bone

toughness [7,17]. Finally, microdamage from microcracking works through crack-tip

shielding [7].

To study the contribution of collagen to the mechanical properties of bone, we developed a

model using the emu tibiae, in which collagen was degraded with 1 M potassium hydroxide

(KOH) while keeping the mineral content of bone unaltered [18]. While the KOH treatment

is not physiological, its role is to improve our understanding of the mechanisms by which

collagen degradation, not removal, affects bone mechanical properties. Endocortical KOH

treatment results in negligible mass loss (0.5%), collagen loss (0.05%), no differences in

geometrical parameters, but significant changes in mechanical properties [18]. Specifically,

male and female emu tibiae showed significant decreases in failure stress and increased

failure strain and toughness with increasing KOH treatment time [18].

The reason for increased failure strain and toughness values of KOH treated bones remains

unclear and in fact, contradicts other work where degraded collagen induces a more brittle-

like behaviour in bone, with decreased toughness values [10,11]. Therefore, we hypothesize

that in situ collagen degradation and/or partial debonding between the mineral and organic

matrix may be responsible for the increased toughness observed with KOH treated bones.

The goal of the present study was to gain insights into the failure mechanisms of female and

male KOH treated emu tibiae in an effort to explain the previously observed mechanical

property changes. A more thorough understanding of the relationship between fracture

surface morphologies and bone toughness can provide new insights into the underlying

mechanisms of bone failure.

5.2. Experimental Details

5.2.1. Emu bone samples and KOH treatment

The emu bone tibiae samples from Chapter 3 – Initial Study, were used in this study. Briefly,

skeletally mature (3-5 years of age) whole emu tibiae were separated from the femora and

108

tarsometatarsi with a scalpel. The shaft of each tibia was then isolated by using a circular saw

to remove the ends (15% of the total bone length from the proximal end and 10% from the

distal end), resulting in bone samples 26 to 32 cm in length. The marrow and trabecular bone

from the diaphysis of the tibiae were removed by drilling longitudinally through the bone

shaft, after which the medullary canal was flushed with tap water [18]. The emu tibiae were

filled with 1 M KOH and the ends were sealed with polymethylmethacrylate (PMMA, SR

Ivolen Liquid and Powder, Ivoclar Vivadent, Mississauga, ON). The bones were held

horizontally over a collection basin and kept moist with a 0.9% saline solution drip. The

bones were rotated 180 ° around the axial axis every 12 hours during the KOH treatment.

After the desired treatment period, the KOH solution was reclaimed, its volume measured

and the endocortical surfaces of the bones rinsed in running tap water for one hour.

Emu tibiae were tested in three-point bending using a servo-hydraulic materials testing

machine (Model 8511, Instron, Canton, MA). Each tibia was placed posterior side down on

the lower supports, 24 cm apart [18]. Anterior-posterior loading is the most commonly used

approach for anatomical and practical reasons. The bone was secured in the three-point

bending jig with a 100 N pre-load and then loaded to failure at a displacement rate of 0.04

mm/s [18]. A load-displacement curve was created for every sample and stiffness, failure

load, failure displacement and energy to failure were determined. Load-displacement data

were normalized using geometric data from Computed Tomography (CT) scans taken at the

mid-point of each emu tibia with an Aquilion 64 CT scanner (Toshiba, Canada) before and

after the different KOH treatment time points. Stress (σ; MPa) and strain (ε; %) were then

calculated using the following equations [19]:

where F is the measured load (N), L is the span of the lower supports (24 cm), APφ is the

external diameter in the anterior-posterior direction (mm), Ixx is the moment of inertia about

the mediolateral axis (mm4), and D is the measured displacement (mm). Stress-strain curves

were created from the normalized data. From these curves, the elastic modulus (E; GPa),

failure stress (σf; MPa), failure strain (εf, %) and toughness (U; mJ/mm3) were determined.

xx

AP

ILF

σ⋅⋅

= 10062 ∗=

LD APφ

ε

109

Bones were divided into two groups of female and male tibiae with ten samples allocated to

each of untreated (0-day) or 14-day KOH treatments. An additional ten female and ten male

bones were filled with saline instead of KOH for 14 days to act as controls.

5.2.2. Fractography analysis

For each sample, the fracture surfaces were cut from the rest of the bone (resulting in twenty

mm long samples) and quantified using digital images (taken by a Nikon D40 SLR digital

camera equipped with a 18-105mm macro lens from Sigma) to assess failure mechanisms of

untreated and 14-day KOH treated emu tibiae. Representative digital images of the fracture

surfaces of untreated and 14-day KOH treated sample are shown in Figure 5.1. The defined

‘tensile’ (T), ‘compressive’ (C) and ‘transition’ (Tr) regions are identified. Bone is known to

be weaker in tension than in compression. Therefore, failure in three-point bending should

typically occur primarily on the tensile side [19]. The terms ‘tension’ and ‘compression’ are

used with the understanding that purely tensile or compressive regions are uncommon in

three-point bending tests of bone [19]. The ‘tensile’ side may also experience shearing and

the ‘compressive’ side may have a portion of tensile failure [9]. From these images, areas of

‘tension’, ‘compression’ and ‘transition’ (areas that did not correspond to obvious tensile or

compressive areas) were identified, using a qualitative protocol previously defined in our

laboratory [9]. Bones were consistently loaded in the anterior-posterior direction, resulting in

medial-lateral regions that could not be defined as ‘tension’ or ‘compression’ and were

therefore termed ‘transition’ regions. Image J (ImageJ 1.28u, National Institutes of Health)

was used to trace the appropriate regions and calculate the relative area. ‘Tensile’ regions

were defined as having relatively flat, smooth surfaces, whereas ‘compressive’ regions were

defined by areas of longitudinal splitting and interlamellar cleavage [9] (Figure 5.1). These

definitions are based on previous observations by Wise et al. [9] that were originally

identified by Jepsen et al. [10] of tensile and compressive sides from four-point bending of

mouse femora.

110

Figure 5.1: Representative digital images of female and male, untreated and 14-day KOH treated fracture surfaces showing defined ‘tensile’ (T), ‘compressive’ (C) and ‘transition’ (Tr) regions.

Next, representative samples from the middle of the fracture surfaces (approximately 5 mm

long) from the compressive and tensile regions were cut, lyophilized, glued to individual

SEM stages with epoxy resin (5 min epoxy, Lepage’s Limited, Brampton, ON) and

subsequently gold-coated in preparation for scanning electron microscopy (SEM) imaging

(XL300ESEM, FEI, Hillsboro, OR). Beam conditions were set at 20 kV accelerating voltage

and a spot size of 4. Samples with debris on the fracture surface or unusual fractures were

excluded from the analysis. The samples were then qualitatively categorized as either

‘smooth’, ‘rough’ or ‘indeterminate’. ‘Indeterminate’ areas were defined as areas that could

not be classified as either ‘smooth’ or ‘rough’. On the tensile side, ‘smooth’ areas had a

relatively clean and flat morphology, whereas ‘rough’ areas displayed a ragged, uneven,

coarse surface [9]. Representative tensile fracture surfaces of male 14-day KOH treated

sample are shown in Figure 5.2 indicating the defined ‘rough’ and ‘smooth’ areas. On the

compressive side, ‘smooth’ areas were identified as having clean, layered lamellae, while

‘rough’ areas exhibited a fragmented, flaky, separated and jagged appearance [9].

Representative compressive fracture surfaces of female 14-day KOH treated sample are

111

shown in Figure 5.3 indicating the defined ‘rough’ and ‘smooth’ areas. These images were

assumed to be representative of the whole surface of each sample.

Figure 5.2: Representative tensile fracture surfaces of male 14-day KOH treated emu tibia showing ‘smooth’ (S) and ‘rough’ (R) regions. ‘Smooth’ regions appear as a flat and clean surface whereas ‘rough’ regions are characterized by an irregular and coarse surface.

Figure 5.3: Representative compressive fracture surfaces of female 14-day KOH treated emu tibia showing ‘smooth’ (S) and ‘rough’ (R) regions. ‘Smooth’ regions are defined as having clean, blunted, layered lamellae whereas ‘rough’ regions have a separated, fragmented and jagged appearance.

112

Percent areas, in the ‘tensile’ and ‘compressive’ areas individually, were then calculated

using Image J (ImageJ 1.28u, National Institutes of Health) to estimate the degree of

‘roughness’ for each sample. ‘Indeterminate’ areas were omitted from the calculations and

therefore the total relative area was calculated as either ‘rough’ or ‘smooth’.

These fracture surfaces were also examined at a higher magnification (1600X) for

characteristic features of bone toughening mechanisms such as microcracks, uncracked

ligament bridging, fiber bridging and crack deflection (Figure 5.4) [14].

Microcracking Uncrackedligament bridging

Crack bridging Crack deflection

(a) (b)

(c) (d) Figure 5.4: Schematic diagram of bone toughening mechanisms: (a) microcracking; (b) uncracked ligament bridging; (c) crack bridging by collagen fibers and (d) crack deflection by osteons [14].

Finally, the number of pores per area (mm-2) and number of small and large pores per area

(mm-2) on the ‘tensile’ side of fracture surfaces were determined using a point-counting

method in Image J. These calculations were not performed on the ‘compressive’ side of

fracture surfaces due to the fragmented and layered nature of these fracture surfaces. A 47

μm diameter threshold was used for the separation between small and large pores. This

threshold value was determined after analyzing a histogram of the overall pore area, which

showed two maximum points (bi-modal distribution). The lowest point of the ‘valley’

between the two maxima was defined as the threshold value.

113

5.2.3. Surface roughness measurements

To validate our qualitative definitions of ‘smooth’ and ‘rough’ fracture surfaces, samples

were assessed with a Veeco Surface Roughness Wyko non-contact optical profiler (AN505-

3-0902, Veeco Instruments Inc., Tucson, AZ). Profiler conditions were set for vertical

scanning interferometry (VSI) mode, 52X magnification. Areas that were previously defined

as ‘rough’ and ‘smooth’ were identified using a referenced SEM mapped image [9]. Ten

measurements were taken in each area per sample. Surface roughness profiles were generated

and analyzed using the accompanying software (WYKO Vision 32 for NT-2000; version

2.2.10). For each ‘smooth’ and ‘rough’ area, the average roughness (Ra, μm) was

determined, which represents the arithmetic average of all deviations from the mean centre

line of the roughness profile. This measurement is the most commonly reported parameter in

studies on surface roughness of bone implant materials [20,21].

5.2.4. Statistical analysis

Statistical analysis was performed using SPSS (SPSS 16.0 for Windows, Chicago, IL)

statistical analysis software. Tests for normality and equality of variances were initially

performed to determine whether parametric or non-parametric t-tests should be used. Two-

way analysis of variance (ANOVA, general linear model) was performed to examine the

effects of sex and KOH treatment on all measured parameters in the emu model and multiple

comparisons using a post-hoc test (Fisher’s Least Significant Difference (LSD)) were made

to detect significant differences between groups. In addition, multiple regression analyses

were used to explore the correlations of bone mechanical properties (elastic modulus, failure

stress, failure strain and toughness) from Chapter 3 (Objective 1 – Initial Study) with the

degree of roughness. Differences between degree of roughness with bending mechanical

parameters were investigated using a generalized linear model to determine any differences

between sexes. All data are presented as mean ± standard error of the mean. A confidence

level of 95% (p=0.05) was considered statistically significant for differences and

correlations.

114

5.3. Results

Significant differences were found in the fracture surface analysis parameters measured

between the untreated (0-day) and 14-day KOH treated groups. No differences were found

between sexes. All parameters measured for the control group (14-day filled with saline)

were similar to those of the untreated (0-day) groups (Appendix A) for both female and male

emu tibiae. Representative stress-strain curves for untreated (solid line) and 14-day KOH

treated (dashed line) emu tibiae (Figure 5.5) reveal changes in the plastic region with KOH

treatment time. The failure stress significantly decreased after 14-day KOH treatment. A

significant increase in failure strain and toughness was seen after 14-day KOH treatment.

Figure 5.5: Representative stress-strain curves for untreated (solid line) and 14-day KOH treated (dashed line) emu tibiae. No apparent changes in the elastic region with KOH treatment time. Significant differences were seen in the plastic region with KOH treatment time.

5.3.1. Tensile versus compressive areas

Representative digital images of the fracture surfaces of female and male untreated and 14-

day KOH treated samples are shown in Figure 5.1. Areas within sample images were

categorized as being ‘tensile’ (T), ‘compressive’ (C) or ‘transition’ (Tr). For both untreated

and 14-day KOH treated female and male groups, there was a significantly greater amount of

‘tensile’ areas compared to ‘compressive’ areas, which is consistent with bending tests [22].

Fracture surfaces of untreated female emu tibiae exhibited 50 ± 2% ‘tensile’ versus 34 ± 1%

115

‘compressive’ (p<0.001) with 15 ± 2% ‘transition’ regions while those of untreated male emu

tibiae exhibited 50 ± 2% ‘tensile’ versus 38 ± 2% ‘compressive’ (p<0.001) with 12 ± 2%

‘transition’ regions. Similarly, fracture surfaces of 14-day KOH treated female emu tibiae

exhibited 50 ± 1% ‘tensile’ versus 37 ± 1% ‘compressive’ (p<0.001) with 13 ± 1%

‘transition’ regions while those of 14-day KOH treated male emu tibiae exhibited 49 ± 1%

‘tensile’ versus 37 ± 1% ‘compressive’ (p<0.001) with 14 ± 1% ‘transition’ regions. All four

groups experienced more ‘tensile’ regions than ‘compressive’ regions. There were no

differences between sexes or with KOH treatment time.

5.3.2. Degree of roughness

Next, the ratios of ‘rough’ to ‘smooth’ areas were then compared for each fracture side

(tensile and compressive, independently) and for each group (female and male, untreated and

14-day KOH treatment), using the previously defined qualitative characterizations of

‘smooth’ and ‘rough’ (Figure 5.2 and Figure 5.3). Indeterminate areas were excluded from

total relative area calculation. Male untreated samples displayed similar percentages of

‘rough’ and ‘smooth’ areas in both tensile and compressive sides (Table 5.1). Specifically, on

the tensile side, untreated male emu tibiae exhibited a non-significant 50:50 ratio of ‘rough’

to ‘smooth’ areas; on the compressive side, a non-significant 44:56 ratio of ‘rough’ to

‘smooth’ areas was also observed. Similarly, untreated female emu tibiae exhibited a non-

significant 48:52 ratio of ‘rough’ to ‘smooth’ areas on the tensile side and a non-significant

46:54 ratio of ‘rough’ to ‘smooth’ areas on the compressive side (Table 5.1). Conversely,

female and male 14-day KOH treated fracture surfaces exhibited significantly more ‘rough’

areas compared to ‘smooth’ areas (Table 5.1). Specifically, on the tensile side, 14-day KOH

treated male tibiae exhibited a significant (p<0.001) 68:32 ratio of ‘rough’ to ‘smooth’ areas;

on the compressive side, a significant (p=0.002) 60:40 ratio of ‘rough’ to ‘smooth’ areas was

observed. Similarly, female 14-day KOH treated emu tibiae exhibited a significant (p=0.002)

62:38 ratio of ‘rough’ to ‘smooth’ areas on the tensile side and a significant (p=0.014) 58:42

ratio of ‘rough’ to ‘smooth’ areas on the compressive side (Table 5.1).

The percentages of ‘rough’ areas between the different treatment groups (untreated vs. KOH

treated) and between sexes (female vs. male) for each failure side (compressive or tensile)

116

were also compared (Table 5.1). No differences were seen in percentage of ‘roughness’

between sexes. However, significant differences were observed in percentage of ‘roughness’

with KOH treatment time for both tensile and compressive sides (Table 5.1). The percentage

of ‘roughness’ was significantly different between untreated and 14-day KOH treatment

groups for female (p=0.047) and male (p=0.031) emu tibiae on the tensile side with 14-day

KOH treated groups having a ‘rougher’ surface. Similarly, on the compressive side, 14-day

KOH treated groups had significantly greater percentages of ‘roughness’ for female

(p=0.035) and male (p=0.017) emu tibiae.

To validate the qualitative definitions of ‘rough’ versus ‘smooth’ regions, average roughness

values of regions within tensile failure sides were measured using an optical profiler. The

average roughness value of regions defined as ‘rough’ was significantly (p<0.001, Table 5.1)

rougher compared to the average roughness value of regions defined as ‘smooth’ for both

untreated and 14-day KOH treated groups. A significant difference was also noted in the

average roughness values of ‘rough’ surfaces between the untreated and 14-day KOH treated

groups (p<0.001, Table 5.1).

117

Table 5.1: Percent area roughness and surface roughness (profiler) measurements for female and male untreated and 14-day KOH treated bones.

Males Females Failure Mode Characteristic

Untreated 14-day Untreated 14-day

%Rough (%) 50 ± 6 68 ± 6a, A 48 ± 5 62 ± 2c, C Tensile Side (TS) %Smooth (%) 50 ± 6 32 ± 6a, A 52 ± 5 38 ± 2c, C

%Rough (%) 44 ± 3 60 ± 5b, B 46 ± 4 58 ± 4d, D Compressive Side (CS) %Smooth (%) 56 ± 3 40 ± 5b, B 54 ± 4 42 ± 4d, D

Males Females Failure

Mode Validation Untreated 14-day Untreated 14-day

Rough (Ra, μm) 11.1 ± 0.4e 13.9 ± 0.9f,E 11.3 ± 0.6g 13.7 ± 0.8h,F Tensile Side (TS) Smooth (Ra, μm) 5.6 ± 0.3e 5.2 ± 0.7f 4.9 ± 0.5g 5.0 ± 0.5h Lower case letters (a,b,c,d) denote significance ( ≤ 0.05) between two marked groups. Uppercase letters (A,B,C,D) denote significance (p≤ 0.05) versus untreated group within each sex.

5.3.3. Regions of interest

Regions of interest on both the tensile and compressive sides were examined at higher

magnification (1600X) with particular focus on the differences observed between untreated

and 14-day KOH treated samples (Figure 5.6). On the tensile side, both female and male

untreated samples showed a relatively ‘smooth’ surface with the presence of deflected cracks

(Figure 5.6 (a), (b), (e) and (f)). On the other hand, both female and male 14-day KOH

treated samples showed a much ‘rougher’ surface coupled with deflected cracks, microcracks

and uncracked ligament bridging (white circle) on the tensile side (Figure 5.6 (c), (d), (g) and

(h)).

118

Figure 5.6: Representative SEM images of (a) untreated male, (c) 14-day KOH treated male, (e) untreated female and (g) 14-day KOH treated female tensile emu tibiae fracture surfaces. Magnified regions of interest are shown in (b) untreated male, (d) 14-day KOH treated male, (f) untreated female and (h) 14-day KOH treated female emu tibiae. A relatively smooth surface can be seen in untreated samples for both sexes, whereas a rougher surface and the presence of toughening mechanisms such as microcracks and uncracked ligament bridging (white circle shown in (h)) are apparent in 14-day KOH treated samples for both sexes.

h)

119

Both female and male untreated samples exhibited a typical clean delamination and layered

structure with no apparent fiber bridging on the compressive side. The surfaces appear

‘smooth’ with tightly packed lamellae, suggesting a brittle-like fracture. In contrast, both

female and male 14-day KOH treated samples showed a coarser and ‘rougher’ appearance

with apparent fiber bridging and connections on the compressive side. Figures 5.7 and 5.8

represent regions of interest on the compressive side for male (Figure 5.7) and female (Figure

5.8) emu tibiae. High magnification images show crack bridging by collagen fibrils (Figure

5.7 (e) for male and Figure 5.8 (f) for female emu tibiae). Fibers appear to be disjointed with

increased splitting and separation between fibers. Further examination between separated

lamellae reveals interwoven mineralized collagen fibers (Figure 5.7 (d) and 5.7 (f)),

indicative of a more ductile-like fracture. These interwoven mineralized collagen fibers are

tightly connected in the left side of Figure 5.7 (d) and appear to be stretched and separated in

the right side of Figure 5.7 (d).

Porosity measurements revealed no differences in large pore density (pore diameter > 47 μm)

and small pore density (pore diameter < 47 μm) between sexes and between untreated and

14-day KOH treated groups (Table 5.2).

Table 5.2: Porosity parameters from tensile side of female and male emu tibiae fracture surfaces.

Males Females Parameter

Untreated 14-day Untreated 14-day

No. of large pores (pore diameter >47 μm) per area (mm-2) 0.8 ± 0.1 1.0 ± 0.3 1.0 ± 0.1 1.2 ± 0.5

No. of small pores (pore diameter<47 μm) per area (mm-2) 75 ± 8 66 ± 7 71 ± 5 70 ± 7

120

Figure 5.7: Representative SEM images of male compressive fracture surfaces of (a) untreated and (c) 14-day KOH treated emu tibiae showing interlamellar cleavage and splitting. Magnified regions of interest are shown in (b) untreated and (d) 14-day KOH treated sample. Characteristic features of bone toughening are shown in (e) and (f) 14-day KOH treated sample. Interlamellar short fibers acting as connections between lamellae are evident in the 14-day KOH treated sample (e), but absent in the untreated sample (b). Interwoven mineralized collagen fibers remain connected upon further examination between separated lamellae (d) and (f).

121

Figure 5.8: Representative SEM images of female compressive fracture surfaces of (a) untreated and (c) 14-day KOH treated emu tibiae showing interlamellar cleaving and splitting. Magnified regions of interest are shown in (b) untreated and (d) 14-day KOH treated sample. Characteristic features of bone toughening are shown in (e) and (f) 14-day KOH treated sample. Interlamellar connecting fibers between lamellae are evident in the 14-day KOH treated sample (e) and (f) but absent in the untreated samples (b).

5.3.4. Correlations

Linear regression analyses were used to explore the correlations of bone mechanical

properties (elastic modulus, failure stress, failure strain and toughness) from Chapter 3 –

Initial Study, with the fracture surface analysis results. Significant correlations were found

between degree of roughness and mechanical properties (Table 5.3). Multiple regression

analyses indicated that the percentage of ‘roughness’ on both the tensile (Figure 5.9) and

compressive (Figure 5.10) sides had strong correlations with the mechanical properties

(elastic modulus, failure stress, failure strain and toughness) for both female and male emu

122

tibiae. The elastic modulus and failure strength of the KOH-treated groups decreased with

increasing percent of ‘rough’ area. The failure strain and toughness increased with increasing

percent of ‘rough’ area. The regression curves were not different between female and male

emu tibiae in any of these tests.

Table 5.3 .Correlations between bone fracture surface features and mechanical properties for female and male emu tibiae. Failure

Strain (%) Failure Stress

(MPa) Toughness (mJ/mm3)

Modulus (GPa)

p-value 0.011 0.028 0.011 0.016 %Rough-TS

R2 0.373 -0.401 0.319 -0.406

p-value 0.036 0.009 0.004 0.006 Males

%Rough-CS R2 0.450 -0.295 0.420 -0.492

p-value 0.006 0.002 0.004 0.003 %Rough-TS

R2 0.402 -0.353 0.414 -0.384

p-value 0.025 0.006 0.007 0.002 Females

%Rough-CS R2 0.429 -0.248 0.365 -0.341

123

Figure 5.9: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the relative ‘roughness’ area on the tensile fracture surface (%Rough-TS) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Significant and strong correlations were found between these parameters.

0 20 40 60 80 1000

1

2

3

4

5 Male

Female

0 20 40 60 80 1000

1

2

3

4

5Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.37p=0.011

R2=0.40p=0.006

R2=0.32p=0.011

R2=0.41p=0.004

Percent Rough - TS (%) Percent Rough - TS (%)0 20 40 60 80 100

0

1

2

3

4

5 Male

Female

0 20 40 60 80 1000

1

2

3

4

5Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.37p=0.011

R2=0.40p=0.006

R2=0.32p=0.011

R2=0.41p=0.004

Percent Rough - TS (%) Percent Rough - TS (%)

0 20 40 60 80 1000

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Percent Rough - TS (%)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.41p=0.016

R2=0.38p=0.003 R2=0.35

p=0.002

R2=0.40p=0.028

Percent Rough - TS (%)0 20 40 60 80 100

0

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Percent Rough - TS (%)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.41p=0.016

R2=0.38p=0.003 R2=0.35

p=0.002

R2=0.40p=0.028

Percent Rough - TS (%)

124

Figure 5.10: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the relative ‘roughness’ area on the compressive fracture surface (%Rough-CS) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Significant and strong correlations were found between these parameters.

5.4. Discussion

The purpose of this study was to investigate the microscopic morphology of the three-point

bending fracture surfaces of untreated and 14-day KOH treated female and male emu cortical

bone (from Chapter 3 – Initial Study) in an effort to understand the altered mechanical

properties previously observed. Analysis of the fracture surfaces can help the interpretation

of mechanical properties and failure behaviours.

0 20 40 60 80 1000

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Percent Rough - CS (%)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.49p=0.006

R2=0.34p=0.002

R2=0.30p=0.009

R2=0.25p=0.006

Percent Rough - CS (%)0 20 40 60 80 100

0

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Percent Rough - CS (%)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.49p=0.006

R2=0.34p=0.002

R2=0.30p=0.009

R2=0.25p=0.006

Percent Rough - CS (%)

0 20 40 60 80 1000

1

2

3

4

5

Male

Female

0 20 40 60 80 1000

1

2

3

4

5 Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.45p=0.036

R2=0.43p=0.025

R2=0.42p=0.004

R2=0.37p=0.007

Percent Rough - CS (%) Percent Rough - CS (%)0 20 40 60 80 100

0

1

2

3

4

5

Male

Female

0 20 40 60 80 1000

1

2

3

4

5 Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.45p=0.036

R2=0.43p=0.025

R2=0.42p=0.004

R2=0.37p=0.007

Percent Rough - CS (%) Percent Rough - CS (%)

125

A significant increase in failure strain was seen with KOH treatment time due to a significant

increase in the post-yield strain for both female and male emu tibiae. This increased post-

yield strain contributed to a significant increase in toughness. The difference in failure strain

in the elastic region between untreated and 14-day KOH treated groups was minimal.

Therefore, the total energy to failure increased due to the accumulation of plastic damage.

These results are in agreement with the fracture surface analysis results.

Crack morphology was visualized qualitatively by taking SEM images of the tensile and

compressive surfaces of untreated and 14-day KOH treated samples. For both sexes, 14-day

KOH treated fracture surfaces exhibited significantly more ‘rough’ areas compared to

‘smooth’ areas on both the tensile and compressive areas. Similarly, the percentage of

‘roughness’ was significantly different between untreated and 14-day KOH treatment groups

for female and male emu tibiae on both the tensile and compressive sides, with 14-day KOH

treated groups exhibiting a relatively larger, ‘rougher’ appearance. These results were

confirmed with the optical profiler surface roughness measurements.

In terms of failure mechanisms, 14-day KOH treated female and male fracture surfaces

showed additional toughening mechanisms that were absent in the untreated samples,

including crack deflection, microcracks, crack bridging by collagen fibers and uncracked

ligaments. These fracture surface features are all indicative of a more ductile-like fracture.

5.4.1. Tensile versus compressive areas

Both female and male, untreated and 14-day KOH treated groups exhibit significantly larger

percent tensile surfaces (approximately 66% percentage tensile areas) compared to

compressive surfaces. This is in agreement with three-point bending fracture behaviour of

most materials. In most materials, tensile strengths are lower than compressive strengths

therefore, materials will initially fail on the tensile side during bending tests. During three-

point bending, one half experiences compression (in this case, anterior side), while the other

half experiences tension (posterior side). As the fracture progresses, the neutral axis shifts

upwards towards the original compressive side, resulting in a dynamic transition from

compressive to tensile regions. Therefore, initial compression regions become tensile regions

126

[9,19], which results in the ‘majority’ of the fracture surface being largely generated in

tension.

5.4.2. Degree of roughness

Examination of the fracture surfaces revealed distinctive differences in appearance between

the untreated and 14-day KOH treated specimens. While the untreated specimens had

‘smoother’ fracture surfaces, the 14-day KOH treated specimen fracture surfaces exhibited a

fibrous and ‘rougher’ appearance.

Rough fracture surfaces have been correlated to increased toughness of bone [4,16]. The

toughness of bone is the capability of bone to absorb energy during the failure process [19].

The degree of roughness observed on a fracture surface is representative of the resistance of a

propagating crack. Therefore, a rougher fracture surface indicates a higher crack propagation

resistance and more energy was required to further propagate the crack. Currey and Brear

[23] have observed that tissues with greater toughness display a rougher fracture surface

profile.

Fracture surface roughness is affected not only by changes in the bone structure but also by

how the loads are applied. In addition, the process of multiple initiation of cracks can result

in enhanced roughness since the majority of cracks in bone are loaded under mixed-mode

(tension and shear) conditions [24]. To avoid shear stresses in bend testing, the gauge length

(span) between the lower supports should ideally be 16 times the thickness of the specimen

[19]. This ideal span is not feasible in emu tibiae due to anatomical reasons. As a result, a

large amount of shear loading is likely present in three-point bending of the emu tibiae.

Specifically, the shear deformation is apparent in the identified ‘transition’ regions, which are

disregarded, as the focus of the fracture surface analysis was to identify only ‘tensile’ and

‘compressive’ regions. Investigation of these ‘transition’ regions would further supplement

the fracture surface analysis technique.

127

5.4.3. Failure mechanisms

Analysis of the 14-day KOH treated fracture surfaces at higher magnifications revealed

additional toughening mechanisms including crack deflection, microcracking and bridging

(as defined by uncracked ligaments and unbroken individual collagen fibers). Interlamellar

delamination and a layered morphology are commonly associated with microcracks whereas

protruding fiber bundles are usually associated with crack bridging by uncracked ligaments

and/or collagen fibers [6]. All of these toughening mechanisms have been proposed for bone

[7]. A prevalence of these features in the 14-day KOH treated fracture surfaces when

compared to untreated bones may explain the reason for the increased toughness observed for

these groups in Chapter 3 – Initial Study.

During failure in bending, collagen fibrils may separate from larger bundles and act as

‘bridges’ between adjacent lamellae. Similarly, crack bridging by uncracked ligaments along

the crack path bridge the opening crack, blunting further propagation [15,16]. As the ‘bridge’

develops, resistance to crack propagation is enhanced until a steady-state bridging zone is

achieved [25].

A discontinuous crack path is present with unbroken regions (uncracked ligaments) in both

female and male 14-day KOH treated bones. Bridges span the crack wake, limiting crack

opening and sustaining the applied load, which would otherwise be used for further crack

propagation. Crack bridging acts to resist crack opening, requiring additional energy for

fracture [15,16]. Support for this theory is provided by the high magnification images inside

cracks of the 14-day KOH treated samples, showing collagen fibrils spanning a microcrack.

Nalla et al. observed similar fibers in microcracks of human cortical bone [7]. Moreover,

fracture surface analysis of abalone shell, a biocomposite composed of calcium carbonate

and insoluble proteins, showed filaments spanning the crack. These filaments were shown to

improve the mechanical strength of the shell [11,26]. Furthermore, 14-day KOH treated male

fracture surfaces revealed interwoven mineralized collagen fibers between two adjacent

lamellae. This interconnected structure indicates resistance to a propagating crack and a

resultant expected increase in energy absorption upon failure. As a result, a ‘rougher’ fracture

surface is created and this fiber bridging may partially explain the increased toughness

128

observed in the 14-day KOH treated samples. For the untreated groups, microcracks show no

filaments spanning cracks. The crack path of untreated bones was relatively straight and

showed no evidence of bridging.

Microcracks have also been suggested as a possible toughening mechanism in bone by

shielding the crack and are constrained by the surrounding undamaged material [27].

However, the extent of microcracking was minimal in the emu tibiae of this study. If

microcrack effects impacted the samples toughness, it is assumed that this effect is secondary

to the easily identified bridging effect as a primary toughening mechanism. Another potential

toughening mechanism is the presence of cement lines and interlamellar interfaces that

deflect crack propagation [7]. In fact, microcracks have been observed to deflect along

cement lines when encountering osteons [14], suggesting that cement lines may provide a

weak path for fracture redirection. Crack deflection increases toughness by increasing the

length of the crack propagation path upon deflection, which implies more energy absorption

during its travel along the longer crack path. Subsequent crack growth is achieved when the

crack finds a more favourable, meaning less energetic, path. Microstructural features, such as

osteons and pores, play a crack arresting or deflecting role [28]. Fracture surfaces from

untreated samples appear to contain fewer features of resistance able to deflect a crack and

produce fractographic irregularities [2]. Therefore, untreated fracture surfaces have a

‘smoother’ surface compared to the ‘rougher’ KOH treated fracture surfaces.

With regards to the cement line as a weak interface, Burr et al. showed that cement lines are

weak interfaces that are more ductile than the surrounding bone matrix, as they allow crack

initiation, but provide resistance to crack growth [29]. Consequently, a propagating crack

will be decelerated by a compliant ductile interface, whereas it will be arrested by a stiffer

interface [29,30]. KOH treatment may have infused KOH between the cement line and the

surrounding tissue, altering the cement line-tissue interface. If this alteration weakened the

interface, it may result in a more compliant interface. This compliant cement line-tissue

interface could increase the ductility of the affected bone.

129

The effect of weak interfaces has long been recognized as a crack stopping mechanism in

engineering composite materials [4]. A weak interface between phases may cause separation

or delamination before the propagating crack is able to advance to the next brittle layer [31].

Conversely, a strong interface leads to high stress concentrations that cause fracture,

decreasing the toughness of the material [32]. As such, to improve toughness, interfacial

debonding and ‘frictional sliding’ are desirable as they relieve these stress concentrations.

However, extensive debonding and low interface friction decrease composite strength due to

poor load transfer. The goal is to achieve a balance between load transfer and pull-out

capabilities while still allowing for an interface to promote interfacial debonding and reduce

fiber strain [32]. In fact, it has been shown that bone with decreased bonding between the

mineral and organic phase would become more ductile [33,34]. KOH treatment may have

altered the interface between the mineral and organic matrix. Infusion of the basic solution

may have permanently affected the organic components of bone, possibly debonding the

links between the mineral and matrix through either affecting collagen and/or other protein

conformation and/or the electrostatic interaction between the mineral surface and surface

proteins. Partial debonding between the mineral and organic matrix may be an additional

mechanism that is responsible for the significant increase in toughness due to KOH

treatment.

5.5. Conclusions

The increased toughness observed in 14-day KOH treated bones compared to untreated bones

is the result of several toughening mechanisms acting together: crack deflection,

microcracking and organic matrix bridging. The increased prevalence of these mechanisms,

identified by the increase in percent area of fracture surface roughness in the 14-day KOH

treated samples are, in turn, direct results of the unique KOH treated bone microstructure.

The information gained from fracture surface analysis using SEM compliments the

mechanical testing results from Chapter 3 – Initial Study, which suggests that fracture surface

analysis is a useful technique for understanding bone fracture mechanisms.

130

5.6. Chapter Summary

The fracture surfaces of 14-day female and male KOH treated bones showed a significantly

higher ‘roughness’ compared to untreated bones, indicating that more energy was consumed

in the 14-day KOH treated fractures. Furthermore, additional toughening mechanisms, which

are important features for dissipating energy during the failure process, were observed in the

14-day KOH treated samples, but were absent in the untreated samples for both sexes. This

suggests that these mechanisms slowed the propagation of the catastrophic crack. These

results indicate that the significant increase in toughness of 14-day KOH treated bones is due

to structural alterations that enhance the ability of the bone microstructure to dissipate energy

during the failure process, thereby slowing crack propagation. Fracture surface analysis has

helped explain why 14-day KOH treated bones have increased toughness compared to

untreated bones, namely via toughening mechanisms on the compressive failure side.

However, KOH treatment may have also altered the links between the mineral and collagen,

resulting in partial debonding, which has been shown to result in a weaker bone with

increased toughness [35].

131

5.7. References

[1] Braidotti P, Branca FP, Stagni L. (1997) Scanning electron microscopy of human cortical bone failure surfaces. J.Biomech. 30:155-162.

[2] Corondan G, Haworth WL. (1986) A fractographic study of human long bone. J.Biomech.

19:207-218. [3] Bonfield W, Li CH. (1966) Deformation and fracture of bone. J.Appl.Phys. 37:869-875. [4] Piekarski K. (1970) Fracture of bone. J.Appl.Phys. 41:215-223. [5] Braidotti P, Bemporad E, D’Alessio T, Sciuto SA, Stagni L. (2000) Tensile experiments and

SEM fractography on bovine subchondral bone. J.Biomech. 33:1153-1157. [6] Sahar ND, Hong S-I, Kohn DH. (2005) Micro- and nano-structural analyses of damage of

bone. Micron. 36:617-629. [7] Nalla RK, Kinney JH, Ritchie RO. (2003) Mechanistic fracture criteria for the failure of

human cortical bone. Nat.Mater. 2:164-168. [8] Saha S, Hayes WC. (1977) Relations between tensile impact properties and microstructure of

compact bone. Calcif.Tissue.Res. 24:5-72. [9] Wise LM, Wang Z, Grynpas MD. (2007) The use of fractography to supplement analysis of

bone mechanical properties in different strains of mice. Bone. 41:620-630. [10] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

[11] Fantner GE, Birkedal H, Kindt JH, Hassenkam T, Weaver JC, Cutroni JA, Bosma BL,

Bawazer L, inch MM, Cidade GA, Morse DE, Stucky GD, Hansma PK. (2004) Influence of the degradation of the organic matrix on the microscopic fracture behavior of trabecular bone. Bone. 35:1013-1022.

[12] George WT, Vashishth D. (2005) Damage mechanisms and failure modes of cortical bone

under components of physiological loading. J.Orthop.Res. 23:1047-1053. [13] Yeni YN, Norman TL. (2000) Fracture toughness of human femoral neck effect of

microstructure, composition and age. Bone. 26:499-504. [14] Ritchie RO, Kinney JH, Druzic JJ, Nalla RK. (2005) A fracture mechanics and mechanistic

approach to the failure of cortical bone. Fat.Fract.Engng.Mater.Struct. 28:345-371. [15] Vashishth D, Behiri JC, Bonfield W. (1997) Crack growth resistance in cortical bone:

concept of microcrack toughening. J.Biomech. 30:763-769. [16] Vashishth D, Tanner KE, Bonfield W. (2000) Contribution, development and morphology of

microcracking in cortical bone during crack propagation. J.Biomech. 33:1169-1174.

132

[17] Yeni YN, Norman TL. (2000) Calculation of porosity and osteonal cement line effects on the effective fracture toughness of cortical bone in longitudinal crack growth. J.Biomed.Mater.Res. 51:504-509.

[18] Wynnyckyj C, Omelon S, Savage K, Damani M, Chachra D, Grynpas MD. (2009) A new

tool to assess the mechanical properties of bone due to collagen degradation. Bone. 44:840-848.

[19] Turner CH, Burr DB. (1993) Basic biomechanical measurements of bone: a tutorial. Bone.

14:595-608. [20] von Knoch M, Bluhm A, Morlock M, von Foerster G. (2003) Absence of surface roughness

changes after insertion of one type of matte cemented femoral component during 2 to 15 years. J.Arthroplasty. 18:471-7.

[21] Szmukler-Moncler S, Testori T, Bernard JP. (2004) Etched implants: a comparative surface

analysis of four implant systems. J.Biomed.Mater.Res.B.Appl.Biomater. 69:46-57. [22] Callister WD. (2003) Materials Science and Engineering: An Introduction 7th ed. John Wiley

& Sons, New York. [23] Currey JD, Brear K. (1992) Fractal analysis of compact bone and antler fracture surfaces.

Biomimetics. 1:103-118. [24] Zimmermann EA, Launey ME, Barth HD, Ritchie RO. (2009) Mixed-mode fracture of

human cortical bone. Biomater. 30:5877-5884. [25] Kruzic JJ, Scott JA, Nalla RK, Ritchie RO. (2006) Propagation of surface fatigue cracks in

human cortical bone. J.Biomech. 39:968-972. [26] Smith BL, Schaffler TE, Tilman E, Viani M, Thompson JB, Frederick NA, Kindt J, Belcher

A, Stucky GD, Morse GE, Hansma PK. (1999) Molecular mechanistic origin of the toughness of natural adhesives, fibres and composites. Nature. 399:761-763.

[27] Kruzic JJ, Nalla RK, Kinney JH, Ritchie RO. (2003) Crack blunting, crack bridging and

resistance-curve fracture mechanics in dentin: effect of hydration. Biomater. 24:5209-5221. [28] Hazenberg JG, Taylor D, Lee TC. (2006) Mechanisms of short crack growth at constant

stress in bone. Biomater. 27:2114-2122. [29] Burr DB, Schaffler MB, Frederickson RG. (1988) Composition of the cement line and its

possible mechanical role as a local interface in human compact bone. J.Biomech. 21:939-945. [30] Kendall K. (1975) Control of cracks by interfaces in composites. Proc.R.Soc.Lond. 341:409-

428. [31] Cooke F, Zeidman H, Scheifele SJ. (1973) The fracture mechanics of bone – another look at

composite modeling. J.Biomed.Mater.Res.Symp. 4:383-399. [32] Beyerlein IJ, Zhu YT, Mahesh S. (2001) On the influence of fiber shape in bone-shaped

short-fiber composites. Compos.Sci.Technol. 61:1341-1357.

133

[33] Bundy KJ. (1985) Determination of mineral-organic bonding effectiveness in bone--theoretical considerations. Ann.Biomed.Eng. 13:119-35.

[34] Kotha SP, Guzelsu N. (2000) The effects of interphase and bonding on the elastic modulus of

bone: changes with age-related osteoporosis. Med.Eng.Phys. 22:575-85. [35] Walsh WR, Ohno M, Guzelsu N. (1994) Bone composite behaviour: effects of mineral-

organic bonding. J.Mater.Sci.Mater.Med. 5:72-79.

134

Chapter 6 Collagen Degradation

6.1. Introduction

6.2. Experimental Details

6.2.1. Reagents 6.2.2. Emu bone samples and KOH treatment

6.2.3. Powder X-ray diffraction 6.2.4. Quantitative backscattered electron imaging 6.2.5. Microhardness testing 6.2.6. Bone powder preparation 6.2.7. α-Chymotrypsin 6.2.8. DSC 6.2.9. SDS-PAGE 6.2.10. Polarized light microscopy 6.2.11. Statistical analysis

6.3. Results

6.3.1. Powder X-ray diffraction 6.3.2. Microhardness 6.3.3. Quantitative backscattered electron imaging 6.3.4. α-Chymotrypsin 6.3.5. DSC 6.3.6. SDS-PAGE 6.3.7. Polarized light microscopy

6.4. Discussion

6.4.1. Mineral characterization 6.4.2. Collagen degradation 6.4.3. Partial debonding of the collagen-mineral interface 6.4.3. Sex differences

6.5. Conclusions

6.6. Chapter summary

6.7. References

135

The following manuscript is currently in revision in Biochimica et Biophysica Acta, General

Subjects under, “Mechanism of bone collagen degradation due to KOH treatment”. This

chapter is a continuation of the work done in Chapter 3. In this chapter, we characterized the

mineral and collagen phases of untreated (0-day) and KOH treated emu tibiae in order to gain

insights into the KOH mechanisms responsible for the altered bone mechanical properties

induced by endocortical KOH treatment observed in Chapter 3.

6.1. Introduction

The mechanical properties of bone are dependent on the quality, specific arrangement and

interaction of an organic matrix and mineral apatite crystals (hydroxyapatite) that form a

carefully designed composite material. Bone consists of approximately 65% mineral and

30% organic matrix, including proteins and cells with the remainder solutes [1]. The mineral

component of bone is composed of small, poorly crystalline and highly substituted apatite

crystals [1,2]. The mineral phase largely contributes to the overall strength and stiffness of

bone [2,3]. The organic matrix consists mainly (~90%) of Type I collagen, a triple helical

molecule that is specifically arranged in several hierarchical levels to provide elasticity and

toughness to bone [2,4,5]. The remainder of the organic matrix components consists of

various noncollagenous proteins that have functions that include initiation and inhibition of

mineral deposition. Some of the most abundant noncollagenous proteins are osteonectin,

osteocalcin, osteopontin and bone sialoprotein. While their individual roles are still not fully

defined, it is well understood that some of these proteins play integral roles in the initiation

and regulation of both mineral deposition and maturation [1]. Finally, it has been reported

that the interaction between the mineral and the collagen matrix play an important role in the

mechanical properties of bone [6,7].

Type I collagen is a heterotrimer composed of two α1(I) chains and one α2(I) chain in a triple

helical structure [8]. The triple helical collagen molecule, which is stabilized by hydrogen

bonds, aggregates with other collagen molecules to form a staggered array that is stabilized

by intermolecular crosslinks. These microfibrils aggregate into collagen fibers to form a

highly hierarchical collagen network structure [9-12]. It has been shown that changes in

collagen molecules, their lattice structure (packing) and crosslinks may affect the mechanical

136

integrity of the collagen network and subsequently lead to altered bone strength and

toughness [12-15].

Various analytical techniques have been developed to detect differences in the collagen

network of bone. Several studies have utilized a selective digestion technique that is

generally thought to distinguish between denatured and intact collagen molecules.

Identification and detection of these different collagen molecules allows for the direct

measurement of the percentage of degraded collagen [6,11,12,16-19]. The thermostability of

the collagen network has been measured frequently using calorimetry and correlated to

collagen network quality [20-27]. Direct mechanical testing has also been used to assess the

properties of the collagen network [11,12,15,17,28,29].

To study the contribution of collagen to the mechanical properties of bone, we developed a

model using the emu tibiae. In this model, the infusion of a basic solution was expected to

affect the collagen component of bone, while leaving the mineral content of bone unaltered,

as apatite is stable in basic solutions [7]. The role of collagen on bone mechanical properties

is gaining more prominence as the changes brought on by osteoporosis and other disorders

are now known to affect the organic matrix of bone and cannot be predicted with the use of

bone mineral density (BMD) measurements [30]. While the KOH treatment is not

physiological, its role is to improve our understanding of the mechanisms by which collagen

degradation, not removal, affects bone mechanical properties.

Female and male emu tibiae were endocortically treated with 1 M potassium hydroxide

(KOH) solution for 1-14 days resulting in negligible mass loss (0.5%), collagen loss (0.05%),

no differences in geometrical parameters, but exhibiting significant changes in mechanical

properties. Female and male emu tibiae showed significant decreases in failure stress and

increased failure strain and toughness with increasing KOH treatment time. These changes

were more significant in male bones compared to female bones [7]. KOH was chosen as a

degradation agent as it has been used previously to affect bone tissue [31,32] and the use of

KOH does not require a demineralization step. Consequently, this degradation method retains

the mineral phase and the geometry of the bone. While the mechanism of action of KOH is

137

not well understood, it is likely to be similar to sodium hypochlorite. Sodium hypochlorite

extraction leaves the mineral phase largely unaffected [31], similar to results seen in our

initial study with KOH treatment [33]. The high pH of 14 of the 1 M KOH solution is likely

to result in conformational changes to the collagen and may also affect mineral-organic

interactions [34].

The goal of the present study was to gain insights into the KOH mechanisms that alter the

collagen properties and/or the interface between the mineral and collagen. The hypothesis of

the present study is that in situ collagen degradation rather than collagen loss may be

responsible for the significant changes in bone mechanical properties induced by

endocortical KOH treatment. To test this hypothesis, bone samples were analyzed using

powder x-ray diffraction, microhardness testing, quantitative backscattered electron imaging,

an α-chymotrypsin selective collagen digestion technique, differential scanning calorimetry,

SDS-PAGE and polarized light microscopy. Correlations of collagen degradation results with

previously reported bone mechanical properties (from Chapter 3 – Initial Study) were also

explored.

6.2. Experimental Details

6.2.1. Reagents

Ethylenediaminetetraacetic acid (EDTA), benzamidine, N-ethylmaleimide, ε-amino-n-

caproic acid, phenylmethylsulfonyl fluoride, Tris, iodoacetamide, pepstatin A, chloramine-T,

Ehrlich’s reagent, pepsin, urea, sodium phosphate, sodium dodecyl sulfate (SDS) and β-

mercaptoethanol were purchased from Sigma Aldrich (Milwaukee, WI). Chloroform,

methanol, perchloric acid, potassium hydroxide, hydrochloric acid, sodium hydroxide and

glacial acetic acid were obtained from Fisher Scientific (Pittsburgh, PA). α-Chymotrypsin

(TLCK-treated) was purchased from MP Biomedical (Solon, OH).

6.2.2. Emu bone samples and KOH treatment

The emu bone tibiae samples from Chapter 3 – Initial Study, were used in this study. These

bones were endocortically treated with 1 M potassium hydroxide (KOH), followed by

mechanical testing in three-point bending [7]. Bones were divided into two groups of female

138

and male tibiae with ten samples allocated to each of 0, 1, 3, 7 or 14-day KOH treatments.

An additional ten female and ten male bones were filled with saline instead of KOH for 14

days to act as controls. Computed Tomography (CT) scans were taken at the mid-point of

each emu tibia with an Aquilion 64 CT scanner (Toshiba, Canada) before and after the

different KOH treatment time points to normalize load-displacement data and to eliminate

differences caused by variation in geometry.

Briefly, skeletally mature (3-5 years of age) whole emu tibiae were separated from the

femora and tarsometatarsi with a scalpel. The shaft of each tibia was then isolated by using a

circular saw to remove the ends (15% of the total bone length from the proximal end and

10% from the distal end). The marrow and trabecular bone from the diaphysis of the tibiae

were removed by drilling longitudinally through the bone shaft, after which the medullary

canal was flushed with tap water [7]. The emu tibiae were filled with 1 M KOH and the ends

were sealed with polymethylmethacrylate (PMMA, SR Ivolen Liquid and Powder, Ivoclar

Vivadent, Mississauga, ON). The bones were held horizontally over a collection basin and

kept moist with a 0.9% saline solution drip. The bones were rotated 180 ° around the axial

axis every 12 hours during the KOH treatment. After the desired treatment period, the KOH

solution was reclaimed, its volume measured and the endocortical surfaces of the bones

rinsed in running tap water for one hour. Emu tibiae were then tested in three-point bending

using a servo-hydraulic materials testing machine (Model 8511, Instron, Canton, MA). The

bone was secured in the three-point bending jig with a 100 N pre-load and then loaded to

failure at a displacement rate of 0.04 mm/s. Force and displacement data were collected and

normalized using geometric data from CT scans to create stress-strain curves to determine

the elastic modulus (E; GPa), ultimate stress (σuts; MPa), failure stress (σf; MPa), failure

strain (εf, %), and toughness (U; mJ/mm3).

6.2.3. Powder X-ray diffraction

To evaluate crystal size and changes in overall crystallinity, powder X-ray diffraction (XRD)

was performed on ground emu tibiae. In powder XRD, x-rays scan the surface of bone

powder over specified incident angles and are reflected. Sharp peaks at specific incident

angles are generated for crystalline materials, which are determined by the regular spacing

139

between the atomic planes in the relevant crystal structure. In poorly crystalline materials and

materials with crystals of very small dimensions (<100 Å), the peaks are broadened.

Broadening of the peaks in bone XRD scans is largely due to the small crystal size of bone

apatite and inherent lattice strain effects. The measurement of this peak broadening can then

be related to the average crystal size/strain [35].

For each bone, ten mm thick cross-sections from each group were cut, manually crushed,

Tris-washed, lyophilized, defatted overnight in a 2:1 chloroform:methanol solution and dried.

Dry specimens were ground to a powder (<45 μm particle size) using a cryogenic freezer

mill (SPEX Certiprep 6750 Freezer Mill, Metuchen, NJ). Each powdered sample was spread

into a thin even layer on a quartz crystal sample holder for powder X-ray diffraction scans

(Rigaku MultiFlex, Rigaku/MSC, Woodland, TX). Using CuKα radiation at 40 kV and 40

mA, samples were scanned from 24.5 to 27.0 degrees (2θ) at a scan speed of 0.1 degrees per

minute for crystal length (002) determination. Additionally, samples were scanned from 37.0

to 42.0 degrees (2θ) at a scan speed of 0.05 degrees per minute for crystal cross-section (310)

analysis. A NIST reference material (2910, Calcium Hydroxyapatite) was used as a standard

for every run.

Crystallite size was calculated from the peak broadening of the powder XRD peaks [36].

This broadening can be estimated by measuring the full width at half the maximum height

(FWHMH) of each of the apatite peaks (β1/2 (002) and β 1/2 (130)), using the JadeTM XRD

pattern processing software (Materials Data Inc., Win-OS V5). Since peak broadening (β)

can also be affected by instrument broadening, the sample peak widths were appropriately

adjusted using values obtained by scanning reference silicon at 26 and 55 °2θ. The

broadening due to the specimen only (corrected β1/2) was calculated from the square root of

the instrument broadening squared subtracted from the measured specimen peak broadening

squared. This corrected β1/2 was used to calculate the crystalline size between the (002) and

(310) planes using the Debye-Scherrer equation:

θβλKD

cos3.57

2/1=

140

where 57.3 is a conversion factor from degrees to radians, K is a correction factor (0.9) used

to reflect the elongated apatite crystals of bone, λ is the K-emission wavelength of copper, θ

is the diffraction angle and D is the size of the crystallite in angstroms along the specific axis.

The means of three independent FWHMH values for each 26 and 40 °2θ peak were

calculated for each sample.

6.2.4. Quantitative backscattered electron imaging

Quantitative backscattered electron (qBSE) imaging allows the determination of the

mineralization profile of a bone section using a scanning electron microscope. A scanning

electron microscope works through a focused beam that is continuously scanned over the

surface of a sample causing backscatter of the primary electrons and ejection of secondary

electrons. Areas containing higher atomic numbers have an increased probability of collision

with these electrons. Heavy elements (high atomic number) backscatter electrons more

strongly compared to light elements (low atomic number) and thus appear brighter in the

image. These backscattered electrons are used to detect contrast between areas with different

chemical compositions. Therefore, an area of bone with a greater density of calcium atomic

nuclei will display a greater intensity of electron collisions [39].

For each bone, ten mm thick sections of bone from each group were cut and subsequently

divided into four sections: anterior, posterior, medial and lateral. Samples were dehydrated in

ascending concentrations of acetone and subsequently infiltrated in ascending ratios of

unpolymerized Spurr resin and acetone. The bones were then embedded in blocks of Spurr

resin that were polymerized in a 60 °C oven for 48 hours. These blocks were ground,

polished, carbon-coated and imaged using quantitative backscattered electron (BSE) imaging

for evaluation of mineralization distribution (solid state BSE detector, FEI Company,

Hillsboro, OR) on a Phillips XL30 ESEM (FEI). Beam conditions were set at 20 kV, working

distance of 15 mm and a spot size of 7. The relative backscattering of the samples was

determined by comparison with a silicon dioxide standard, which was measured between

every specimen measurement. This was required to correct for the variation (drift of the

machine) that could potentially occur between specimens.

141

Histograms of the grey level distribution were created for each of the anterior, posterior,

medial and lateral sections of the cortical samples, with increasing brightness representing

increasing mineralization [40]. A Quips program on a Quantimet 500 IW system sets up a

series of bins based on the intensity of the pixels and calculates the number of pixels falling

in each bin. This information is used to create a histogram of each image. From the

histograms, the grey level of the histogram peak was determined and used to represent the

overall degree of mineralization. The full width at half the maximum height (FWHMH) of

each histogram was also noted, which represents the heterogeneity of the distribution [40].

Images obtained from quantitative BSE were next analyzed for the number of large and small

pores as well as the total width of the dark band from the endocortical region towards the

periosteal region for the medial, lateral, anterior and posterior zones of the bones using a

Quips program on a Quantimet 500 IW system. A 47 μm diameter threshold was used for the

separation between small and large pores. This threshold value was determined after

analyzing a histogram of the overall pore area, which showed two maximum points (bi-

modal distribution). The lowest point of the ‘valley’ between the two maxima was defined as

the threshold value.

6.2.5. Microhardness testing

Hardness refers to the resistance of a material to indentation. Strong correlations have been

found between microhardness and mechanical properties of bone [37,38].

Following quantitative BSE imaging, the same Spurr embedded blocks were subsequently

used for microhardness testing on a microhardness testing machine (Mitutoyo HM-122, S/N

260113). During the test, a pyramidal diamond indenter of known geometry was lowered

onto the sample under a known load for 10 s, leaving an indentation on its surface. The

resulting length of the indentation and applied force are related to the microhardness of the

bone using the following equation:

2 2 1891.02sin2

102.0102.0D F

D

θF

SF

SF k HV ====

142

where HV is the Vickers Hardness, k is a constant (0.102), F is the test force (0.025 kg), S is

the surface area of indentation (mm2), D is the average length of two diagonals (mm) and θ is

the face-to-face apex angle of the diamond indenter (136°). Indentations were taken from the

endocortical region to the periosteal region (250 μm from the bone edge; 125 μm between

each measurement). Five measurements were taken at the medial, lateral, anterior and

posterior zones of the bone in order to obtain the hardness profile of each sample.

6.2.6. Bone powder preparation

After mechanical testing, ten mm thick cross-sections of bone from each group were cut into

smaller pieces using a table saw, washed by vortexing three times in 0.2 M Tris-HCl (pH 7.4)

containing the following protease inhibitors (PI): benzamidine (5 mM); N-ethylmaleimide (5

mM); ε-amino-n-caproic acid (10 mM) and phenylmethylsulfonyl fluoride (PMSF) (1 M)

and then dried overnight on a lyophilizer. Samples were then defatted in a 2:1

chloroform:methanol mixture and left to agitate for 24 hours at room temperature. Defatted

samples were placed in 100% methanol for one hour and left to dry overnight. Dried samples

were ground to a fine powder using a cryogenic freezer mill (SPEX certiprep 6750 Freezer

Mill, Metuchen, NJ) [41-43]. Next, the powder was demineralized in buffered (pH 7.4) 0.5 M

ethylenediaminetetraacetic acid (EDTA), 4 M guanidine hydrochloride (GuHCl) with PI at 4

°C for three weeks with daily exchanges and then thoroughly washed three times in distilled

water [19,41-43]. The extracts of collagen-rich powder were lyophilized in preparation for α-

chymotrypsin digestion, DSC and SDS-PAGE.

6.2.7. α-Chymotrypsin

The amount of collagen degraded by KOH treatment was determined by the assay described

by Bank et al. [16], which is based on the observation that α-chymotrypsin digests

denatured/degraded collagen but not intact triple helical polymeric collagen [44]. Ten mg of

demineralized bone powder were digested at 37 °C for 24 hours in one ml of incubation

buffer (0.1 M Tris HCl pH 7.3, 1 mM iodoacetamide, 1 mM EDTA and 10 μg/ml pepstatin-

A), containing one mg/ml α-chymotrypsin. Since α-chymotrypsin selectively digests

degraded collagen, the solubilized (degraded) collagen was separated from intact polymeric

143

collagen by removing the supernatant (containing the degraded collagen molecules) from the

remaining insoluble matrix (containing the intact polymeric collagen molecules) [6].

The amounts of collagen in the supernatant and the pellet were determined using the

colourimetric hydroxyproline assay according to the method of Woessner [45], assuming that

collagen is 10% hydroxyproline [45,46]. Specimens were hydrolyzed in one ml of 6 N HCl

for 18 hours at 110 °C. Hydrolyzates were then neutralized with 2.4 ml of 2.5 N NaOH per

ml of HCl and diluted with distilled water. The samples were prepared for colourimetric

analysis with the addition of 0.05 N chloramine-T, 3.15 N perchloric acid and Ehrlich’s

Reagent. The colourimetric reaction was quantified with a μ-Quant Microplate

Spectrophotometer (BioTek Instruments, Winooski, VT) at 560 nm. Absorbance values were

plotted against the concentration of standard hydroxyproline (0-5 μg) and the quantity of

hydroxyproline determined from the standard curve [45]. Finally, the percentage of digested

collagen was calculated by dividing the amount of hydroxyproline present in the supernatant

by the sum of amount of hydroxyproline present in the supernatant and remaining matrix

[16]. The α-chymotrypsin digestion assay was performed in triplicates for each specimen.

6.2.8. DSC

Differential scanning calorimetry (DSC) was utilized in order to detect collagen denaturation

and/or changes in intermolecular structure (collagen degradation) [21,22,27]. Fifty μl of PBS

was added to ten mg of bone powder to maintain the collagen in a fully hydrated condition

[47]. The hydrated demineralized bone powder was then equally divided into three volumes

in order to perform triplicate measurements of each sample. Specimens were loaded in

hermetically sealed, aluminum DSC pans (TA Instruments Inc., New Castle, DE) and

scanned at 5 °C per minute from 25 °C to 85 °C [21,47,48]. An empty pan was used as a

reference. Samples were run on a TA Instruments DSC Q2000 calorimeter with refrigerated

cooling system (TA Instruments Inc., New Castle, DE). Temperature and heat flow scales

were calibrated with Indium standards.

After DSC analysis, the collagen content of each pan was determined using the

hydroxyproline assay, as described above [45], to normalize thermal properties to collagen

144

content. From the resulting thermogram of heat flow versus temperature (Figure 6.1), the

following thermodynamic parameters were determined: denaturation temperatures (Tonset and

Tpeak), enthalpy of denaturation (ΔH), Height and full width at half maximum height

(FWHMH) of the endothermic peak. Specifically, Tonset (°C) and Tpeak (°C) provide measures

of the thermal stability. Enthalpy of denaturation, ΔH (J/g) is the area under the thermogram

normalized to collagen mass and is the amount of heat required for denaturation. Height

(mW/mg) is the height of the thermogram peak normalized to collagen mass and FWHMH

(°C) is the full width at half maximum height of the endothermic peak. Together, the height

of the thermogram peak and FWHMH values provide a measure of the heterogeneity of the

molecular structure (shape of the curve). All parameters were determined using the DSC

system software (TA Instruments, Universal Analysis 2000).

Figure 6.1: Schematic of a typical DSC curve and the definitions of the DSC parameters measured. Tonset (°C) is the intersection of a line that is tangent to the steepest section of the leading edge and the baseline of the thermogram. Tpeak (°C) is the denaturation temperature at maximum heat flow. Enthalpy of denaturation, ΔH (J/g), is the area under the thermogram peak normalized to collagen weight. Height (mW/mg) is the height of the thermogram peak normalized to collagen weight and FWHMH (°C) is the full width at the half maximum of the thermogram.

Temperature (degrees C)

FWHMH

ΔH

Height

Tonset Tpeak

area =

Temperature (degrees C)

FWHMH

ΔH

Height

Tonset Tpeak

area =

145

6.2.9. SDS-PAGE

Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) was used to

determine the amount of collagen fragmentation due to KOH treatment. Using an established

protocol [42,49], demineralized bone powder was solubilized in 0.5 M acetic acid solution

with pepsin in a 10:1 weight ratio of bone collagen to enzyme for 72 hours at 4 °C [50,51].

The reaction was neutralized and the samples centrifuged for 30 minutes at 14,000 rpm to

remove insoluble collagen. The concentration of soluble collagen in the supernatant was

determined by measuring the hydroxyproline content, as described above.

A one ml aliquot of solubilized collagen and fragments from each sample was lyophilized

and resuspended in a buffer containing 0.01 M sodium phosphate pH 7.2, 2 M urea and 0.1%

sodium dodecyl sulfate (SDS) [52-54]. Samples loaded onto the gel contained a 1:1 mixture

of protein (20 μg) to Laemmli buffer (BioRad) with 10% β-mercaptoethanol. The samples

were loaded in duplicates and then electrophoresed for 2 hours at 140 V on a 4-20% gradient

polyacrylamide gel (BioRad). A BioRad Precision Plus Protein Dual Colour Standard (5 μL)

and acid-soluble rat tail tendon collagen Type I (5 μg, Sigma) were loaded to act as a protein

standard and a pure collagen control, respectively. The gels were stained with Coomassie

Blue staining solution (BioRad) for one hour and then washed overnight in a methanol:acetic

acid solution until destained. Protein bands were visualized using densitometry and the

images were analyzed using ImageJ (ImageJ 1.28u, National Institutes of Health). Only

intact α-chain bands (α1(I) and α2(I)) were quantified by defining each lane with the

rectangular tool, which generated lane profile plots. Lines were drawn to enclose the peaks of

interest (α1(I) and α2(I)) and to determine the average collagen peak intensity (area under

peak).

6.2.10. Polarized light microscopy

Polarized light microscopy was used to analyze the arrangement of the collagen fibril

network in bone sections [55]. Ten mm thick cross sections were cut from untreated (0-day)

and 14-day KOH treated female and male emu tibiae. From these samples, transcortical

sections were cut (10 mm long, 5 mm circumferentially). These samples were then

demineralized in buffered (pH 7.4) 0.5 M EDTA with protease inhibitors (PI) at 4 °C for six

146

weeks with daily exchanges and then thoroughly washed three times in distilled water

[19,41-43]. After dehydration and paraffin embedding, five μm sections were cut and stained

with 0.1% picro-sirius red F3B (BDH) [56]. Picro-sirius red is used to identify collagen

fibres, by the reaction between its sulphonic acid groups and the basic groups found in

collagen fibres [55]. Picro-sirius red stains collagen red in a bright-field microscope, whereas

under polarization microscope collagen appears bright orange/red and/or bright green [55].

The sections were examined at 150X magnification at the endocortical and periosteal regions

on a Leica microscope with the use of polarized light and filter analyzer. Images were

acquired using a Leica CCD camera.

6.2.11. Statistical analysis

Statistical analysis was performed using SPSS (SPSS 16.0 for Windows, Chicago, IL)

statistical analysis software. Tests for normality and equality of variances were initially

performed to determine whether parametric or non-parametric t-tests should be used. Two-

way analysis of variance (ANOVA, general linear model) was performed to examine the

effects of sex and KOH treatment on all measured parameters in the emu model and multiple

comparisons (Fisher’s Least Significant Difference (LSD)) were made to detect significant

differences between the groups. In addition, multiple regression analyses were used to

explore the correlations of bone mechanical properties (elastic modulus, failure stress, failure

strain and toughness) from Chapter 3 – Initial Study with the percent degraded collagen from

each technique. Differences between percent degraded collagen with bending mechanical

parameters were investigated using a generalized linear model to determine any differences

between sexes. All data are presented as mean ± standard error of the mean. A confidence

level of 95% (p=0.05) was considered statistically significant for differences and

correlations.

6.3. Results

Significant differences were found in the biochemical parameters measured between the

KOH treated groups and the untreated groups, as well as between sexes. All parameters

measured for the control group (14-day filled with saline) were similar to those of the

untreated (0-day) groups (Appendix A) for both female and male emu tibiae. The data for all

147

treated groups were combined for the regression analysis of the parameters for female and

male data.

6.3.1. Powder X-ray diffraction

The KOH treatment did not affect the crystal size (length and cross section) as there were no

differences between crystal size at the periosteum (unaffected by KOH) and the endocortex

(region affected by KOH) (Table 6.1). There were no differences observed between sexes.

6.3.2. Microhardness

Microhardness testing did not show differences between locations (medial, lateral, anterior

and posterior zones), nor with increasing distance from the endocortical region. The data was

thus grouped and no differences were observed between sexes at any KOH treatment time

point (Table 6.1).

Table 6.1: Emu bone mineral crystal length (002) and cross section (310) estimated by XRD and microhardness testing results.

KOH treatment time in days Parameter

Sex

0 1 3 7 14

Female 205 ± 2 203 ± 3 204 ± 2 201 ± 3 206 ± 2 Crystal length 26° peak (Å) Male 209 ± 2 204 ± 3 207 ± 4 203 ± 2 211 ± 2

Female 62.1 ± 0.5 62.9 ± 0.7 62.8 ± 0.6 63.3 ± 0.6 62.8 ± 0.6 Crystal cross section 40° peak (Å) Male 62.9 ± 0.6 62.8 ± 0.8 63.4 ± 0.9 63.5 ± 0.6 63.9 ± 0.6

Female 61 ± 0.4 62 ± 0.8 61 ± 0.4 61 ± 0.6 63 ± 0.5 Microhardness, HV (kg/mm2) Male 62 ± 0.5 61 ± 0.8 61 ± 1.0 62 ± 0.9 63 ± 0.7

6.3.3. Quantitative backscattered electron imaging

The average peak grey level and FWHMH, which measures the heterogeneity of the

mineralization distribution, did not show any differences between sexes as well as with KOH

treatment time (Table 6.2). Porosity measurements revealed no differences in large pore

density (pore diameter > 47 μm) and small pore density (pore diameter < 47 μm) between

sexes and KOH treatment time (Table 6.2). BSE images revealed a darker coloured banding

on the endorcortical region of the samples. The total distance of dark band was measured:

148

there were no differences in the total width of the dark endocortical band between medial,

lateral, anterior and posterior locations, allowing pooling of measurements per specimen.

Once grouped, KOH treated samples (7-day and 14-day) for both sexes had significantly

(p<0.001) greater total width of the dark endocortical band compared to untreated (0-day)

samples (Table 6.2, Figure 6.2). There were no differences between sexes.

Table 6.2: Quantitative BSE results for tibiae of female and male KOH treated bone. KOH treatment time in days

Parameter

Sex 0 1 3 7 14

Female 177 ± 7 181 ± 7 184 ± 8 182 ± 7 175 ± 8 Peak grey level (pixels) Male 175 ± 9 181 ± 5 179 ± 8 188 ± 9 174 ± 8

Female 25 ± 3 27 ± 3 26 ± 2 27 ± 3 23 ± 2 FWHMH (pixels) Male 23 ± 3 25 ± 3 25 ± 2 26 ± 3 24 ± 3

Female 0.23 ± 0.06 0.20 ± 0.06 0.14 ± 0.07 0.14 ± 0.09 0.16 ± 0.06 Large Pore Density (#/mm2) Male 0.22 ± 0.07 0.26 ± 0.06 0.18 ± 0.11 0.18 ± 0.07 0.11 ± 0.06

Female 110 ± 4 106 ± 5 101 ± 5 105 ± 7 108 ± 4 Small Pore Density (#/mm2) Male 110 ± 4 115 ± 5 107 ± 8 108 ± 5 110 ± 5

Female 190 ± 45 259 ± 47 326 ± 47 357 ± 50a 409 ± 50a Total Length of Dark Band (μm) Male 220 ± 45 268 ± 53 298 ± 71 375 ± 43a 445 ± 47a

ap ≤ 0.05 versus untreated (0-day) KOH treatment time

149

Figure 6.2: BSE Images of (a) 0-day (untreated) female, (b) 0-day (untreated) male, (c) 14-day female and (d) 14-day male emu tibiae. Magnified regions of the width of the dark endocortical band are shown in (e) 14-day female and (f) 14-day male emu tibiae. For each image, the endocortical region is found on the right side. After 14-day KOH treatment, a dark band is apparent beginning from the endocortical region and moving towards the periosteal region (see white enclosed box, magnified images in (e) and (f) and the black arrows spanning from the endocortical side to the end of the dark band).

150

6.3.4. α-Chymotrypsin

The results for the percent collagen digested versus KOH treatment time are presented in

Figure 6.3. Two-way ANOVA indicated that the percent digested collagen varied

significantly with KOH treatment time and sex. Compared with untreated samples, the 14-

day KOH treated specimens exhibited an almost two-fold increase in the percentage of

digestible collagen (%DC). Furthermore, the amount of digestible collagen was significantly

greater in male emu tibiae compared to female emu tibiae at 3-day (p=0.03), 7-day (p=0.02)

and 14-day (p<0.001) KOH treatment time points.

Figure 6.3: Percent digested collagen as a function of KOH treatment time for female (pink circles) and male (blue circles) emu tibiae. Significant increases of percent digested collagen (%DC) were observed with increasing KOH treatment time. The * and ** indicate that both female and male 7-day and 14-day KOH treated samples respectively, had a significantly greater amount of digested collagen compared to untreated samples, as determined by the α-chymotrypsin assay. Significant differences were seen between sexes at 3-day, 7-day and 14-day KOH treatment time.

Multiple regression analyses indicated that the percentage of digested collagen had a strong

correlation with the mechanical properties (elastic modulus, failure stress, failure strain and

toughness) for both female and male emu tibiae (Figure 6.4). The elastic modulus (Figure 6.4

(a)) and failure strength (Figure 6.4 (b)) of the KOH treated groups decreased gradually with

0 1 3 7 140

10

20

30

40MaleFemale

KOH Treatment Time (days)

Dig

este

d C

olla

gen

(%

)

*

**

0 1 3 7 140

10

20

30

40MaleFemale

KOH Treatment Time (days)

Dig

este

d C

olla

gen

(%

)

*

**

151

increasing collagen degradation. Conversely, the failure strain (Figure 6.4 (c)) and toughness

(Figure 6.4 (d)) increased with increasing collagen degradation. KOH induced collagen

degradation exhibited the least impact on the elastic modulus of bone. The regression curves

were not different between female and male emu tibiae in any of these tests.

0 10 20 30 40 500

20

40

60

80

100

120

140

160

180

Male

Female

0 10 20 30 40 500

5

10

15

20

25

30Male

Female

Digested Collagen (%)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.29p<0.001

R2=0.25p=0.001

R2=0.47p<0.001

R2=0.49p<0.001

Digested Collagen (%)

0 10 20 30 40 500

1

2

3

4

5

Male

Female

0 10 20 30 40 500

1

2

3

4

5 Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.38p<0.001

R2=0.32p<0.001

R2=0.50p<0.001

R2=0.48p<0.001

Digested Collagen (%) Digested Collagen (%)

Figure 6.4: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the percent digested collagen (%DC) for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Strong correlations were found between these parameters.

152

6.3.5. DSC

The temperature at which the main denaturation peak occurred (Tpeak) and the enthalpy

values were unaffected by KOH treatment. The shape (peak width and height) of the

endotherm as well as Tonset were affected for the 7-day and 14-day KOH treated groups for

both sexes (Table 6.3). However, there were no differences between sexes. The major

transition peak during thermal denaturation for all groups occurred at the same temperature

(approximately 60 °C). However, the 14-day KOH treated groups showed a peak that was

lower and broader compared to the untreated (0-day) groups for both female and male emu

tibiae. The FWHMH, a measure of the heterogeneity of the molecular structure, increased 3

°C (p=0.03) for female bones and 5 °C (p<0.001) for male bones after 14-day KOH

treatment. This increase was coupled with a significant decrease in Tonset (a measure of

thermal stability) of 7 °C (p<0.001) for female and 8 °C (p<0.001) for male bones and a

significant (p<0.001 for both sexes) decrease in the Height of the peak (a measure of the

heterogeneity of the molecular structure). The changes in shape of the thermogram were

significant with KOH treatment (p<0.001). The enthalpies calculated from the area under the

peak were not found to be different (Table 6.3). Multiple regression analyses revealed limited

correlations between the FWHMH (Figure 6.5), Height (Figure 6.6) and Tonset (Figure 6.7)

with bone mechanical properties.

153

Table 6.3: Average female and male emu tibiae thermal characteristics from DSC. KOH treatment time in days Thermal

Characteristics

Sex

0 1 3 7 14 Female 48.5 ± 1.6 48.0 ± 1.6 47.0 ± 1.0 44.9 ± 1.3a 41.7 ± 1.0a

Tonset (°C) Male 47.7 ± 1.5 46.7 ± 1.4 45.3 ± 1.1 42.9 ± 1.3a 40.2 ± 0.7a

Female 60.9 ± 0.2 61.3 ± 0.4 60.2 ± 0.5 59.8 ± 0.3 59.8 ± 0.4

Tpeak (°C) Male 60.6 ± 0.4 61.0 ± 0.4 61.2 ± 0.6 59.9 ± 0.3 59.9 ± 0.4

Female 18.4 ± 1.1 19.4 ± 1.1 19.7 ± 1.0 20.8 ± 1.1 21.5 ± 0.9a

FWHMH (°C) Male 17.6 ± 0.8 19.0 ± 0.6 19.4 ± 1.2 21.1 ± 0.8a 23.0 ± 1.0a

Female 38.7 ± 0.8 36.7 ± 1.3 39.1 ± 1.3 38.5 ± 1.4 38.9 ± 0.9 Enthalpy (ΔH; J/g of collagen) Male 38.6 ± 1.2 39.5 ± 1.6 37.8 ± 1.1 40.4 ± 1.6 39.5 ± 1.8

Female 0.17 ± 0.01 0.16 ± 0.01 0.15 ± 0.01 0.15 ± 0.01 0.13 ± 0.01a Height (mW/mg of collagen) Male 0.16 ± 0.01 0.17 ± 0.01 0.15 ± 0.01 0.14 ± 0.01 0.13 ± 0.01a

ap ≤ 0.05 versus 0-day KOH treatment time

154

Figure 6.5: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the FWHM (from DSC curves) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Limited correlations were found between these parameters.

0 10 20 300

1

2

3

4

5

Male

Female

0 10 20 300

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.07p=0.068

R2=0.01p=0.427

R2=0.11p=0.019

R2=0.001p=0.957

FWHM (°C) FWHM (°C)0 10 20 30

0

1

2

3

4

5

Male

Female

0 10 20 300

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.07p=0.068

R2=0.01p=0.427

R2=0.11p=0.019

R2=0.001p=0.957

FWHM (°C) FWHM (°C)

0 10 20 300

20

40

60

80

100

120

140

160

180

Male

Female

0 10 20 300

5

10

15

20

25

30Male

Female

FWHM (°C)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.08p=0.055

R2=0.19p=0.002 R2=0.02

p=0.307

R2=0.002p=0.777

FWHM (°C)0 10 20 30

0

20

40

60

80

100

120

140

160

180

Male

Female

0 10 20 300

5

10

15

20

25

30Male

Female

FWHM (°C)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.08p=0.055

R2=0.19p=0.002 R2=0.02

p=0.307

R2=0.002p=0.777

FWHM (°C)

155

Figure 6.6: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the Height (from DSC curves) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Limited correlations were found between these parameters.

0.00 0.05 0.10 0.15 0.20 0.250

1

2

3

4

5

Male

Female

0.00 0.05 0.10 0.15 0.20 0.250

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.02p=0.328

R2=0.0002p=0.938

R2=0.007p=0.581

R2=0.004p=0.623

Height (mW/mg of collagen) Height (mW/mg of collagen)0.00 0.05 0.10 0.15 0.20 0.25

0

1

2

3

4

5

Male

Female

0.00 0.05 0.10 0.15 0.20 0.250

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.02p=0.328

R2=0.0002p=0.938

R2=0.007p=0.581

R2=0.004p=0.623

Height (mW/mg of collagen) Height (mW/mg of collagen)

0.00 0.05 0.10 0.15 0.20 0.250

20

40

60

80

100

120

140

160

180

Male

Female

0.00 0.05 0.10 0.15 0.20 0.250

5

10

15

20

25

30Male

Female

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.03p=0.18

R2=0.04p=0.228

R2=0.06p=0.068

R2=0.01p=0.689

Height (mW/mg of collagen) Height (mW/mg of collagen)0.00 0.05 0.10 0.15 0.20 0.25

0

20

40

60

80

100

120

140

160

180

Male

Female

0.00 0.05 0.10 0.15 0.20 0.250

5

10

15

20

25

30Male

Female

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.03p=0.18

R2=0.04p=0.228

R2=0.06p=0.068

R2=0.01p=0.689

Height (mW/mg of collagen) Height (mW/mg of collagen)

156

Figure 6.7: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the Tonset (from DSC curves) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Limited correlations were found between these parameters.

0 20 40 60 80 1000

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Tonset (°C)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.06p=0.083

R2=0.16p=0.006

R2=0.06p=0.068

R2=0.01p=0.689

Tonset (°C)0 20 40 60 80 100

0

20

40

60

80

100

120

140

160

180

Male

Female

0 20 40 60 80 1000

5

10

15

20

25

30Male

Female

Tonset (°C)

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.06p=0.083

R2=0.16p=0.006

R2=0.06p=0.068

R2=0.01p=0.689

Tonset (°C)

0 20 40 60 80 1000

1

2

3

4

5

Male

Female

0 20 40 60 80 1000

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.03p=0.278

R2=0.01p=0.637

R2=0.001p=0.868

R2=0.05p=0.162

Tonset (°C) Tonset (°C)0 20 40 60 80 100

0

1

2

3

4

5

Male

Female

0 20 40 60 80 1000

1

2

3

4

5

6Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.03p=0.278

R2=0.01p=0.637

R2=0.001p=0.868

R2=0.05p=0.162

Tonset (°C) Tonset (°C)

157

6.3.6. SDS-PAGE

The electrophoretic profiles for select samples from each treatment group are shown in

Figure 6.8 (a)) for female (left) and male (right) emu tibiae. The representative densitometric

scans for the selected samples from each treatment group of female (left) and male (right)

emu tibiae are shown in Figure 6.8 (b)), where the peaks correspond to protein bands on the

gel and the area under the peak is related to the amount of protein in the band. Although the

same amount of protein was loaded in each lane, as determined by the hydroxyproline assay,

the peak areas changed when comparing the different groups and between sexes. The

intensity of the α1(I) and α2(I) protein bands appear to decrease with KOH treatment time

where male emu tibiae seem to be more affected by KOH treatment compared to female emu

tibiae. The pepsin band intensity remained constant in each lane, indicating consistent

loading. Collagen degradation fragment area (shown to the extreme right on the

representative densitometric scans for both sexes) increases with KOH treatment time with

no differences between sexes. The results from these scans, represented as relative peak

intensity band area for the different treatment groups, are presented in Figure 6.8 (c)). A

decrease in α-chain band intensity was observed with KOH treatment time for both sexes.

This decrease was greater for male samples compared to female samples for the scans shown

in Figure 6.8 (b)).

Figure 6.8: SDS-PAGE results: (a) Analysis of proteins from select samples from each treatment group for female (left) and male (right) emu tibiae. Lane 1: Std, standard; Lane 2: RTC, rat tail collagen; Lanes 3-4: untreated (0-day); Lanes 5-6: 1-day; Lanes 7-8: 3-day; Lanes 9-10: 7-day; Lanes 11-12: 14-day KOH treatment time.

158

Figure 6.8: (b) Representative densitometric scans of proteins from (a) are shown to indicate more clearly changes in protein revealed on SDS-PAGE gels. Note that the intensity of α1(I) and α2(I) protein bands decrease with KOH treatment time. Figure 6.8: (c) Plot of average relative peak intensity band area versus KOH treatment time for female (pink circles) and male (blue circles) emu tibiae from (a). Decreased α-chain band intensity indicates collagen fragmentation due to KOH treatment.

0 1 3 7 148000

10000

12000

14000

16000

18000

20000

0 1 3 7 148000

10000

12000

14000

16000

18000

20000

KOH Treatment Time (days)

Col

lage

n pe

ak a

rea

inte

nsity

KOH Treatment Time (days)

Col

lage

n pe

ak a

rea

inte

nsity

c)

0 1 3 7 148000

10000

12000

14000

16000

18000

20000

0 1 3 7 148000

10000

12000

14000

16000

18000

20000

KOH Treatment Time (days)

Col

lage

n pe

ak a

rea

inte

nsity

KOH Treatment Time (days)

Col

lage

n pe

ak a

rea

inte

nsity

c)

159

Similar results were seen after analysis of all samples within each group. Figure 6.9

represents the average percent change in α-chain band intensity as a function of KOH

treatment time for all samples. Overall, α-chain band intensity decreased 38% for male emu

tibiae and 28% for female emu tibiae after 14-day KOH treatment. Significant differences in

α-chain band intensity between female and male emu tibiae were observed for 7-day

(p=0.002) and 14-day (p=0.005) KOH treated groups. Multiple regression analyses of the

average α-chain band intensity with the mechanical properties yielded similar correlations to

the α-chymotrypsin digestion results (Figure 6.10).

Figure 6.9: Average percent change in α-chain band intensity area for all samples as a function of KOH treatment time for female (pink circles) and male (blue circles) emu tibiae. Significant decreases in average α-chain peak intensity with increasing KOH treatment time and significant differences between female and male emu tibiae at 7-day and 14-day KOH treated groups. The * and ** indicate that both female and male 7-day and 14-day KOH treated samples respectively, had a significantly greater decrease in α-chain band intensity compared to untreated groups.

0 1 3 7 1450

60

70

80

90

100

FemaleMale

KOH Treatment Time (days)

Ave

rage

α-C

hai

n b

and

Inte

nsi

tyP

erce

nt

Ch

ange

(%

) ***

0 1 3 7 1450

60

70

80

90

100

FemaleMale

KOH Treatment Time (days)

Ave

rage

α-C

hai

n b

and

Inte

nsi

tyP

erce

nt

Ch

ange

(%

) ***

160

Figure 6.10: Regression of (a) elastic modulus, (b) failure stress, (c) failure strain and (d) toughness with respect to the collagen peak area intensity (from SDS-PAGE) of bone for female (pink circles) and male (blue circles) emu tibiae, with the correlation coefficient (R2) and p-values being presented. Strong correlations were found between these parameters.

0 10000 20000 300000

20

40

60

80

100

120

140

160

180

Male

Female

0 10000 20000 300000

5

10

15

20

25

30Male

Female

Collagen peak area intensity Collagen peak area intensity

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.23p<0.001

R2=0.18p=0.001

R2=0.40p<0.001

R2=0.37p<0.001

0 10000 20000 300000

20

40

60

80

100

120

140

160

180

Male

Female

0 10000 20000 300000

5

10

15

20

25

30Male

Female

Collagen peak area intensity Collagen peak area intensity

a) Elastic Modulus (GPa) b) Failure Strength (MPa)

R2=0.23p<0.001

R2=0.18p=0.001

R2=0.40p<0.001

R2=0.37p<0.001

0 10000 20000 300000

1

2

3

4

5

Male

Female

0 10000 20000 300000

1

2

3

4

5Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.33p<0.001

R2=0.36p<0.001

R2=0.32p<0.001

R2=0.33p<0.001

Collagen peak area intensity Collagen peak area intensity0 10000 20000 30000

0

1

2

3

4

5

Male

Female

0 10000 20000 300000

1

2

3

4

5Male

Female

c) Failure Strain (%) d) Toughness (mJ/mm3)

R2=0.33p<0.001

R2=0.36p<0.001

R2=0.32p<0.001

R2=0.33p<0.001

Collagen peak area intensity Collagen peak area intensity

161

6.3.7. Polarized light microscopy

Histological sections stained with picro-sirius red with and without polarization are shown in

Figure 6.11 for both female and male emu tibiae after 14-day KOH treatment. Images from

polarized light microscopy revealed differences in structure in terms of pattern distribution of

preferentially oriented collagen between the periosteal and endocortical regions for both

sexes. Collagen oriented in a similar direction will have the same brightness/colour, whereas

collagen oriented in other directions result in different brightness/colours. The endocortical

regions appear to have different amounts of layers with alternating birefringence (Figure

6.11), indicating a less organized structure. Both endocortical and periosteal regions of

untreated samples for both sexes had similar appearances to the periosteal region seen in the

14-day KOH treated samples (Figure 6.11).

162

Figure 6.11: Images of transcortical sections of demineralized 14-day KOH treated male ((a), (b), (c), (d)) and female ((e), (f), (g), (h)) emu tibiae, viewed under non-polarized ((a), (c), (e) and (g)) and polarized ((b), (d), (f) and (h)) light. Note the differences in structure in terms of pattern distribution of preferentially oriented collagen between the periosteal and endocortical regions for both sexes.

163

6.4. Discussion

In our initial study [7], KOH treatment was shown to significantly compromise the

mechanical properties of bone with a minimal amount of collagen loss and no significant

changes in geometry or mineral content, as determined by DXA. It was hypothesized that

these observed changes may result from in situ collagen degradation. The purpose of this

study was to determine if the collagen and mineral were indeed affected by the KOH

treatment and could be correlated to previously observed altered mechanical properties.

6.4.1. Mineral characterization

KOH treatment had no effect on bone mineralization and bone microhardness for both

female and male emu tibiae. These results may be influenced by the lack of sensitivity to

detecting changes in bone mineral at the microstructure level and the fact that both

techniques were performed on specimens at one cross-section only. Similarly, XRD analysis

revealed no differences in crystal size for both sexes. These results are not surprising as KOH

does not affect bone mineral [7,31,32] and may be of importance in understanding the

mechanisms by which certain bone disorders alter bone collagen without causing changes to

the bone mineral.

For example, cortical bone samples irradiated with gamma radiation showed small changes

in the elastic behaviour under static testing but a significant decrease in fracture resistance

under cyclic loading [50]. The authors determined by SDS-PAGE that gamma radiation

caused a decrease of intact alpha-chains resulting in cleavage of the collagen backbone [50].

Similarly, cortical bone samples incubated in ribose caused an accumulation of advanced

glycation end products (AGEs), which increased the formation of collagen crosslinks,

resulting in brittle bone [57]. Finally, patients with Type 2 diabetes have been shown to have

increases in fractures despite having increased BMD [58], suggesting the collagen may be

affected.

BSE images of female and male 7-day and 14-day KOH treated samples revealed darker

coloured banding on the endocortical region, suggesting that KOH treatment affected the

bone mineral. However, this was not reflected in the bone mineralization analysis,

164

microhardness testing or XRD analysis results. This suggests that alternative techniques may

be needed to detect the changes in bone mineral due to KOH treatment. The darker coloured

banding on the endocortical region may be due to KOH penetration through the samples

during the endocortical treatment time with longer time periods resulting in a larger distance

of penetration. The KOH may be penetrating into the bone and affecting not only the

collagen but also the mineral-collagen interface, as discussed below.

6.4.2. Collagen degradation

The percentage of collagen digested by α-chymotrypsin significantly increased with

increasing KOH treatment time for both female and male emu tibiae. The degraded collagen

consequently altered the properties of the matrix, leading to impaired bone mechanical

properties, consistent with previous investigations [7,28,59,60]. These results suggest that

KOH may be unwinding the collagen (denaturation). Alternatively, KOH may be penetrating

into the bone, swelling the collagen, resulting in a disorganized structure that is more

susceptible to general proteolysis. Swelling followed by in situ collagen degradation by KOH

is also an additional possibility. α-chymotrypsin may also digest damaged and not solely

denatured collagen, which may cause an overestimation in digested content. It has been

shown that α-chymotrypsin digests degraded collagen without affecting intact native

polymeric collagen [61,62], but KOH treated collagen may not necessarily be native.

Furthermore, some studies suggest that the α-chymotrypsin preparations may be

contaminated with traces of trypsin, which would result in overestimation of the amount of

degraded collagen. In this study, α-chymotrypsin-TLCK (treated with 1-chloro-3-tosylamido-

7-amino-2-heptanone) was chosen as the TLCK treatment inhibits trypsin activity, leading to

greater specificity [63,64].

DSC was used in this study to characterize the thermal stability of collagen (the resistance of

the protein molecule to unfolding) and the heat required for thermal denaturation in female

and male untreated and KOH treated groups. DSC has been shown to be sensitive to the

amount of collagen crosslinks [65] and to the level of hydration of collagen [25]. DSC has

also been used in the research of thermodynamic characteristics of various collagenous

tissues such as human cartilage and intervetebral discs [66-68], bovine tail tendons [21],

165

mineralizing turkey leg tendons [69] and bone [22-25,27]. Furthermore, DSC has been

successfully used as a tool for investigating collagen and its degradation by correlating

changes in thermal denaturation temperature and/or changes in enthalpy of denaturation

[67,70,71].

In this study, both the temperature of maximum heat flow, Tpeak and the enthalpy were

unaffected in KOH treated groups. However, the shape of the thermogram was altered for

both sexes, indicating a change in the intermolecular collagen structure of the KOH treated

groups [22]. A change in enthalpy would indicate denaturation of the collagen molecule

[21,22], whereas changes in Tonset, Height and FWHMH would suggest that KOH treatment

causes disruption of fibrillar structure and collagen degradation [21,22]. In fact, a broad

shape of the peak has been shown to correspond to an increased dispersion in terms of

stability of the collagen [21,22,47]. Knott et al. examined the thermal characteristics of

normal and osteoporotic avian bone and observed lower and broader thermogram peaks for

the osteoporotic bone with no differences in peak temperature or enthalpies compared to

normal bone [22]. These DSC thermogram results are in agreement with our findings. The

authors postulated this was due to increased variability in thermal stability in the osteoporotic

bone collagen, perhaps due to increased disorder of the fibrillar structure [22]. Furthermore, a

recent in vitro tensile overload tendon model that used DSC to measure the thermal

behaviour of collagen, reported similar results due to intermolecular mechanical deformation,

which alters the fibrillar structure and organization [21]. Changes in the shape of the

thermogram have also been associated with differences in post-translational modifications of

collagen [22]. Furthermore, Herbage et al. showed that the DSC curves obtained for

demineralized normal bone and demineralized Osteogenesis Imperfecta (OI) bone were

identical and contained a single peak at 60 °C [66]. They concluded that the collagen thermal

stability had not been influenced by changes of the nature of the reducible intermolecular

crosslinks [66]. Therefore, it may be inferred that the mechanism of KOH on emu tibiae did

not affect the collagen crosslinks, as there were no changes in Tpeak or enthalpy observed

between treatment groups or between sexes. The DSC results of no enthalpy change, increase

in FWHMH and decrease in Height and Tonset, measured in this study, suggest that KOH

166

treatment increases disorder of the fibrillar structure causing collagen degradation rather than

denaturation.

The interpretation of the DSC curves must be done with regards to the way in which the

sample was prepared [23]. It has been reported that bone shows different thermal behaviour

if it is ground or studied in pieces and this may affect the shape of the DSC curve and the

change in the enthalpy of denaturation values [23,24]. Sample preparation may be the reason

for the unaffected enthalpy values as well as the weak correlations observed in this study and

this interesting issue is currently being investigated by our group.

Analysis by SDS-PAGE revealed a general decrease in α-chain band intensity with KOH

treatment time for both female and male emu tibiae, indicating collagen fragmentation due to

KOH treatment. Furthermore, densitometric scans revealed possible lower molecular weight

fragments with increasing KOH treatment time for both sexes, indicating chain scission in

helical regions [53]. These results are in agreement with the α-chymotrypsin digestion and

DSC results and further supports the hypothesis that KOH treatment causes in situ collagen

degradation rather than denaturation. The high pH of 14 of the KOH solution may have

caused fibrillar swelling and α-chain hydrolysis (degradation). This combination could

explain the results from all these assays. It would be interesting to investigate the

susceptibility of the proteolytic treatment to known, bone-rich collagenases such as cathepsin

K and MMPs. Furthermore, pepsin digestion, used to prepare for gel electrophoresis,

specifically cleaves native polymeric collagen in the telopeptide regions, circumventing the

covalent crosslinks and therefore, should liberate both intact collagen molecules and any

fragments from the insoluble crosslinked matrix without significant additional degradation

[42,49,51]. However, KOH treated collagen may not be native, which pepsin may also be

digesting.

Collagen degradation is also confirmed by the images obtained from polarized light

microscopy where the endocortical region of the 14-day KOH treated bones show a different

microstructure (due to the endocortical KOH treatment) compared to the periosteal region for

both sexes. The different regional structures are an indication of collagen that is no longer

167

organized in the same orientation as the surrounding collagen. This disorientation may be due

to KOH penetration, which causes swelling and a change in the lattice structure, resulting in

a less densely packed and less ordered structure.

Multiple regression analyses from α-chymotrypsin digestion and gel electrophoresis showed

the amount of degraded collagen had a strong correlation with the failure stress, failure strain

and toughness. The weakest correlation was found with the elastic modulus, indicating that

collagen degradation had little effect on the stiffness of bone. This is not surprising as the

maximum contribution of collagen to bone stiffness has been shown to be less than 1% [17].

However, it is important to note that after KOH treatment, the specimens remained intact,

suggesting that KOH treatment did not lead to a total disintegration of the collagen network

but rather in situ collagen degradation. The endocortical treatment with KOH likely resulted

in an inner zone where collagen was significantly altered, with less disruption as the distance

increased [7].

The regression curves were not significantly different between sexes, contradictory to results

shown in Chapter 3 – Initial Study, which indicated that male emu tibiae had significantly

decreased failure stress and increased failure strain and toughness compared to female emu

tibiae [7]. Since these differences are no longer prevalent when correlating to the percentage

of degraded collagen and regression analysis explained on average only 40% of the original

variability for both sexes, this suggests an additional mechanism of action by KOH other

than simply in situ collagen degradation, such as damage to the mineral-collagen interface.

6.4.3. Partial debonding of the collagen-mineral interface

Bone is a composite material whose mechanical properties depend on the characteristics of

the mineral, collagen and the interaction between the mineral and the collagen [6,7]. KOH

treatment causes collagen degradation, which in turn may affect the mineral-organic

interface, altering the way the overall bone behaves. Several authors have tested and

published data regarding decreases of bone quality due to changes in bonding between the

mineral and organic phases [7,72-75]. Kotha and Guzelsu modeled the stress transfer

between the mineral and organic components of bone in order to investigate the mineral-

168

collagen interface [74]. Their model showed that a composite with lower bonding would

have decreased elastic modulus, yield and ultimate stresses, while the ultimate strain would

be increased. They concluded that bone with decreased bonding between the mineral and

collagen would become more ductile [74]. Their results are very similar to our observations

and suggests a mechanism for the significant alteration of mechanical properties in the

absence of geometric or compositional changes.

Walsh and Guzelsu postulated that ion permeation modification of chemical bonds between

the collagen and mineral may affect bone interfacial bonding [76]. Bone fluid is rich in

potassium and correspondingly poor in Na+ and Ca+ [77]. The bone fluid is actively

maintained by the metabolic activity of bone cells [78]. As such, the bonding strength and

bond distribution at the collagen-mineral interface could easily be altered due to the different

binding and displacing capabilities of these inorganic free ions in competition with collagen

polar groups for mineral binding sites. Specifically, it has been shown that in vitro soaking of

bone specimens in fluoride or phosphate ion solutions for several days alters the mechanical

properties similarly to those reported in Chapter 3 – Initial Study (lower elastic modulus,

failure stress and higher failure strain and toughness) [7,76,79]. The authors attributed these

alterations to compromised interfacial bonding between the mineral and collagen as a result

of the ability of these free ions to compete with the organic negative groups on the collagen

or noncollagenous proteins for the binding sites on the bone mineral surface [76,79]. This

may be a mechanism of action of KOH, where the free potassium and/or hydroxyl ions are

permanently incorporated into the bone structure, debonding the links between the mineral

and matrix. Partial debonding between the mineral and organic matrix may be an additional

mechanism together with collagen degradation that is responsible for the significant observed

alteration of mechanical properties.

6.4.4. Sex differences

There were significant differences observed in the amount of digested collagen from the α-

chymotrypsin digestion and gel electrophoresis between KOH treated female and male emu

tibiae. Specifically, male samples had a significantly larger amount of digested collagen

compared to female emu tibiae at 7-day and 14-day KOH treatment. In skeletally mature

169

emus, sexual dimorphism in size (geometry), mass and BMD have been reported [7,80,81].

Male emus carry out the egg incubation period (approximately 8 weeks) [81], which may

lead to adverse effects of disuse on bone. During periods of disuse or immobilization, bones

experience altered magnitudes of loading. Disuse decreases the stresses in bones, which leads

to an adaptive remodeling response, bone atrophy and decreased mechanical properties [82].

For example, sixteen weeks of immobilization significantly decreased the mechanical

properties of canine cortical bone [82]. It has been shown that the remobilization period to

recover lost bone is much longer than the disuse period. Dogs and rats require a

remobilization period of 2-3 times the length of disuse period to recover all the bone lost

during disuse [82-84] and bone loss may continue during remobilization [85]. It is unknown

whether the male emus used in this study were sacrificed before or after incubation.

However, the inactivity or insufficient remobilization period and the sexual dimorphisms

may be the reasons why male emu tibiae are more susceptible to KOH attack, resulting in a

higher amount of degraded collagen.

6.5. Conclusions

Bone quality is a complex property that we do not fully understand and the way that KOH

exposure alters bone quality is still partially unclear. In addition to the collagen degradation

results presented in this study, the observed decreases in bone mechanical properties may be

the manifestation of the involvement of other mechanisms induced by KOH treatment such

as partial debonding of the collagen-mineral interface. To clarify these underlying

mechanisms, further investigations are needed. The results of this study support our

hypothesis that the compromised mechanical properties previously reported [7], are partly a

result of in situ collagen degradation. It is also important to remember that bone is a

composite material comprising mineral, collagen, noncollagenous proteins and solutes. It is a

limitation of this study that the noncollagenous proteins were not studied. Bone mechanical

properties are dependent on the mineral, the collagen and the interaction between mineral and

collagen [1,7]. Thus, the in situ collagen degradation measured in this study serves only as a

measure of a combined effect of the above factors.

170

6.6. Chapter Summary

This study has shown, with the use of several mineral and collagen characterization

techniques, that endocortical KOH treatment causes collagen degradation without affecting

the mineral phase. Furthermore, polarized light microscopy images revealed differences in

collagen structure in terms of pattern distribution of preferentially oriented collagen between

the periosteal and endocortical regions of 14-day KOH treated samples. These results help

explain the previously reported altered mechanical properties in Chapter 3 – Initial Study as

well as the altered fatigue properties in Chapter 4 - Fatigue. However, multiple regression

analysis could not fully explain the observed results (average R2=0.40), suggesting an

additional mechanism of action by KOH, such as changes to the mineral-collagen interface.

171

6.7. References

[1] Boskey A. (2001) Bone Mineralization. In: Cowin SC, editor. Bone Mechanics Handbook. CRC Press. Boca Raton. Florida. pp. 1-33.

[2] Rho JY, Kuhn-Spearing L, Zioupos P. (1998) Mechanical properties and the hierarchical

structure of bone. Med.Eng.Phys. 20:92-102. [3] Jee WSS. (2001) Integrated Bone Tissue Physiology. In: Cowin SC, editor. Bone Mechanics

Handbook. CRC Press. Boca Raton. Florida. pp. 1-68. [4] Nyman JS, Reyes M, Wang X. (2005) Effect of ultrastructural changes on the toughness of

bone. Micron. 36:566-582. [5] Currey JD, Brear K, Zioupos P. (1996) The effects of ageing and changes in mineral content

in degrading the toughness of human femora. J.Biomech. 29:257-260. [6] Li X, Agrawal M, Wang X. (2003) Age dependence of in situ thermostability of collagen in

human bone. Calcif.Tissue.Int. 72:513-518. [7] Wynnyckyj C, Omelon S, Savage K, Damani M, Chachra D, Grynpas MD. (2009) A new

tool to assess the mechanical properties of bone due to collagen degradation. Bone. 44:840-848.

[8] Garnero P, Schott AM, Prockop D, Chevrel G. (2009) Bone turnover and type I collagen C-

telopeptide isomerization in adult osteogenesis imperfecta: associations with collagen gene mutations. Bone. 44:461-466.

[9] Baron R. Anatomy and Ultrasturcture of Bone. (1993) In: Favus MJ, editor. Primer on the

Metabolic Bone Diseases and Disorders of Mineral Metabolism. Lippincott-Raven. New York. pp. 3-9.

[10] Termine JD. Bone Matrix Proteins and the Mineralization Process. (1993) In: Favus M,J

editor. Primer on the Metabolic Bone Diseases and Disorders of Mineral Metabolism. Lippincott-Raven. New York. pp. 21-25.

[11] Wang X, TeKoppele JM, Athanasiou KA. (2000) Effect of collagen denaturation on the

toughness of bone. Clin.Orthop.Relat.Res. 371:228-239. [12] Wang X, Li X, Bank RA, Agrawal CM. (2002) Effects of collagen unwinding and cleavage

on the mechanical integrity of the collagen network in bone. Calcif.Tissue.Int. 71:186-192. [13] Bailey AJ, Wotton SF, Sims TJ, Thompson PW. (1993) Biochemical changes in the collagen

of human osteoporotic bone matrix. Connect.Tissue.Res. 29:119-132. [14] Oxlund H, Barckman M, Ortoft G, Andreassen TT. (1995) Reduced concentrations of

collagen cross-links are associated with reduced strength of bone. Bone. 17:365S-371S. [15] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

172

[16] Bank RA, Krikken M, Beekman B, Sttop R, Maroudas A, Lafebers FPJG, TeKoppele JM. (1997) A simplified measurement of degraded collagen in tissues: application in healthy, fibrillated and osteoarthritic cartilage. Matrix.Biology. 16:233-243.

[17] Wang X, Bank RA, TeKoppele JM, Agrawal CM. (2001) The role of collagen in determining

bone mechanical properties. J.Orthop.Res. 19:1021-1026. [18] Viguet-Carrin S, Farlay D, Bala Y, Munoz F, Bouxsein ML, Delmas PD. (2008) An in vitro

model to test the contribution of advanced glycation end products to bone biomechanical properties. Bone. 42:139-149.

[19] Garnero P, Borel O, Gineyts E, Duboeuf F, Solbreg H, Bouxsein ML, Christiansen C, Delmas

PD. (2006) Extracellular post-translational modifications of collagen are major determinants of biomechanical properties of fetal bovine cortical bone. Bone. 38:300-309.

[20] Danielsen CC. (1990) Age-related thermal stability and susceptibility to proteolysis of rat

bone collagen. Biochem.J. 272:697-701. [21] Willett TL, Labow RS, Lee JM. (2008) Mechanical overload decreases the thermal stability

of collagen in an in vitro tensile overload tendon model. J.Orthop.Res. 12:1605-1610. [22] Knott L, Whitehead CC, Fleming RH, Bailey AJ. (1995) Biochemical changes in the

collagenous matrix of osteoporotic avian bone. Biochem.J. 310:1045-1051. [23] Lozano LF, Pena-Rico MA, Heredia A, Ocotlan-Flores J, Gomez-Cortes A, Velazquez R,

Belio IA. (2003) Thermal analysis study of human bone. J.Mater.Sci. 38:4777-4782. [24] Nielsen-Marsh CM, Hedges REM, Mann T, Collins MJ. (2000) A preliminary investigation

of the application of differential scanning calorimetry to the study of collagen degradation in archaeological bone. Thermochim.Acta. 365:129-139.

[25] Trebacz H, Wojtowicz K. (2005) Thermal stabilization of collagen molecules in bone tissue.

Int.J.Biol.Macromol. 37:257-262. [26] Kronick PL, Cooke P. (1996) Thermal stabilization of collagen fibers by calcification.

Connect.Tissue.Res. 33:275-282. [27] Kubisz L, Mielcarek S. (2005) Differential scanning calorimetry and temperature dependence

of electric conductivity in studies on denaturation process of bone collagen. J.Non.Cry.Sol. 351:2935-2939.

[28] Wang X, Shen X, Li X, Agrawal CM. (2002) Age-related changes in the collagen network

and toughness of bone. Bone. 31:1-7. [29] Wang X, Li X, Shen X, Agrawal CM. (2003) Age-related changes of noncalcified collagen in

human cortical bone. Ann.Biomed.Eng. 31:1365-1371. [30] Burr DB. (2002) The contribution of the organic matrix to bone's material properties. Bone.

31:8-11.

173

[31] Abe K, Hashizume H, Ushiki T. (1992) An EDTA-KOH method to expose bone cells for scanning electron microscopy. J.Electron.Microsc.(Tokyo). 41:113-115.

[32] Arsenault AL, Grynpas MD. (1988) Crystals in calcified epiphyseal cartilage and cortical

bone of the rat. Calcif.Tissue.Int. 43:219-25. [33] Broz JJ, Simske SJ, Corley WD, Greenberg AR. (1997) Effects of deproteinization and

ashing on site-specific properties of cortical bone. J.Mater.Sci.Mater.Med. 8:395-401. [34] Walsh WR, Labrador DP, Kim HD, Guzelsu N. (1994) The effect of in vitro fluoride ion

treatment on the ultrasonic properties of cortical bone. Ann.Biomed.Eng. 22:404-15. [35] Burr DB, Miller L, Grynpas MD, Li J, Boyde A, Mashiba T, Hirano T, Johnston CC. (2003)

Tissue mineralization is increased following 1-year treatment with high doses of bisphosphonates in dogs. Bone. 33:960-969.

[36] Klug H, Alexander L. (1974) X-ray diffraction procedures for polycrystalline and amorphous

materials. Wiley. New York. [37] Currey JD, Brear K. (1990) Hardness, Young's modulus and yield stress in mammalian

mineralized tissues. J.Mat.Sci.Mat.Med. 1:14-20. [38] Grynpas M. (1993) Age and disease-related changes in the mineral of bone. Calcif.Tissue.Int.

53 Suppl. 1:S57-S64. [39] An and Draughn [Eds.] (2000) Mechanical Testing of Bone and the Bone-Implant Interface.

CRC Press LLC. Boca Raton. Florida. [40] Boyde A, Jones SJ. (1983) Back-scattered electron imaging of skeletal tissues.

Metab.Bone.Dis.Relat.Res. 5:145-150. [41] Domenicucci C, Goldberg HA, Hofmann T, Isenman D, Wasi S, Sodek J. (1988)

Characterization of porcine osteonectin extracted from foetal calvariae. Biochem.J. 253:139-151.

[42] Grynpas MD, Tupy JH, Sodek J. (1994) The distribution of soluble, mineral-bound and

matrix-bound proteins in osteoporotic and normal bones. Bone. 15:505-513. [43] Sodek KL, Tupy JH, Sodek J, Grynpas MD. (2000) Relationships between bone protein and

mineral in developing porcine long bone and calvaria. Bone. 26:189-198. [44] van der Harst MR, Brama PAJ, van de Lest CHA, Kiers GH, DeGroot J, van Weeren PR.

(2004) An integral biochemical analysis of the main constituents of articular cartilage, subchondral and trabecular bone. Osteoarthritis.Cartilage. 12:752-761.

[45] Woessner JF Jr. (1976) Determination of hydroxyproline in connective tissues. In: Hall D,

editor. The methodology of connective tissue research. Joynson-Bruvvers Ltd. Oxford. pp. 227-233.

[46] Reddy GK, Enwemeka CS. (1996) A simplified method for the analysis of hydroxyproline in

biological tissues. Clin.Biochem. 29:225-229.

174

[47] Gelinsky M, Welzel PB, Simon P, Bernhardt A, Konig U. (2008) Porous three-dimensional scaffolds made of mineralised collagen: preparation and properties of a biomimetic nanocomposite material for tissue engineering of bone. Chem.Eng.J. 137:84-96.

[48] Koon HEC, Loreille OM, Covington AD, Christensen AF, Parsons TJ, Collins MJ. (2008)

Diagnosing post-mortem treatments which inhibit DNA amplification from US MIAs buried at the Punchbowl. Forensic.Sci.Int. 178:171-177.

[49] Butkowski RJ, Noelken ME, Hudson BG. (1982) Estimation of the size of collagenous

proteins by electrophoresis and gel chromatography. In: Cunningham LW, Frederiksen DW, editors. Methods of Enzymology. Academic Press. New York. pp. 410-423.

[50] Akkus O, Belaney RM, Das P. (2005) Free radical scavenging alleviates the biochemical

impairment of gamma radiation sterilized bone tissue. J.Orthop.Res. 23:838-845. [51] Baht GS, Hunter GK, Goldberg HA. (2008) Bone sialoprotein-collagen interaction promotes

hydroxyapatite nucleation. Matrix.Biology. 27:600-608. [52] Furthmayr H, Timpl R. (1971) Characterization of collagen peptides by sodium

dodecylsulfate-polyacrylamide electrophoresis. Anal.Biochem. 41:510-516. [53] Carmichael DJ, Dodd CM, Veis A. (1977) The solubilization of bone and dentin collagens by

pepsin effect of cross-linkages and non-collagen components. Biochim.Biophys.Acta. 491:177-192.

[54] Penkova R, Goshev I, Gorinstein S, Nedkov P. (1999) Stability of collagen during

denaturation. J.Protein.Chem. 18:397-401. [55] Junqueira LC, Bignolas G, Brentani RR. (1979) Picrosirius staining plus polarization

microscopy, a specific method for collagen detection in tissue sections. Histochem.J. 11:447-455.

[56] Dayan D, Hiss Y, Hirshberg A, Bubis JJ, Wolman M. (1989) Are the polarization colors of

picrosirius red-stained collagen determined only by the diameter of the fibers? Histochemistry. 93:27-29.

[57] Vashishth D, Gibson GJ, Khoury JI, Schaffler MB, Kimura J, Fyhrie DP. (2001) Influence of

nonenzymatic glycation on biomechanical properties of cortical bone. Bone. 28:195-201. [58] Melton LJ III, Leibson CL, Achenbach SJ, Therneau TM, Khosla S. (2008) Fracture risk in

type 2 diabetes: update of a population-based study. J.Bone.Miner.Res. 23:1334-1342. [59] Bailey AJ, Wotton SF, Sims TJ, Thompson PW. (1993) Biochemical changes in the collagen

of human osteoporotic bone matrix. Connect.Tissue.Res. 29:119-132. [60] Miles CA, Bailey AJ. (2001) Thermally labile domains in the collagen molecule. Micron.

32:325-332. [61] Hollander AP, Heathfield TF, Webber C, Iwata Y, Bourne R, Rorabeck C, Poole AR. (1994)

Increased damage to type II collagen in osteoarthritic articular cartilage detected by a new immunoassay. J.Clin.Invest. 93:1722-1732.

175

[62] van Wart HJ, Bond MD. (1982) An accurate, quantitative assay for collagenase activity based on the synergistic hydrolysis of collagen. Anal.Biochem. 120:151-158.

[63] Shaw E, Mares-Guia M, Cohen W. (1965) Evidence for an active-center histidine in trypsin

through use of a specific reagent, 1-chloro-3-tosylamido-7-amino-2-heptanone, the chloromethyl ketone derived from n-tosyl-l-lysine. Biochem. 4:2219-2224.

[64] Willet TL, Labow RS, Avery NC, Lee MJ. (2007) Increased proteolysis of collagen in an in

vitro tensile overload tendon model. Ann.Biomed.Eng. 35:1961-1972. [65] Knott L, Bailey AJ. (1998) Collagen cross-links in mineralizing tissues: a review of their

chemistry, function, and clinical relevance. Bone. 22:181-187. [66] Herbage D, Borsali F, Buffevant C, Flandin F, Aguercif M. (1982) Composition, cross-

linking and thermal stability of bone and skin collagens in patients with osteogenesis imperfecta. Metab.Bone.Dis.Rel.Res. 4:95-101.

[67] Doman I, Toth Gy, Illes T, Lorinczy D. (2001) Differential scanning calorimetric

examination of the human intervertebral disc: A preliminary study. Thermochim.Acta. 376:117-122.

[68] Than P, Vermes Cs, Schaffer B, Lorinczy D. (2000) Differential scanning calorimetric

examination of human cartilage. Thermochim.Acta. 346:147-151. [69] Knott L, Tarlton JF, Bailey AJ. (1997) Chemistry of collagen cross-linking: biochemical

changes in collagen during the partial mineralization of turkey leg tendon. Biochem.J. 322:535-542.

[70] Miles CA, Wardale RJ, Birch HL, Bailey AJ. (1994) Differential scanning calorimetric

studies of superficial digital flexor tendon degeneration in the horse. Equine.Vet.J. 26:291-296.

[71] Miles CA. (1994) Differential scanning calorimetry (DSC): protein structure probe useful for

the study of damaged tendons. Equine.Vet.J. 26:255-256. [72] Bundy KJ. (1985) Determination of mineral-organic bonding effectiveness in bone--

theoretical considerations. Ann.Biomed.Eng. 13:119-135. [73] Walsh WR, Ohno M, Guzelsu N. (1994) Bone composite behaviour: effects of mineral-

organic bonding. J.Mater.Sci.Mater.Med. 5:72-79. [74] Kotha SP, Guzelsu N. (2000) The effects of interphase and bonding on the elastic modulus of

bone: changes with age-related osteoporosis. Med.Eng.Phys. 22:575-585. [75] Gupta HS, Fratzl P, Kerschnitzki M, Benecke G, Wagermaier W, Kirchner HO. (2007)

Evidence for an elementary process in bone plasticity with an activation enthalpy of 1 eV. J.R.Soc.Interface. 4:277-282.

[76] Walsh WR, Guzelsu N. (1993) The role of ions and mineral-organic interfacial bonding on

the compressive properties of cortical bone. Biomed.Mater.Eng. 3:75-84.

176

[77] Neuman WF, Ramp WK. (1971) The concept of a bone membrane: some implications. In: Nichols G Jr, Wasserman RH, editors. Cellular Mechanisms for Calcium Transfer and Homeostasis. Academic Press. New York. pp. 197-209.

[78] Rasmussen H, Border P. (1974) The Physiological and Cellular Basis of Metabolic Bone

Disease. The Williams and Wilkins Company. Baltimore. pp. 71-104. [79] DePaula CA, Abjornson C, Pan Y, Kotha SP, Koike K, Guzelsu N. (2002) Changing the

structurally effective mineral content of bone with in vitro fluoride treatment. J.Biomech. 35:355-361.

[80] Maloney SK, Dawson TJ. (1993) Sexual dimorphism in basal metabolism and body

temperature of a large bird, the emu. Condor. 95:1034-1037. [81] Davies SJJF. (2002) Ratites and Tinamous. Oxford University Press. New York. [82] Kaneps AJ, Stover SM, Lane NE. (1997) Changes in canine cortical and cancellous bone

mechanical properties following immobilization and remobilization with exercise. Bone. 21: 419-423.

[83] Donahue SW, McGee ME, Harvey KB, Vaughan MR, Robbins CT. (2006) Hibernating bears

as a model for preventing disuse osteoporosis. J.Biomech. 39: 1480-1488. [84] Weinreb M, Patael H, Preisler O, Ben-Shemen S. (1997) Short-term healing kinetics of

cortical and cancellous bone osteopenia induced by unloading during the reloading period in young rats. Virchow.Arch. 431: 449-452.

[85] Trebacz H. (2001) Disuse-induced deterioration of bone strength is not stopped after free

remobilization in young adult rats. J.Biomech. 34: 1631-1636.

177

Chapter 7 Interface

7.1. Introduction

7.2. Experimental Details

7.2.1. Emu bone samples and KOH treatment 7.2.2. Raman spectroscopy data acquisition 7.2.3. Raman spectroscopy data analysis 7.2.4. Atomic force microscopy imaging 7.2.5. Secondary surface roughness measurements 7.2.6. Statistical analysis

7.3. Results

7.3.1. Raman spectroscopy analysis 7.3.2. Atomic force microscopy

7.4. Discussion

7.4.1. Raman spectroscopy 7.4.2. Atomic force microscopy

7.5. Conclusions

7.6. Acknowledgement

7.7. Chapter Summary

7.8. References

178

The following manuscript has been submitted to Calcified Tissues International for review in

May 2010 under, “Bone collagen degradation affects the mineral-matrix interface”. The

complete list of authors includes: C. Wynnyckyj, S. Omelon and M.D. Grynpas. This chapter

is a continuation of the work done in Chapter 3. In this chapter, we use Raman spectroscopy

and atomic force microscopy to investigate the interface alterations in KOH treated emu

tibiae.

7.1. Introduction

Bone is a complex composite material, consisting of two main phases: a dense mineral phase

embedded within a compliant organic matrix. The mechanical properties of bone depend on

the characteristics of the mineral and organic matrix phases as well as the interactions

between the mineral and organic phases [1-3]. The importance of bonding strength between

the mineral and collagen components was shown by studies that investigated the effects of

decreased bone quality due to changes in bonding between the mineral and organic phases

[2-5].

An emu tibia model was developed to study the contribution of the organic component of

bone, which is approximately 90% collagen matrix and 10% noncollagenous proteins, to the

mechanical properties of bone. In this model, the collagen matrix of bone was degraded with

1 M potassium hydroxide (KOH), while maintaining the mineral content unaltered [6]. While

KOH treatment of bone is non physiological, it helps to understand the mechanisms by which

the degradation of the organic component of bone affects bone mechanical properties, as this

strategy leaves the apatite mineral content of bone unaffected. Endocortical treatment

resulted in negligible mass loss (0.5%), collagen loss (0.05%), no differences in geometrical

parameters, but with significant changes in mechanical properties [6]. Specifically, female

and male emu tibiae showed significant decreases in failure stress and increased failure strain

and toughness with increasing KOH treatment time.

We have recently shown using several mineral (qBSE, microhardness, powder XRD) and

collagen (α-chymotrypsin, DSC, gel electrophoresis, polarized light microscopy)

characterization techniques, that endocortical KOH treatment causes in situ collagen

179

degradation without affecting the mineral phase [7]. However, the reasons for the increased

failure strain and toughness values of KOH treated bones are not fully understood. Previous

studies that have affected collagen by removal with sodium hypochlorite showed a more

brittle-like behaviour with decreased toughness values [8,9], an opposing result to our study

[6]. Therefore, we hypothesize that partial debonding between the mineral and organic matrix

may be an additional mechanism together with in situ collagen degradation that is

responsible for the significant observed alterations of mechanical properties. We probed the

effects of KOH on the organic component of bone with Raman spectroscopy and atomic

force microscopy (AFM).

Raman spectroscopy has been used for chemical identification, characterization of molecular

structures, effects of bonding, environment and stress on a sample material [10]. Raman

spectroscopy measures the non-elastic scattering of the monochromatic light due to light-

induced changes in a molecule. The frequency of this scattered light depends on the structure

and composition of the molecular units in the sample being measured [11]. In biological

tissues, Raman spectroscopy probes the molecular and ionic vibrations of mineral species as

well as the many vibrations that arise from the organic matrix. The Raman spectra of bone is

complex, so researchers have mainly focused on specific bands including the phosphate ν1

band at ~960 cm-1, carbonate ν1 band at ~1070 cm-1 and the bands associated with collagen

(amide III at ~1270 cm-1, C-H bending at ~1450 cm-1 and amide I at ~1665 cm-1) for

spectroscopic chemical characterization of the bone sample [12].

The typically strong, Raman phosphate ν1 band is a prominent marker for mineral content in

bone [10]. The amide I, amide III, and C-H bending (methylene) are typical markers for the

protein or organic matrix of bone. Changes in the amide I band are associated with changes

in collagen crosslinks in bone. The bands of amide I and amide III have been shown to be

good indicators of protein conformation [13]. Furthermore, it has been reported that collagen

denaturation is reflected by the shift of amide III line and broadening of the peak in that

region [14]. Carden et al. recently showed that a shifted amide band is indicative of the

presence of collagen that has undergone a transformation from its normal triple-helical state

to a disordered state [12]. Finally, the phosphate ν1/amide I [15] or amide III [16-18] or C-H

180

bending [12] band height ratio have been used to estimate the mineral-to-organic matrix

ratio.

AFM has been used to study the tissue structure in bone [19], in ossified tendon [20] and at

the calcification front of growing bone [21]. AFM uses the measurement of the surface

topography of a sample by tracing the sample surface with a minute mechanical probe (sharp

needle) attached to a softspring (cantilever). When the probe approaches the sample surface,

tiny interaction forces, such as van der Waals and electrostatic forces, occur between the

probe and the sample. The resulting cantilever deflection is recorded by measuring the

displacement of a laser beam reflected from the back side of the cantilever [22]. A recent

study used AFM to investigate the effect of sodium floride (NaF) on trabecular bone fracture

surfaces in situ and in a time-lapsed fashion [23]. NaF exposure has been shown to influence

the interfacial bonding of the organic components of bone [24]. AFM imaging showed that in

vitro treatment of bone fracture surfaces with highly concentrated NaF solutions resulted in

observable mineral detachment from the underlying collagen fibrils and extraction of more

bone proteins than sodium chloride exposure [23]. The authors suggested that NaF exposure,

which did not significantly reduce bone mass, interfered with the noncollagenous proteins

involved in bonding between the collagen-mineral interface, resulting in weakening of the

organic-mineral interface that was observed in previous NaF exposure studies [24].

The goal of using Raman spectroscopy and AFM to analyze KOH affected bone tissue in this

study was to understand the effect of KOH treatment on the bone mineral-collagen interface.

The major mineral and matrix bands in the spectrum of untreated and KOH treated bone

samples were analyzed using Raman spectroscopy. AFM surface topology data supported by

surface roughness measurements using profilometry was used to investigate changes in bone

surface following KOH treatment. The insights gained were used to explain the significant

changes in the bone mechanical properties from Chapter 3.

181

7.2. Experimental Details

7.2.1. Emu bone samples and KOH treatment

The emu bone tibiae samples from Chapter 3 – Initial Study, were used in this study. These

bones were endocortically treated with 1 M potassium hydroxide (KOH), followed by

mechanical testing in three-point bending [6]. Bones were divided into two groups of female

and male tibiae, with ten samples each allocated to 0-day (untreated) or 14-day KOH

treatments. An additional ten female and ten male bones were filled with PBS instead of

KOH for 14 days to act as controls.

Briefly, skeletally mature (3-5 years of age) whole emu tibiae were separated from the

femora and tarsometatarsi with a scalpel. The shaft of each tibia was then isolated by using a

circular saw to remove the ends (15% of the total bone length from the proximal end and

10% from the distal end). The marrow and trabecular bone from the diaphysis of the tibiae

were removed by drilling longitudinally through the bone shaft, after which the medullary

canal was flushed with tap water [6]. The emu tibiae were filled with 1 M KOH and the ends

were sealed with polymethylmethacrylate (PMMA, SR Ivolen Liquid and Powder, Ivoclar

Vivadent, Mississauga, ON). The bones were held horizontally over a collection basin and

kept moist with a 0.9% saline solution drip. The bones were rotated 180 ° around the axial

axis every 12 hours during the KOH treatment. After the desired treatment period, the KOH

solution was reclaimed, its volume measured and the endocortical surfaces of the bones

rinsed in running tap water for one hour.

7.2.2. Raman spectroscopy data acquisition

Ten mm thick cross sections were cut from untreated and 14-day KOH treated female and

male bones. From these samples, transcortical sections were cut (~ 5 mm x 5 mm x 5mm)

under constant irrigation. These samples were assumed to be representative of the whole

cross-sectional area of each sample. The exposed endocortical surface of the cut sample was

placed under a Hiroba Jobin Yvon LabRam Raman microscope (integrated Raman system)

for spectral scanning over the range 400-1800 cm-1 using a 532 nm laser with a 100X

microscope objective and a laser power of ≤0.5 mW at the sample surface. This wavenumber

range covers the signature bands of mineral and collagen phases [25]. For a given sample, the

182

spectrum was acquired with the LabSpec software (v4.18-05, Jobin Yvon, France) at a

distance of 100 µm from the edges. The 100X objective provided a laser spot size of

approximately 2 µm in diameter. The measured spectra consisted of five accumulations with

an integration time of 25 seconds each and a scan step of 1 cm-1. These settings resulted in

signal to noise ratios ranging from 5:1 (weak amide bands) to 50:1 (prominent phosphate

band). Samples were kept hydrated in between consecutive measurements using a saline-

soaked brush.

7.2.3. Raman spectroscopy data analysis

The sample background fluorescence was subtracted and spectra were filtered using the

LabSpec software (Jobin Yvon, France). Data analysis was performed using MatLab 6.5

(MathWorks Inc., MA) to determine the peak heights and full width at half maximum height

(FWHM) values of the phosphate, carbonate, amide III, methylene (C-H bending) and amide

I bands.

The analysis was confined to the Raman scattering region between 800 and 1800 cm-1. The

average of measurements at five locations was taken to determine the mean mineralization,

the mean carbonate substitution and the mean crystallinity, as shown in Figure 7.1 [15]. The

degree of mineralization was determined by the mineral to matrix ratio. Recent studies have

found that the intensity of the amide I band is affected by the orientation of the collagen

fibrils relative to the incident Raman laser beam [16,17]. Therefore, the ratio of

phosphate/amide I may not be sensitive enough to detect differences in tissue composition.

Therefore, the degree of mineralization was calculated by dividing the peak intensities of the

ν1 phosphate band (at ~960 cm1) with each of the amide III, C-H bending or amide I bands

(at ~1270, 1450, 1665 cm-1), respectively. An increasing mineral:matrix ratio indicates a

more mineralized collagen matrix [15]. Two types of carbonate substitutions have been

described for bone: type A (OH- substituted by CO32-) and type B (PO4

3- substituted by

CO32-) [15]. Raman spectroscopy is not able to detect type A carbonation due to peak overlap

[26]. The extent of type B carbonate substitution was quantified by dividing the intensity of

the phosphate symmetric stretch (ν1) band (at ~960 cm-1) with the type B carbonate

symmetric stretch band (at ~1072 cm-1). An increase in carbonate substitution indicates more

183

mature bone composition [15]. Mineral crystallinity (a combination of crystal size/strain) was

calculated as the inverse of the width of the phosphate symmetric stretch band at half the

maximum intensity value (FWHM). An increasing FWHM indicates a disordered crystal

lattice and hence, lower crystallinity [15]. Furthermore, the average FWHM and Raman shift

values (wavenumber) at the phosphate, carbonate, amide III, C-H bending (methylene) and

amide I were determined (Figure 7.1). Shifts of the collagen bands have been shown to

indicate collagen that has undergone a transformation from its normal tripe-helical state to a

disordered state [12].

Figure 7.1: Typical Raman spectra of bone showing the calculation of degree of mineralization, carbonate substitution and crystallinity (adapted from [15]).

7.2.4. Atomic force microscopy imaging

Following Raman spectroscopy, the same samples were subsequently used for atomic force

microscopy (AFM). Samples were glued onto small discs using epoxy resin (5 min epoxy,

Lepage’s Limited, Brampton, ON) with the endocortical surface face-up for AFM analysis.

Tapping-mode AFM imaging in air was performed with PPP-NCH-50 tips (manufacturer

184

specifications: length=125μm, k=42 N/m; Nanosensors, Neuchalet, Switzerland) using a

Digital Instruments Bioscope™ scanning probe microscope equipped with a Nanoscope IIIa

controller and a dual-range J-scanner with a maximum possible scan size of 90 μm x 90 μm

(Veeco instruments, Santa Barbara, CA). All AFM images were acquired as 16-bit, 512 x

512 pixel images in tapping mode under amplitude feedback, at a typical line scan rate of 1.2

Hz. All AFM images were analyzed with Digital Instruments Nanoscope software (version

5.30r3). Height deviations from the x-y plane, which gives a statistical roughness (r, nm2)

histogram with Gaussian distribution, were measured. Images shown are representative of

features observed from all groups.

7.2.5. Secondary surface roughness measurements

Histograms of the grey level distribution were created for each sample, with decreasing

brightness representing increasing distance from the AFM cantilever tip. A Quips program

on a Quantimet 500 IW system sets up a series of bins based on the intensity of the pixels

and calculates the number of pixels corresponding to each bin. This information is used to

create a histogram of each image. From the histograms, the grey level of the histogram peak

was determined as well as the width at the half peak height. The grey level was used to

represent the overall height profile (dark grey represents the sample surface whereas light

grey corresponds to increased height from the surface) and the width at the half peak height

relates to the heterogeneity of the surface (height distribution). To validate the surface

roughness values determined from the AFM software (Bioscope™) and histograms, samples

were assessed with a Veeco Surface Roughness Wyko non-contact optical profiler (AN505-

3-0902, Veeco Instruments Inc., Tucson, AZ). Profiler conditions were set for vertical

scanning interferometry (VSI) mode, 52X magnification. Five areas were measured per

sample. Surface roughness profiles were generated and analyzed using the accompanying

software (WYKO Vision 32 for NT-2000; version 2.2.10). For each sample, the average

roughness (Ra, μm) was determined, which represents the arithmetic average of all deviations

from the mean center line of the roughness profile. This measurement is the most commonly

reported parameter in studies on surface roughness of bone implant materials [27].

185

7.2.6. Statistical analysis

Statistical analysis was performed using SPSS (SPSS 16.0 for Windows, Chicago, IL)

statistical analysis software. Tests for normality and equality of variances were initially

performed to determine whether parametric or non-parametric t-tests should be used. Two-

way analysis of variance (ANOVA, general linear model) was performed to examine the

effects of sex and KOH treatment on all measured parameters in the emu model and multiple

comparisons (Fisher’s Least Significant Difference (LSD)) were made to detect significant

differences between the groups. All data are presented as mean ± standard error of the mean.

A confidence level of 95% (p=0.05) was considered statistically significant.

7.3. Results

All parameters measured for the control group (14-day filled with saline) were similar to

those of the untreated (0-day) groups (Appendix A) for both female and male emu bones.

7.3.1. Raman spectroscopy analysis

The typical Raman spectra profiles of male (Figure 7.2) and female (Figure 7.3) untreated

(solid lines) and 14-day KOH treated groups (dashed lines) have the same peak positions and

relative intensity, with minor differences. The prominent bands are labeled. The ν1 phosphate

band at ~965 cm-1 is the strongest marker for bone mineral. The band at ~1072 cm-1 indicates

type B carbonate substitution in the bone (carbonate substituting for phosphate in the apatite

lattice). The broad bands in the high frequency region are amide III (~1265 cm-1), the C-H

bending (~1457 cm-1) and amide I (~1665 cm-1). The amide I and amide III peaks are mainly

associated with the presence of collagen while the C-H bending band is present in collagen

and noncollagenous proteins [28,29]. These assignments are in good agreement with those

described previously in the literature [10-12,15].

186

Figure 7.2: Typical Raman spectra scans taken from male untreated (solid line) and 14-day KOH treated (dashed line) samples. No differences in intensities and position of the mineral and collagen peaks between treatment groups.

Figure 7.3: Typical Raman spectra scans taken from female untreated (solid line) and 14-day KOH treated (dashed line) samples. No differences in intensities and position of the mineral and collagen peaks between treatment groups.

187

The resulting Raman spectra showed little differences amongst the groups. The intensity of

each labeled peak and the FWHM of the phosphate peak were used to calculate the degree of

mineralization, type-B carbonate substitution and crystallinity, as shown in Figure 7.1. The

results are presented in Table 7.1. There were no differences observed between the different

groups (untreated vs. 14-day KOH treated), or between sexes, for any of the listed Raman

spectra parameters.

Table 7.1: Average female and male emu tibiae Raman spectroscopy parameters. KOH treatment time in days

Parameter

Sex Untreated 14 days

Female 9.17 ± 0.55 9.02 ± 0.86

Mineralization (phosphate/amide I) Male 10.18 ± 0.91 8.75 ± 0.53

Female 18.47 ± 0.71 17.11 ± 1.27

Mineralization (phosphate/amide III) Male 18.39 ± 1.48 19.22 ± 3.02

Female 11.20 ± 0.70 12.63 ± 0.97

Mineralization (phosphate/C-H bending) Male 12.25 ± 1.19 11.29 ± 1.90

Female 4.76 ± 0.14 4.38 ± 0.22 Carbonate Substitution (phosphate/carbonate)

Male 4.56 ± 0.29 4.35 ± 0.21

Female 0.057 ± 0.001 0.058 ± 0.002

Mineral Crystallinity (1/phosphate FWHM) Male 0.056 ± 0.001 0.057 ± 0.002

Finally, the FWHM versus Raman shift values (wavenumber) at the phosphate, carbonate,

amide III, C-H bending and amide I bands were calculated and plotted to determine any

changes in peak position and band width of the mineral (Figure 7.4) and collagen bands

(Figure 7.5). In both figures, untreated groups are filled circles whereas 14-day KOH

treatment groups are shown as hollow circles (female samples are represented in pink, male

samples in blue). KOH treatment showed no effect on peak position or FWHM for the

phosphate and carbonate bands (mineral) (Figure 7.4) for both sexes. In terms of the collagen

bands, 14-day KOH treatment increased the peak position of the amide III band from 1261

cm-1 ± 2 to 1267 cm-1 ± 2 for female emu tibiae (p=0.041) and from 1263 cm-1 ± 2 to 1270

cm-1 ± 2 for male emu tibiae (p=0.034) while simultaneously, significantly increasing the

188

FWHM of the amide III band from 40 cm-1 ± 2 to 46 cm-1 ± 2 for male emu tibiae (p=0.027)

only (Figure 7.5). The increase in FWHM for females from 41 cm-1 ± 1 to 44 cm-1 ± 2 was a

trend (p=0.076). There were no differences in peak position and FWHM of the collagen

bands between sexes. 14-day KOH treatment had no effect on the C-H bending or amide I

peak position or FWHM (Figure 7.5).

Figure 7.4: Changes in Raman peak widths (FWHM = full width at half maximum) and Raman peak positions of (a) phosphate band and (b) carbonate band for female (pink) and male (blue), untreated (filled circles) and 14-day KOH treated (hollow circles) samples. Error bars represent standard error from the ten samples allocated to each group. No differences were observed between treatment groups as well as between sexes.

189

Figure 7.5: Changes in Raman peak widths (FWHM = full width at half maximum) and Raman peak positions of (a) amide III band, (b) C-H bending band and (c) amide I band for female (pink) and male (blue), untreated (filled circles) and 14-day KOH treated (hollow circles) samples . Error bars represent standard error from the ten samples allocated to each group. 14-day KOH treatment significantly increased the peak position of the amide III band from 1261 cm-1 ± 2 to 1267 cm-1 ± 2 for female emu tibiae and from 1263 cm-1 ± 2 to 1270 cm-1 ± 2 for male emu tibiae while simultaneously increasing the FWHM of the amide III band from 41 cm-1 ± 1 to 44 cm-1 ± 2 for female emu tibiae and from 40 cm-1 ± 2 to 46 cm-1 ± 2 for male emu tibiae. There were no differences between sexes. 14-day KOH treatment had no effect on the C-H bending or amide I position or FWHM.

190

7.3.2. Atomic force microscopy

Figure 7.6 shows AFM images of male, untreated, endocortical samples in Figure 7.6 (a) and

(b) and 14-day KOH treated samples in Figure 7.6 (c) and (d). The bright pixels in the

images correspond to higher points in the topography whereas the darker pixels represent

lower points. The surface of the untreated male sample is covered with rounded bumps,

giving a granular texture. The 67 nm banding pattern characteristic of collagen Type I is not

discernible due to the overlying mineral. The diameter of the spherical subparticles ranges

from 200 nm to 1.25 μm, with the larger particles appearing to be clusters of two or three

smaller particles. The surface morphology of untreated and 14-day KOH treated samples

appears different. Densely packed agglomerated spheroidal particles are present with a

‘swollen’ appearance in the high resolution image (Figure 7.6 (d)) of the 14-day KOH treated

male bone sample.

191

Figure 7.6: A 20 x 20 μm2 AFM tapping mode image of the endocortical surface of male cortical bone, showing (a) densely packed agglomerated spheroidal particles and (b) a higher resolution image from (a), of untreated male samples. Particles appear to be homogeneously closely packed to each other. Figure (c) represents a 14-day KOH treated image and (d) a higher resolution image from (c). The densely packed agglomerated spheroidal particles are still present but appear to be separated into clusters and ‘swollen’.

Figure 7.7 shows AFM images of female untreated endocortical samples (Figure 7.7 (a) and

(b)) and 14-day KOH treated samples (Figure 7.7 (c) and (d)). Untreated female samples had

irregular features compared to the homogeneous surface in male samples and further

magnification could not identify a consistent pattern of spatial structural organization. As a

result of imaging the endocortical surface of the bone samples and without typical bone

processing for AFM (demineralization), these particles are believed to be collagen fibrils

192

coated with mineral. The diameter of the spherical subparticles ranges from 200 nm to 1.25

μm, with the larger particles appearing to be clusters of numerous smaller particles. There

appears to be a greater amount of large particle clusters in the female samples compared to

the male samples. Similarly to the male samples, densely packed agglomerated spheroidal

particles are present with a ‘swollen’ appearance in the high resolution image (Figure 7.7 (d))

of the 14-day KOH treated female bone sample.

Figure 7.7: A 20 x 20 μm2 AFM tapping mode image of the endocortical surface of female cortical bone, showing (a) densely packed agglomerated spheroidal particles in clusters and (b) a higher-resolution image from (a), of untreated female samples. Particles appear to have an irregular surface. Figure (c) represents a 14-day KOH treated image and (d) a higher resolution image from (c). The densely packed agglomerated spheroidal particles are still present but appear to be separated into clusters and ‘swollen’.

193

To quantitatively validate these observations, average roughness values of the samples were

measured using an optical profiler (Table 7.2). The 14-day KOH treated female and male

samples had significantly higher average roughness values compared to untreated surfaces

(p<0.001 for both sexes). Furthermore, the average roughness value of female surfaces was

greater compared to the average roughness value of male surfaces for both untreated

(p=0.066) and 14-day KOH treated groups (p=0.041) (Table 7.2).

Analysis of the histograms created from the grey level distribution from each sample

confirmed these results (Table 7.2). There were no differences between sexes or between

treatment groups in the average grey level (representation of overall surface height profile).

Conversely, the width at half the peak height (representation of the heterogeneity of the

surface) of these histograms increased after 14-day KOH treatment for both female (p=0.062)

and male (p=0.056) emu tibiae, indicating a greater distribution of profile heights.

Furthermore, the width at half the peak height of female samples had a trend towards larger

values compared to male samples for both untreated (p=0.079) and 14-day KOH treated

groups (p=0.057).

Height deviations from the x-y plane of the AFM images (Bioscope™), showed similar

results (Table 7.2). Male untreated samples presented lower roughness (r = 101 ± 10 nm2)

compared to female untreated samples (r = 124 ± 12 nm2) (p=0.061). Finally, 14-day KOH

treated samples showed higher roughness compared to untreated samples for female samples

(r = 191 ± 11 nm2) and male samples (r = 167 ± 123 nm2) (p<0.001 for both sexes).

194

Table 7.2 Quantitative grey level distribution and surface roughness results for female and male emu tibiae.

KOH treatment time in days Parameter

Sex

Untreated 14 days

Female 77 ± 3 75 ± 4 Peak grey level (pixels)

Male 78 ± 4 76 ± 2

Female 53 ± 2 56 ± 2a

Width at half peak height (pixels) Male 50 ± 1A 53 ± 2a,A

Female 124 ± 12 191 ± 11b

Surface Roughness (Bioscope™: r, nm2) Male 101 ± 10A 167 ± 13b,A

Female 1.7 ± 0.2 2.8 ± 0.2b

Surface Roughness (Optical profiler: Ra, μm) Male 1.4 ± 0.1A 2.3 ± 0.2b,B

a0.01 ≤ p ≤ 0.05 versus untreated group; b p < 0.05 versus untreated group. A0.01 ≤ p ≤ 0.05 versus female; Bp< 0.05 versus female.

7.4. Discussion

In our previous study [6], KOH treatment was shown to significantly compromise the

mechanical properties of bone with minimal amount of collagen loss (0.05%), mineral

content loss (0.5%) and no significant changes in geometry. We determined that these

observed changes resulted from in situ collagen degradation [7]. We hypothesized that partial

debonding between the mineral and organic matrix may be an additional mechanism,

together with collagen degradation, that is responsible for the significant observed altered

mechanical properties. The reduction in the bone mineral-interface strength observed with

KOH treatment is comparable to adversely affected bone mechanical properties observed

after NaF treatment [23]. Specifically, both KOH and NaF treatment resulted in significant

decreases in failure stress and increased failure strain and toughness with treatment time,

without significant bone loss [6,23]. The purpose of this study was to determine if

endocortical KOH treatment of whole emu tibiae affected the mineral-organic matrix

interface by investigating endocortical chemical and surface roughness changes. Endocortical

surfaces (mineral and organic components) of female and male, untreated and 14-day KOH

treated samples were characterized using Raman spectroscopy. Subsequently, topographic

195

profiles of female and male, untreated and 14-day KOH treated bone samples were studied

using AFM and profilometry.

7.4.1. Raman spectroscopy

For all groups, the position of the prominent peaks found in emu bone was comparable to

those found in other studies [15-17,26,30]. The Raman signal is dependent upon a

combination of the composition and the structure of the sample [16,31]. There were no

differences between sexes or with KOH treatment of the degree of mineralization (calculated

as the ratio of peak intensities of the phosphate band to either of the amide III, C-H bending

or amide I bands). The three calculations were performed, as recent studies suggested that the

phosphate/amide I ratio may be sensitive to the orientation of the incident light [16]. It was

also shown that the amide III band would be less susceptible to orientation effects [16].

Nevertheless, no differences in degree of mineralization were observed in this study.

Furthermore, the degree of carbonate substitution and crystallinity were also unaffected by

KOH treatment and there were no differences between sexes. In addition, the position,

intensity and shape of the Raman bands assigned to the mineral components of bone

remained unchanged between untreated and 14-day KOH treated samples for both sexes.

These results confirm previous observations that KOH treatment of bone tissue does not

affect bone mineral [6,7].

Shifts of the collagen bands and broader peaks have been shown to indicate collagen that has

undergone a transformation from its normal triple-helical state to a disordered state [12,30].

Increased shifts in amide I and amide III bands are typical features of collagen that has lost

its structure through crosslink rupture (amide I) [32] or collagen denaturation (amide III)

[33]. Changes in peak shape were observed for the amide III band, particularly in the spectra

from the 14-day KOH treated samples. Specifically, the height of the amide III peak

decreased and increased in FWHM and the position of the peak was shifted to a higher

wavenumber after KOH treatment for both sexes, suggesting a change in the secondary

structure of the collagen. It is not clear if the amide III peak changes are due to denatured

collagen, altered noncollagenous proteins, or a combination of both. KOH treatment showed

no effect in peak position and band width of the C-H bending or amide I bands.

196

The amide bands of the Raman spectral pattern are mostly composed of peptide vibrations

[34]. Since collagen consists of amino acids linked by peptide bonds, the vibrational modes

most sensitive to secondary structure (the backbone) are those of the amino class [35]. The

amide I band is mostly due to the C=O stretching vibration of the peptide groups with some

C-N stretching and N-H bending and has been shown to be sensitive to changes in protein

structure [35-37]. Specifically, the amide I band responds to changes in secondary structure

[38] and a shape change of the amide I band has been shown to indicate the rupture of

crosslinks [39]. Based on these observations, it may be inferred that the mechanism of KOH

on emu tibiae did not rupture collagen crosslinks as there were no significant changes in

amide I band peak position or shape after 14-day KOH treatment. However, modifications to

these crosslinks cannot be ruled out. The C-H bending stems from side-chains thus, it is less

susceptible to changes in the collagen structure [34]. The amide III mode is associated with

the C-N vibration [37] and has also been used to estimate conformation characteristics of the

protein backbone [35,37,40]. Wang et al. determined molecular changes in collagen of rat tail

tendon under strain using Raman spectroscopy and showed a decrease in FWHM in several

key collagen bands, indicating that the structure had become more ordered [35]. Therefore,

an increase in FWHM indicates an increase in bending of the molecular kinks, resulting in an

increase in disorder of the protein structure. In a previous study [7], we have shown that

endocortical KOH treatment causes in situ collagen degradation without affecting the mineral

phase. These resulting degradation fragments as well as the high pH of 14 of the 1 M KOH

solution would in turn cause an increased susceptibility to conformational changes in the

protein structure, which may also result in mineral-matrix interface debonding. Therefore,

AFM was utilized to investigate the interface.

7.4.2. Atomic force microscopy

AFM images of untreated and 14-day KOH treated samples revealed a large number of

spherical subparticles packed closely together on the surface. Larger particles in the images

appear to be clusters of two or three smaller particles. Similar spherical particles have been

observed using AFM of bovine bone [41,42]. Specifically, Sasaki et al. treated bovine bone

with collagenase to eliminate the collagen fibers from the bone surface, leaving the mineral

phase unchanged [42]. The resulting AFM images showed spherical particles that the authors

197

described as the mineral component of bone. Conversely, treatment of bovine bone with

EDTA, which is known to eliminate the mineral component in bone, causes the spherical

particles to disappear, revealing thread-like objects with a banding pattern [42]. The authors

attributed the thread-like objects to the collagen fibrils of bone. Clusters of mineralized

spherical structures were seen in AFM images of dentine treated with phosphoric acid. The

residual clusters were a result of incomplete mineral dissolution [43]. Phosphoric-acid treated

samples exhibited irregular features and further magnification could not identify a consistent

pattern of spatial structural organization. The presence of the 67 nm periodicity of gap zones

was not observed, but would become apparent and increasingly pronounced if the samples

were to be demineralized, as shown by Balooch et al. [44]. The lack of visible collagen fibrils

in the AFM images of this study further confirms the fact that KOH treatment does not affect

the bone mineral and collagen fibrils are not normally visible on the surface of bone, without

appropriate treatment to remove the mineral component.

In our study, significant differences in surface topography were observed by AFM scanning

between sexes in untreated bone samples as well as between 14-day KOH treated and

untreated samples. For the untreated samples, the spherical particles appear to be connected

to each other relatively homogeneously in the male bone samples, whereas a higher number

of particle clusters appear to be present in the female bone samples. Similar differences were

observed when comparing the 14-day KOH treated images to the untreated images for both

sexes. The 14-day KOH treated samples appear to have increasing spacing between adjacent

particle clusters. These qualitative measures are validated with the width at half the peak

height of the grey level distribution histograms: for both female and male 14-day KOH

treated samples, the width at half the peak height increased, indicating an increase in height

distribution. Therefore, untreated samples appear to have a more homogeneous surface

compared to the more irregular, uneven 14-day KOH treated samples.

The irregular surface profile is likely due to the endocortical KOH treatment. Furthermore,

the measured surface roughness values from the optical profiler as well as the height

deviations from the x-y plane of the AFM images show similar results. The topographical

differences observed in these AFM images may be due to the KOH penetrating into the bone,

198

altering the collagen structure and possibly the noncollagenous protein structures or the

bonding between the noncollagenous proteins and the mineral or the collagen. Altering the

collagen conformation may change the surface interactions with the mineral component of

bone, which in turn, may alter the bone mineral-matrix interface. Should the noncollagenous

proteins be altered, the role they play in providing a bond between the charged bone mineral

surface and the surrounding collagen may be compromised. This interfacial debonding would

result in reduced mechanical properties.

Closer examination of high magnification AFM images of the 14-day KOH treated samples

revealed a ‘swollen’ appearance. This swollen appearance may be due to an altered mineral-

matrix interface. The bone mineral-collagen interface proteins hold the mineral in close

proximity to the collagen. If KOH treatment denatures or degrades these proteins or

interfaces, the collagen may separate from the mineral surface, causing it to look ‘bigger’ or

‘swollen’. Recently, Xu et al. analyzed the structure of bovine tibiae before and after

demineralization using AFM [45]. The authors observed that demineralization dissolved the

bone mineral phase and revealed a layer of small spherical particles deposited on the surface

of collagen bundles, preventing mineral crystals of being in direct contact with collagen

bundles. These particles were identified as lipids since organic solvents were capable of

dissolving them [45]. If a layer of lipids is present in between the mineral and collagen, KOH

treatment may be interacting with these lipids, causing a modification in the mineral-collagen

interface, resulting in a ‘swollen’ appearance. A ‘swollen’ appearance may also be due to an

increase in bone hydration after KOH treatment. Silva and Ulrich observed increased bone

hydration following NaF treatment [46]. Increased water content may be possible with the

loss of the strong links between the mineral, the collagen and the noncollagenous proteins.

Therefore, increased bone hydration might be observed as a ‘swollen’ bone tissue.

Similarly, AFM studies of boiled bone samples revealed a soft appearance compared to

untreated bone samples and the authors suggested that even though the organic phase was

still present, the cohesion of the organic matrix was reduced [8]. This modified bonding

would explain the increased flexible behaviour of boiled bone: the organic component is still

present for energy dissipation functions however, the organic matrix no longer binds the

199

inorganic component tightly together with the compliant organic matrix [8]. This may be

analogous to the mechanism of action of KOH. KOH may have affected the binding between

the collagen and mineral as well as the interactions within the collagen, leading to an increase

in toughness but decrease in strength. It has been postulated that the bonding between bone

apatite and collagen fibrils involves sharing of hydroxyl groups since bone apatite lacks

hydroxyl groups, whereas collagen fibrils are abundant in hydroxyl groups [47]. This shared

bonding would allow for rapid breaking or reformation of the mineral-collagen bond and

variations in pH would be able to accomplish such bonding and debonding [47]. The high pH

of 14 of the 1 M KOH solution may have taken place within the emu bone samples and may

have reduced interfacial bonding [24]. Alternatively, KOH may be altering the behaviour of

noncollagenous proteins or peptides [47]. Various researchers have shown noncollagenous

proteins to be either mineral-bound or collagen-bound [48,49], suggesting that the mineral-

collagen interface is formed by a complex noncollagenous protein network, which involves

the binding of noncollagenous proteins to both the mineral and the collagen [50]. The

endocortical KOH treatment may be affecting the binding of the noncollagenous proteins

however, it still remains unclear which proteins may be involved and further research is

needed.

7.5. Conclusions

The results of this study support our hypothesis that KOH treatment causes alterations to the

collagen-mineral interface. While Raman analysis of 14-day KOH treated bones showed

changes only to the amide III bands, AFM images provide visual proof of a swollen

mineralized surface, which we believe to be due to KOH treatment modifying the underlying

interactions between the mineral and collagen, as well as modifying the collagen matrix. The

effect of KOH on the collagen and possibly the noncollagenous proteins involved in bonding

between the collagen and the mineral phases of bone may be responsible for the altered

mechanical properties previously observed in endocortically, KOH treated emu bones.

7.6. Acknowledgement

The author acknowledges assistance from Gary Mo, who performed the AFM imaging and

measurements.

200

7.7. Chapter Summary

The endocortical surface of female and male untreated and 14-day KOH treated emu tibiae

were assessed using Raman spectroscopy and atomic force microscopy (AFM) to identify

chemical and morphological changes. The Raman spectra of 14-day KOH treated bone

samples showed a singular increase in amide III band peak position and width for both sexes.

This observation supports the theory that KOH treatment causes disorder in the organic

component of bone. The disorganization of collagen and noncollagenous proteins would

adversely affect collagen matrix integrity and the mineral-organic interface. AFM images

revealed a swollen appearance in 14-day KOH treated samples compared to untreated

samples. These results suggest that degradation of the organic matrix of bone weakens the

mineral-collagen interface.

201

7.8. References [1] Li X, Agrawal M, Wang X. (2003) Age dependence of in situ thermostability of collagen in

human bone. Calcif.Tissue.Int. 72:513-518. [2] Bundy KJ. (1985) Determination of mineral-organic bonding effectiveness in bone--

theoretical considerations. Ann.Biomed.Eng. 13:119-35. [3] Walsh WR, Ohno M, Guzelsu N. (1994) Bone composite behaviour: effects of mineral-

organic bonding. J.Mater.Sci.Mater.Med. 5:72-9. [4] Kotha SP, Guzelsu N. (2000) The effects of interphase and bonding on the elastic modulus of

bone: changes with age-related osteoporosis. Med.Eng.Phys. 22:575-85. [5] Gupta HS, Fratzl P, Kerschnitzki M, Benecke G, Wagermaier W, Kirchner HO. (2007)

Evidence for an elementary process in bone plasticity with an activation enthalpy of 1 eV. J.R.Soc.Interface. 4:277-82.

[6] Wynnyckyj C, Omelon S, Savage K, Damani M, Chachra D, Grynpas MD. (2009) A new

tool to assess the mechanical properties of bone due to collagen degradation. Bone. 44:840-848.

[7] Wynnyckyj C, Willet TL, Omelon S, Kyle K, Goldberg H, Grynpas MD. (2010) Mechanism

of bone collagen degradation due to KOH treatment. Biochim.Biophys.Acta. In revision. [8] Fantner GE, Birkedal H, Kindt JH, Hassenkam T, Weaver JC, Cutroni JA, Bosma BL,

Bawazer L, Finch MM, Cidade GA, Morse DE, Stucky GD, Hansma PK. (2004) Influence of the degradation of the organic matrix on the microscopic fracture behavior of trabecular bone. Bone. 35:1013-1022.

[9] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

[10] Morris MD, Finney WF. (2004) Recent developments in Raman and infrared spectroscopy

and imaging of bone tissue. Spectroscopy. 18:155-159. [11] Carden A, Morris MD. (2000) Application of vibrational spectroscopy to the study of

mineralized tissues (review). J.Biomed.Opt. 5:259-68. [12] Carden A, Rajachar RM, Morris MD, Kohn DH. (2003) Ultrastructural changes

accompanying the mechanical deformation of bone tissue: a raman imaging study. Calcif.Tissue.Int. 72:166-175.

[13] Kubisz L, Polomska M. (2007) FT NIR Raman studies on γ-irradiated bone.

Spectrochim.Acta.A.Mol.Biomol.Spectrosc. 66:616-625. [14] Bandekar J. (1992) Amide modes and protein conformation. Biochim.Biophys.Acta.

1120:123-143.

202

[15] Akkus O, Addar F, Schaffler MB. (2004) Age-related changes in physiochemical properties of mineral crystals are related to impaired mechanical function of cortical bone. Bone. 34:443-453.

[16] Kazanci M, Roschger P, Paschalis EP, Klaushofer K, Fratzl P. (2006) Bone osteonal tissues

by Raman spectral mapping: orientation-composition. J.Structur.Biol. 156:489-496. [17] Kazanci M, Wagner HD, Manjubala NI, Gupta HS, Paschalis EP, Roschger P, Fratzl P.

(2007) Raman imaging of two orthogonal planes within cortical bone. Bone. 41:456-461. [18] Nyman JS, Masui P, Patil CA, O’Quinn EC, Edwards JR, Guelcher S, Nicollela DP,

Mahadevan-Jansen A. (2009) Detecting compositional differences by raman spectroscopy in non-fixed and embedded human cortical bone. Proceedings of the 55th annual meeting of the Orthopaedic Research Society.

[19] Reilly GC, Knapp HF, Stemmer A, Niederer P, Knothe Tate ML. (2001) Investigation of the

morphology of the lacunocanalicular system of cortical bone using atomic force microscopy. Ann.Biomed.Eng. 29:1074-1081.

[20] Lees S, Prostak KS, Ingle VK, Kjoller K. (1994) The loci of mineral in turkey leg tendon as

seen by atomic force microscope and electron microscopy. Calcif.Tissue.Int. 55:180-189. [21] Wiesmann HP, Chi L, Stratmann U, Plate U, Fuchs H, Joos U, Hohling HJ. (1998) Sutural

mineralization of rat calvaria characterized by atomic force microscopy and transmission electron microscopy. Cell.Tissue.Res. 294:93-97.

[22] Thalhammer S, Heckl WM, Zink A, Nerlich AG. (2001) Atomic force microscopy for high

resolution imaging of collagen fibrils – A new technique to investigate collagen structure in historic bone tissues. J.Arch.Sci. 28:1061-1068.

[23] Kindt JH, Thurner PJ, Lauer ME, Bosma BL, Schitter G, Fantner GE, Izumi M, Weaver JC,

Morse DE, Hansma PK. (2007) In situ observation of fluoride-ion-induced hydroxyapatite-collagen detachment on bone fracture surfaces by atomic force microscopy. Nanotech. 18:1-8.

[24] Walsh WR, Labrador DP, Kim HD, Guzelsu N. (1994) The effect of in vitro fluoride ion

treatment on the ultrasonic properties of cortical bone. Ann.Biomed.Eng. 22:404-15. [25] Yeni YN, Yerramshetty J, Akkus O, Pechey C, Les CM. (2006) Effect of fixation and

embedding on raman spectroscopic analysis of bone tissue. Calcif.Tissue.Int. 78:363-371. [26] Shen J, Fan J, Yang A, Shen G, Hu JM. (2010) A longitudinal raman microspectroscopic

study of osteoporosis induced by spinal cord injury. Osteoporos.Int. 21:81-87. [27] Szmukler-Moncler S, Testori T, Bernard JP. (2004) Etched implants: a comparative surface

analysis of four implant systems. J.Biomed.Mater.Res.B.Appl.Biomater. 69:46-57. [28] Timlin JA, Carden A, Morris MD. (1999a) Chemical microstructure of cortical bone probed

by Raman transects. App.Spect. 53:1429-1435.

203

[29] Timlin JA, Carden A, Morris MD, Bonadio JF, Hoffler II, Kozloff CE, Goldstein KM. (1999b) Spatial distribution of phosphate species in mature and newly generated mammalian bone by hyperspectral Raman imaging. J.Biomed.Opt. 4:28-34.

[30] Freeman JJ, Wopenka B, Silva MJ, Pasteris JD. (2001) Raman spectroscopic detection of

changes in bioapatite in mouse femora as a function of age and in vitro fluoride treatment. Calcif.Tissue.Int. 68:156-162.

[31] Leroy G, Penel G, Leroy N, Bres E. (2002) Human tooth enamel: a Raman polarized

approach. Appl.Spect. 56:1030-1034. [32] Paschalis EP, Verdelis K, Doty SB, Boskey AL, Mendelsohn R, Yamauchi M. (2001)

Spectroscopic characterization of collagen cross-links in bone. J.Bone.Min.Res. 16:1821-1828.

[33] Frushour BG, Koenig JL. (1975) Raman scattering of collagen, gelatin, and elastin.

Biopolym. 14:379-391. [34] Yeni YN, Yerramshetty J, Akkus O, Pechey C, Les CM. (2006) Effect of fixation and

embedding on raman spectroscopic analysis of bone tissue. Calcif.Tissue.Int. 78:363-371. [35] Wang YN, Galiotis C, Bader DL. (2000) Determination of molecular changes in soft tissues

under strain using laser Raman microscopy. J.Biomech. 33:483-486. [36] McCreadie BR, Morris MD, Chen TC, Sudhaker Rao D, Finney WF, Widjaja E, Goldstein

SA. (2006) Bone tissue compositional differences in women with and without osteoporotic fracture. Bone. 39:1190-1195.

[37] Rousseau ME, Lefevre T, Beaulieu L, Asakura T, Pezolet M. (2004) Study of protein

conformation and orientation in silkworm and spider silk fibers using raman microspectroscopy. Biomacro. 5:2247-2257.

[38] Carmejane O, Morris MD, Davis MK, Stixrude L, Tecklenburg M, Rajachar RM, Kohn DH.

(2005) Bone chemical structure response to mechanical stress studied by high pressure raman spectroscopy. Calcif.Tissue.Int. 76:207-213.

[39] Sahar N, Kohn DH, Golcuk K, Morris MD. (2009) Effects of exercise on bone quality as

shown by raman microspectroscopy. In proceedings of the 55th annual meeting of the Orthopaedic Research Society.

[40] Dong R, Yan X, Pang X, Liu S. (2004) Temperature-dependent Raman spectra of collagen

and DNA. Spectrochim.Acta.A.Mol.Biomol.Spectrosc. 60:557-561. [41] Tong W, Glimcher MJ, Katz JL, Kuhn L, Eppell SJ. (2003) Size and shape of mineralites in

young bovine bone measured by atomic force microscopy. Calcif.Tissue.Int. 72:592-598. [42] Sasaki N, Tagami A, Goto T, Taniguchi M, Nakata M, Hikichi K. (2002) Atomic force

microscopic studies on the structure of bovine femoral cortical bone at the collagen fibril-mineral level. J.Mat.Sci.Mat.Med. 13:333-337.

204

[43] Bozec L, de Groot J, Odlyha M, Nicholls B, Horton MA. (2002) Mineralised tissues as nanomaterials: analysis by atomic force microscopy. IEE.Proc.Nanobiotechnol. 152:183-186.

[44] Balooch M, Habelitz S, Kinney JH, Marshall SJ, Marshall GW. (2008) Mechanical properties

of mineralized collagen fibrils as influenced by demineralization. J.Struct.Biol. 162:404-410. [45] Xu S, Yu JJ. (2006) Beneath the minerals, a layer of round lipid particles was identified to

mediate collagen calcification in compact bone formation. Biophys.J. 91:4221-4229. [46] Silva MJ, Ulrich SR. (2000) In vitro sodium fluoride exposure decreases torsional and

bending strength and increases ductility of mouse femora. J.Biomech. 33:231-234. [47] Pasteris JD, Wopenka B, Freeman JJ, Rogers K, Valsami-Jones E, van der Houwen JA, Silva

MJ. (2004) Lack of OH in nanocrystalline apatite as a function of degree of atomic order: implications for bone and biomaterials. Biomater. 25:229-238.

[48] Tye CE, Hunter GK, Goldberg HA. (2005) Identification of type I collagen-binding domain

of bone sialoprotein and characterization of the mechanism of interaction. J.Biol.Chem. 280:13487-13492.

[49] Grynpas MD, Tupy JH, Sodek J. (1995) The distribution of soluble, mineral-bound, and

matrix-bound proteins in osteoporotic and normal bones. Bone. 15:505-513. [50] Raif EM, Harmand MF. (1993) Molecular interface characterization in human bone matrix. I.

Biochemical and IR spectroscopic studies. Biomater. 14:978-984.

205

Chapter 8 Conclusions

206

As outlined in Chapter 1, the work in this thesis was divided into five separate objectives.

This section outlines the overall conclusions that have been drawn based on the culmination

of work performed within the five objectives.

1. Untreated female emu tibiae had significantly higher cross-sectional geometry, mass,

BMD and BMC values as well as ultimate stress and failure stress compared to

untreated male bone samples. However, male emu tibiae had significantly higher

failure strain and post-yield strain compared to female bone samples. The inactivity

or insufficient remobilization period experienced by male emus during incubation (8-

weeks) and the sexual dimorphisms may be the reasons for the observed mechanical

property differences in untreated bones.

2. Endocortical KOH treatment of whole emu tibiae does not affect bone mineral.

3. Endocortical KOH treatment of whole emu tibiae resulted in negligible bone mass

loss (0.5%), collagen loss (0.05%), no changes in geometrical parameters, but the

mechanical properties were significantly affected. Specifically, a significant decrease

in modulus and failure strength as well as a significant increase in failure strain and

toughness was observed with increasing KOH treatment time for both sexes. The

KOH treated bones behaved in a more ductile-like manner.

4. Endocortical KOH treatment of whole emu tibiae causes in situ collagen degradation

rather than removal. This occurs through KOH penetration into the bone, which

causes swelling and a change in the lattice structure, resulting in a less densely

packed and less ordered structure. Male emu tibiae had significantly more degraded

collagen compared to female emu tibiae.

5. Polarized light microscopy revealed differences in collagen organization in KOH

treated bones on the endocortical side compared to the periosteal side for both sexes.

207

6. Endocortical KOH treatment lowered the fatigue resistance of emu tibiae at high

stresses (>60 MPa) only. It was determined that cyclic creep was an important

mechanism in the fast deterioration rate of the KOH treated bones. Collagen

degradation and disorganization caused by KOH treatment is the reason for the

observed altered fatigue behaviour at high stresses only, since collagen is responsible

for the creep behaviour in bone. At low stresses (<60 MPa), the degraded and

disorganized collagen in KOH treated bones act as additional microstructural barriers,

leading to an increased amount of crack deflection and thus, resistance to crack

propagation.

7. DXA and QUS did not detect changes in bone mechanical properties as a result of

collagen degradation as well as fatigue.

8. MRTA detected the changes in bone mechanical properties induced by collagen

degradation as well as fatigue.

9. Fracture surface analysis of KOH treated samples revealed a higher degree of

‘roughness’ as well as the presence of additional toughening mechanisms, indicating

higher resistance to crack propagation and hence, increased toughness.

10. The high pH of the KOH solution causes conformational changes in the protein

structure (as seen from the shape change of the amide III band in the Raman spectra),

as well as degradation fragments, which in turn, may also lead to partial interface

debonding (as seen from the swollen appearance of 14-day KOH treated AFM

images).

208

Chapter 9 Future work

9.1. Introduction

9.2. Future work

9.2.1. Objective 1 - Initial study 9.2.2. Objective 2 - Fatigue 9.2.3. Objective 4 - Collagen degradation 9.2.4. Objective 5 - Interface

9.3. References

209

9.1. Introduction

In this study, we investigated the role of collagen on the mechanical and fatigue properties of

female and male bone, using an emu model. This study increased our knowledge of bone

mechanical properties, mineral-matrix interface, collagen degradation and sex differences on

bone quality at the various hierarchical levels of bone. An understanding of the role of

collagen degradation from the mineral phase may provide new avenues of treatment of bone

fragility. To more extensively investigate the consequences of collagen degradation on bone

mechanical properties, some future work is recommended here.

9.2. Future work

9.2.1. Objective 1 - Initial study

The MRTA has been successfully developed to clinically test the human ulna. Other sites

such as the femur and tibia are of greater clinical concern. The human tibia is the most

common site of stress fractures in athletes and military recruits [1]. Therefore, it would be

advantageous to further develop the MRTA to measure the human tibia and other possible

sites.

Collagen is not quantifiable by current imaging modalities (DXA and QUS). It has been

demonstrated that when the collagen phase is damaged, the fatigue and fracture resistance of

bone are greatly reduced [2-6]. It would be interesting to modify the collagen in whole emu

tibiae using known techniques that affect collagen such as GuHCl (collagen denaturation)

[7], gamma radiation (cleavage of collagen) [8] and ribose incubation (increased crosslink

formation) [9] and determine if CT, DXA, QUS and MRTA can detect these changes.

Similarly, investigations should be undertaken to determine the ability of CT, DXA, QUS

and MRTA to detect induced changes in bone mineral of whole emu tibiae with known

demineralization agents (EDTA/HCl) [7,10].

Three-point bend testing offers the advantage of being a simple test however, it has the

disadvantage of creating high shear stresses in the region around the midsection of the bone

[11]. Conversely, four-point bend testing ensures transverse shear stresses are zero by

producing pure bending between the two upper loading points. Unfortunately, the force at

210

each loading point needs to be equal in order for this to occur, a difficult requirement to

achieve when testing whole bones [11]. As a result, three-point bend testing is the common

testing methodology for measuring the mechanical properties of bones. However, four-point

bend testing of emu tibiae should be investigated.

Various factors can affect the fracture properties of bone including porosity and mineral

content as well as strain rate. It would be interesting to investigate the behaviour of KOH

treated bones at varying strain rates.

9.2.2. Objective 2 - Fatigue

In terms of fatigue testing, with the S/N approach, the measured fatigue lifetime represents

the number of cycles to both initiate and propagate a (dominant) crack to failure. As a result,

the S/N fatigue results are difficult to interpret in terms of mechanisms responsible for

fatigue failure, as the factors affecting crack initiation and subsequent crack growth cannot

readily be differentiated [12]. To analyze fatigue crack propagation, a fracture mechanics

approach needs to be considered where the crack-propagation rate, da/dN, is assessed in

terms of the range in stress-intensity factor, ΔK, defined as the difference between the stress

intensity at the maximum and minimum of the loading cycle [12]. The stress-intensity factor

fully characterizes the local stress and deformation fields in the immediate vicinity of a crack

tip in a linear-elastic solid and thus can be used to correlate to the extent of crack advance.

By using this approach, it is possible to isolate the specific mechanisms responsible for

fatigue crack growth and determine the microstructural or external factors which affect

growth separately from crack initiation. This may be deemed a more important aspect of

fatigue damage as bone is known to possess an inherent population of microcracks which

may minimize the role of the crack initiation stage [13].

The ability to characterize fracture resistance in emu tibiae (untreated and KOH treated) is of

importance to understanding bone fragility. Studies have measured the critical stress intensity

factor, Kc, and the critical strain energy release rate, Gc, in various bone tissues [14,15].

However, a linear elastic fracture mechanics (LEFM) approach is often inadequate to

characterize bone because of its complex microstructure and nonlinear, anisotropic

211

behaviour. This is especially valid when nonlinear processes absorb energy at the crack tip,

resulting in stable crack growth and increasing fracture resistance with increasing crack

length [15]. Such materials are commonly evaluated using the R-curve method, in which

fracture resistance is measured as a function of crack length as the crack propagates in a

stable, rather than catastrophic manner [15]. In order to provide a better estimate of bone’s

resistance to fracture, application of fracture mechanics should be utilized. As such, to study

fracture characteristics of untreated and KOH treated bones, a controlled notch can be

introduced into bone samples.

Microdamage assessment should be attempted using confocal microscopy as this technique

offers better visualization of microdamage than brightfield or fluorescence microscopy. The

small aperture of the confocal microscope improves brightness and spatial resolution by

providing light that does not scatter, hence enhancing the contrast between the stained crack

and the unstained background [16]. An alternative staining agent, such as lead uranyl acetate,

could be used to assess microdamage, as it has been shown to stain for microdamage in bone

[17].

It would also be interesting to investigate the fatigue damage accumulation (loss of stiffness)

and microdamage accumulation at different intervals during fatigue testing of emu tibiae

(untreated and KOH treated) in order to determine at what point during fatigue testing the

majority of the modulus degradation occurs. The samples could be fatigued to a set number

of cycle intervals (such as every 10,000 cycles) and then assessed using CT, DXA, QUS and

MRTA.

9.2.3. Objective 4 - Collagen degradation

Bone is a composite material composed of mineral, collagen, noncollagenous proteins and

solutes. Noncollagenous proteins play various roles in regulating mineralization and

maintaining the strength, stability and integrity of bone [18]. Noncollagenous protein

extraction experiments should be undertaken in order to verify any effect due to KOH

treatment.

212

The collagen molecules are stabilized by intermolecular crosslinks. Crosslinks have been

shown to inhibit intermolecular sliding and are important for enhancement of bone toughness

[19]. The location, identification and quantification of crosslinks in untreated and KOH

treated bones should also be investigated.

9.2.4. Objective 5 - Interface

In terms of investigating the nature of the collagen-mineral interface in KOH treated bone,

solid-state NMR can be used. It has been shown that 13C spins are mostly confined to the

organic matrix while 31P spins are largely confined to the inorganic component and thus, any 13C-31P distances measured, can be related to the organic-inorganic interface [20].

Finally, time-resolved sequence AFM imaging should be performed on emu tibiae samples

during chemical treatment with a highly concentrated KOH solution and afterwards with

EDTA. Untreated samples should be flushed with EDTA only. The EDTA treatment will

show the loss of mineral and may reveal any changes in the underlying collagen structure due

to KOH treatment. Subsequently, these samples should be analyzed using Raman

spectroscopy.

With respect to Raman spectroscopy, amide I profiles contain contributions from β-sheets, β-

strands, β-turns and disordered residues as well as from the predominant α-helical portions

[21]. Therefore, to resolve contributions from α-helix and β-sheet/strand, the Raman amide I

band should be deconvoluted for secondary structure quantification.

213

9.3. References

[1] Bennell K, Matheson G, Meeuwisse W, Brukner P. (1999) Risk factors for stress fractures. Sports.Med. 28:91-122.

[2] Li X, Agrawal M, Wang X. (2003) Age dependence of in situ thermostability of collagen in

human bone. Calcif.Tissue.Int. 72:513-518. [3] Wang X, TeKoppele JM, Athanasiou KA. (2000) Effect of collagen denaturation on the

toughness of bone. Clin.Orthop.Relat.Res. 371:228-239. [4] Wang X, Li X, Bank RA, Agrawal CM. (2002) Effects of collagen unwinding and cleavage

on the mechanical integrity of the collagen network in bone. Calcif.Tissue.Int. 71:186-192. [5] Oxlund H, Barckman M, Ortoft G, Andreassen TT. (1995) Reduced concentrations of

collagen cross-links are associated with reduced strength of bone. Bone. 17:365S-371S. [6] Jepsen KJ, Goldstein SA, Kuhn JL, Schaffler MB, Bonadio J. (1996) Type-I collagen

mutation compromises the post-yield behavior of Mov13 long bone. J.Orthop.Res. 14:493-499.

[7] Grynpas MD, Tupy JH, Sodek J. (1994) The distribution of soluble, mineral-bound and

matrix-bound proteins in osteoporotic and normal bones. Bone. 15:505-513. [8] Akkus O, Belaney RM, Das P. (2005) Free radical scavenging alleviates the biochemical

impairment of gamma radiation sterilized bone tissue. J.Orthop.Res. 23:838-845. [9] Vashishth D, Gibson GJ, Khoury JI, Schaffler MB, Kimura J, Fyhrie DP. (2001) Influence of

nonenzymatic glycation on biomechanical properties of cortical bone. Bone. 28:195-201. [10] Abe K, Hashizume H, Ushiki T. (1992) An EDTA-KOH method to expose bone cells for

scanning electron microscopy. J.Electron.Microsc.(Tokyo). 41:113-115. [11] Turner CH, Burr DB. (2001) Experimental Techniques for Bone Mechanics. In: Cowin SC,

editor. Bone Mechanics Handbook. CRC Press. Florida. pp. 1-35. [12] Kruzic JJ, Ritchie RO. (2008) Fatigue of mineralized tissues: cortical bone and dentin.

J.Mech.Beh.Biomed.Mat. 3-17. [13] Nalla RK, Kruzic JJ, Kinney JH, Ritchie RO. (2005) Aspects of in vitro fatigue in human

cortical bone: time and cycle dependent crack growth. Biomater. 26:2183-2195. [14] Bonfield W. (1987) Advances in the fracture mechanics of cortical bone. J.Biomech.

20:1071-1081. [15] Malik CL, Stover SM, Martin RB, Gibeling JC. (2003) Equine cortical bone exhibits rising

R-curve fracture mechanics. J.Biomech. 36:191-198. [16] O’Brien FJ, Taylor D, Dickson GR, Lee TC. (2000) Visualisation of three-dimensional

microcracks in compact bone. J.Anat. 197:413-420.

214

[17] Schaffler MB, Pitchford WC, Choi K, Riddle JM. (1994) Examination of compact bone microdamage using back-scattered electron microscopy. Bone. 15:483-488.

[18] Gundberg CM. (2003) Matrix proteins. Osteoporos.Int. 14:S37-S42. [19] Knott L, Bailey AJ. (1998) Collagen cross-links in mineralizing tissues: a review of their

chemistry, function, and clinical relevance. Bone. 22:181-187. [20] Jaeger C, Groom NS, Bowe EA, Horner A, Davies ME, Murray RC, Duer MJ. (2005)

Investigation of the nature of the protein-mineral interface in bone by solid-state NMR. Chem.Mater. 17:3059-3061.

[21] Shen J, Fan L, Yang J, Shen AG, Hu JM. (2010) A longitudinal raman microspectroscopic

study of osteoporosis induced by spinal cord injury. Osteoporos.Int. 21:81-87.

215

Appendix A Untreated (0-day) versus Control (14-day filled with saline) results A.1. Objective 1 - Initial study A.2. Objective 2 - Fatigue A.3. Objective 3 - Fractography A.4. Objective 4 - Collagen degradation A.5. Objective 5 - Interface

216

For each objective, two additional 14-day groups of ten female and ten male bones were

filled with saline instead of KOH to act as controls. This appendix presents the results from

each objective comparing the untreated (0-day) versus control (14-day filled with saline)

groups for both sexes.

A.1. Objective 1 - Initial study

In Objective 1, female and male emu tibiae were treated with 1 M potassium hydroxide

(KOH) solutions for 1-14 days. DXA, QUS, MRTA and three-point bending measurements

were performed on the tibiae before and after KOH treatment. KOH treatment did not change

the geometry or mineral content of the bone samples and only a minimal amount of collagen

was extracted. This was confirmed with the control (14-day filled with saline) group, whose

data was comparable to the untreated (0-day) group. Below is a summary of this data.

Table A.1: Average geometrical parameters, BMD, BMC values and normalized mechanical properties of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in these parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Medial-Lateral diameter (mm) 28.6 ± 0.3 27.8 ± 0.3 28.5 ± 0.3 27.9 ± 0.3

Anterior-Posterior diameter (mm) 23.4 ± 0.2 22.8 ± 0.3 23.5 ± 0.2 23.2 ± 0.3

Cortical Thickness (mm) 4.2 ± 0.1 3.9 ± 0.1 4.2 ± 0.1 3.8 ± 0.1

Cortical area (mm2) 275 ± 5 240 ± 5 272 ± 5 239 ± 5

Moment of Inertia (mm4) 14132 ± 416 12068 ± 455 141115 ± 423 11995 ± 437

BMD (g/cm2) 1.3 ± 0.02 1.2 ± 0.02 1.3 ± 0.01 1.2 ± 0.01

BMC (g) 73.7 ± 1.1 65.4 ± 1.6 75.0 ± 1.5 64.7 ± 1.9

Ultimate Strength (MPa) 141.9 ± 5.1 124.7 ± 6.8 144.9 ± 7.5 129.6 ± 5.4

Failure Strength (MPa) 111.3 ± 6.0 70.8 ± 13.7 120.1 ± 9.5 76.4 ± 15.0

Failure strain (%) 1.6 ± 0.2 2.2 ± 0.3 1.5 ± 0.2 2.2 ± 0.4

Post Yield strain (%) 1.1 ± 0.2 1.8 ± 0.3 1.0 ± 0.2 1.7 ± 0.2

Toughness (mJ/mm3) 1.2 ± 0.2 1.6 ± 0.2 1.2 ± 0.1 1.6 ± 0.1

‘Post Yield’ Toughness (mJ/mm3) 1.0 ± 0.2 1.4 ± 0.2 0.9 ± 0.1 1.3 ± 0.2

Elastic Modulus (GPa) 16.9 ± 0.4 17.8 ± 0.6 15.9 ± 0.6 17.4 ± 0.9

217

Table A.2: Average geometrical parameter changes of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in geometrical parameter changes between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Change in medial-lateral diameter (%) 0.19 ± 0.2 0.34 ± 0.3 0.42 ± 0.8 0.13 ± 0.5

Change in anterior-posterior diameter (%) 0.34 ± 0.2 0.48 ± 0.4 0.40 ± 0.6 0.23 ± 0.4

Change in cortical thickness (%) 0.29 ± 0.6 0.39 ± 0.9 0.32 ± 0.3 1.0 ± 0.7

Change in cortical area (%) 0.11 ± 0.2 0.26 ± 0.4 0.17 ± 0.1 0.23 ± 0.4

Change in moment of inertia (%) 0.93 ± 0.5 0.89 ± 0.9 0.88 ± 0.7 0.71 ± 0.8 Table A.3: Average percent collagen weight removed and bone weight loss of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in the collagen weight removed and bone weight loss between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Percent collagen loss (%) 0.0024 ± 0.001 0.0023 ± 0.001 0.0024 ± 0.001 0.0023 ± 0.001

Percent bone mass change (%) 0.0004 ± 0.001 0.0004 ± 0.001 0.0004 ± 0.001 0.0004 ± 0.001

Table A.4: Percent changes of bone quality measurements reported by the different measurement techniques of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in the measured output from DXA, QUS, MRTA and three-point bending between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Percent change in BMD from DXA (%) 0.14 ± 0.2 0.26 ± 0.4 0.28 ± 0.3 0.15 ± 0.2

Percent change in SOS from QUS (%) 0.38 ± 0.2 0.48 ± 0.2 0.21 ± 0.3 0.27 ± 0.4

Percent change in EI from MRTA (%) 0.82 ± 1.2 1.28 ± 1.6 0.90 ± 1.0 1.13 ± 0.9

Percent change in E from three-point bending (%) 0.66 ± 0.6 0.87 ± 0.8 0.76 ± 1.0 1.01 ± 0.7

218

A.2. Objective 2 - Fatigue

The purpose of Objective 2 was to determine how emu bone fatigue properties are affected

by KOH treatment, a treatment targeted to affect only the organic component of bone and to

determine if current clinical tools could detect the effects of partial fatigue testing in

untreated and KOH treated bones. Similarly to Objective 1, KOH treatment did not change

the geometry or mineral content of the bone samples and a negligible amount of collagen was

extracted. Fatigue testing to 100,000 cycles also did not alter the geometry or mineral content

of the bone samples. This was confirmed with the control (14-day filled with saline) group,

whose data was comparable to the untreated (0-day) group. Below is a summary of this data.

Table A.5: Average geometrical parameters and BMD values of untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in these parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Medial-Lateral diameter (mm) 28.1 ± 0.3 27.9 ± 0.3 27.9 ± 0.4 28.0 ± 0.3

Anterior-Posterior diameter (mm) 24.0 ± 0.2 23.2 ± 0.3 24.2 ± 0.2 23.3 ± 0.2

Cortical Thickness (mm) 4.55 ± 0.14 3.91 ± 0.14 4.49 ± 0.23 3.94 ± 0.10

Cortical area (mm2) 292 ± 11 251 ± 10 285 ± 17 256 ± 13

Moment of Inertia (mm4) 15234 ± 729 11805 ± 685 14789 ± 865 11467 ± 601

BMD (g/cm2) 1.32 ± 0.03 1.20 ± 0.04 1.35 ± 0.02 1.19 ± 0.05

Table A.6: Average geometrical parameter changes of untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in geometrical parameter changes between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Change in medial-lateral diameter (%) 0.35 ± 0.9 0.56 ± 0.8 0.49 ± 0.6 0.60 ± 0.9

Change in anterior-posterior diameter (%) 0.38 ± 0.7 0.68 ± 0.7 0.47 ± 0.6 0.59 ± 0.9

Change in cortical thickness (%) 0.50 ± 0.8 1.02 ± 0.9 0.62 ± 0.9 0.76 ± 0.7

Change in cortical area (%) 0.40 ± 0.6 0.38 ± 0.4 0.45 ± 0.5 0.37 ± 0.5

Change in moment of inertia (%) 0.42 ± 0.6 0.69 ± 0.6 0.56 ± 0.7 0.63 ± 0.9

219

Table A.7: Initial secant modulus, creep strain at fracture and maximum strain at fracture for untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in these parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Eo (MPa) 24947 ± 1110 27821 ± 1139 26651 ± 1019 25113 ± 1110

εcreep_fracture (με) 3340 ± 793 3394 ± 800 3514 ± 1037 3487 ± 955

εmax_fracture (με) 7983 ± 1059 6457 ± 814 6582 ± 1033 6993 ± 934

Table A.8: Average percent collagen weight removed and bone weight loss of untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in the collagen weight removed and bone weight loss between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Percent collagen loss (%) 0.0034 ± 0.001 0.0017 ± 0.001 0.0029 ± 0.001 0.0023 ± 0.001

Percent bone mass change (%) 0.0009 ± 0.001 0.0007 ± 0.001 0.0004 ± 0.001 0.0006 ± 0.001

Table A.9: Strain levels during partial fatigue testing of untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in the strain levels between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Percent collagen loss (%) 1980 ± 34 2003 ± 20 2001 ± 42 1989 ± 37

220

Figure A.1: Peak stress versus log(N) curves for untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in fatigue behaviour between the untreated and control samples for both sexes. All tests were subjected to a constant minimum stress of 5 MPa.

Figure A.2: Peak stress versus damage index rate curves for untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in fatigue behaviour between the untreated and control samples for both sexes. All tests were subjected to a constant minimum stress of 5 MPa.

1 10 100 1000 10000 100000 10000000

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Cycles to failure (N)

Pea

k St

ress

(M

Pa)

1 10 100 1000 10000 100000 10000000

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Cycles to failure (N)

Pea

k St

ress

(M

Pa)

10-910-810-710-610-510-410-310-210-11000

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Damage index rate (1/sec)

Pea

k St

ress

(M

Pa)

10-910-810-710-610-510-410-310-210-11000

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Damage index rate (1/sec)

Pea

k St

ress

(M

Pa)

221

Figure A.3: Peak stress versus creep rate curves for untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in fatigue behaviour between the untreated and control samples for both sexes. All tests were subjected to a constant minimum stress of 5 MPa. Table A.10: Percent changes of bone quality measurements reported by the different measurement techniques of untreated (0-day) and control (14-day filled with saline) fatigue samples. No differences were observed in the measured output from DXA, QUS, MRTA and three-point bending between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Percent change in BMD from DXA (%) 0.26 ± 0.5 1.09 ± 0.3 0.34 ± 0.4 0.58 ± 0.7

Percent change in SOS from QUS (%) 1.38 ± 0.5 0.68 ± 0.5 1.72 ± 0.4 1.12 ± 0.6

Percent change in EI from MRTA (%) 19.28 ± 4.0 21.98 ± 4.1 20.99 ± 3.6 19.71 ± 4.7

Percent change in E from three-point bending (%) 20.16 ± 4.8 19.45 ± 3.5 20.02 ± 4.0 20.03 ± 3.8

10-910-810-710-610-510-410-310-20

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Creep rate (1/sec)

Pea

k St

ress

(M

Pa)

10-910-810-710-610-510-410-310-20

50

100

150

200Female Untreated (0-day)Male Untreated (0-day)Female Control (14-day saline)Male Control (14-day saline)

Creep rate (1/sec)

Pea

k St

ress

(M

Pa)

222

A.3. Objective 3 - Fractography

In Objective 3, we attempted to determine the failure mechanisms responsible for the altered

fracture behaviour of female and male emu tibiae endocortically treated with KOH. We

investigated the microscopic morphology of the three-point bending fracture surfaces of

untreated and KOH treated female and male emu cortical bone (from Objective 1) in an effort

to understand the altered mechanical properties previously observed. Crack morphology was

visualized qualitatively by taking SEM images of the tensile and compressive surfaces of

untreated and 14-day KOH treated samples. The fracture surfaces of 14-day female and male

KOH treated bones showed a significantly higher ‘roughness’ compared to untreated bones.

Furthermore, additional toughening mechanisms, which are important features for dissipating

energy during the failure process, were observed in the KOH treated samples, but were

absent in the untreated samples for both sexes. This was confirmed with the control (14-day

filled with saline) group, whose data was comparable to the untreated (0-day) group. Below

is a summary of this data.

Table A.11: Average tensile, compressive and transition areas of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in the percent tensile, compressive or transition areas between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Percent tensile region (%) 50 ± 2 50 ± 2 48 ± 2 49 ± 2

Percent compressive region (%) 34 ± 1 38 ± 2 37 ± 2 37 ± 2

Percent transition region (%) 15 ± 2 12 ± 2 15 ± 2 14 ± 2 Table A.12: Average percent area roughness measurements of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in the percent area roughness between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Failure Mode Characteristic

Female Male Female Male

%Rough (%) 50 ± 6 48 ± 5 51 ± 5 50 ± 4 Tensile Side (TS)

%Smooth (%) 50 ± 6 52 ± 5 49 ± 5 50 ± 4

%Rough (%) 44 ± 3 46 ± 4 43 ± 4 45 ± 4 Compressive Side (CS)

%Smooth (%) 56 ± 3 54 ± 4 57 ± 4 55 ± 4

223

Figure A.4: Representative SEM images of (a) untreated (0-day) male, (c) control (14-day saline) male, (e) untreated (0-day) female and (g) control (14-day saline) female tensile emu tibiae fracture surfaces. Magnified regions of interest are shown in (b) untreated male (0-day), (d) control (14-day saline) male, (f) untreated female (0-day) and (h) control (14-day saline) female emu tibiae. A relatively smooth surface can be seen in untreated and control samples for both sexes.

224

Figure A.5: Representative SEM images of (a) untreated (0-day) male, (c) control (14-day saline) male, (e) untreated (0-day) female and (g) control (14-day saline) female compressive emu tibiae fracture surfaces. Magnified regions of interest are shown in (b) untreated male (0-day), (d) control (14-day saline) male, (f) untreated female (0-day) and (h) control (14-day saline) female emu tibiae. A relatively smooth surface can be seen in untreated and control samples for both sexes.

225

Table A.13: Average porosity parameters from tensile side of untreated (0-day) and control (14-day filled with saline) fracture surfaces. No differences were observed in the percent area roughness between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female 14-day

No. of large pores (pore diameter >47 μm) per area (mm-2) 0.8 ± 0.1 1.0 ± 0.1 1.1 ± 0.2 1.0 ± 0.2

No. of small pores (pore diameter<47 μm) per area (mm-2) 75 ± 8 71 ± 5 72 ± 6 69 ± 7

226

A.4. Objective 4 - Collagen degradation

The purpose of Objective 4 was to determine if the mineral and collagen were affected by

KOH treatment using several mineral (powder XRD, microhardness testing, qBSE) and

collagen (α-chymotrypsin, DSC, SDS-PAGE, polarized light microscopy) characterization

techniques. KOH treatment causes in situ collagen degradation without affecting the mineral

phase. 14-day KOH treated emu tibiae for both sexes had significantly greater amounts of

degraded collagen compared to untreated samples. This was also confirmed with the control

(14-day filled with saline) group, whose data was comparable to the untreated (0-day) group.

Below is a summary of this data.

Table A.14: Emu bone mineral crystal length (002) and cross section (310) estimated by XRD and microhardness testing results of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in BSE parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Crystal length 26° peak (Å) 205 ± 2 209 ± 2 207 ± 3 206 ± 2

Crystal cross section 40° peak (Å) 62 ± 0.5 63 ± 0.6 63 ± 0.7 62 ± 0.7

Microhardness, HV (kg/mm2) 61 ± 0.4 62 ± 0.5 60 ± 0.7 61 ± 0.6

Table A.15: Quantitative BSE results of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in BSE parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Peak grey level (pixels) 177 ± 7 175 ± 9 180 ± 8 179 ± 8

FWHMH (pixels) 25 ± 3 23 ± 3 24 ± 3 23 ± 4

Large Pore Density (#/mm2) 0.23 ± 0.06 0.22 ± 0.07 0.18 ± 0.08 0.19 ± 0.06

Small Pore Density (#/mm2) 110 ± 4 110 ± 4 108 ± 4 109 ± 4

Total Length of Dark Band (μm) 190 ± 45 220 ± 45 198 ± 51 195 ± 46

227

Table A.16: Average percent digested collagen from the α-chymotrypsin technique of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in percent digested collagen between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Percent digested collagen (%) 16.53 ± 0.93 18.89 ± 0.64 17.48 ± 0.53 18.55 ± 0.47

Table A.17: Average female and male emu tibiae thermal characteristics from DSC of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in DSC parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male

Tonset (°C) 48.5 ± 1.6 47.7 ± 1.5 48.2 ± 1.8 48.7 ± 1.4

Tpeak (°C) 60.9 ± 0.2 60.6 ± 0.4 60.7 ± 0.4 60.4 ± 0.3

FWHMH (°C) 18.4 ± 1.1 17.6 ± 0.8 17.2 ± 0.8 18.1 ± 0.8

Enthalpy (ΔH; J/g of collagen) 38.7 ± 0.8 38.6 ± 1.2 38.9 ± 0.9 38.5 ± 1.4

Height (mW/mg of collagen) 0.17 ± 0.01 0.16 ± 0.01 0.17 ± 0.01 0.17 ± 0.01

Table A.18: Average relative peak intensity from SDS-PAGE of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in percent digested collagen between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Average relative peak intensity (Arbitrary Units) 16944 ± 189 16948 ± 218 16714 ± 345 17019 ± 287

228

A.5. Objective 5 - Interface

The goal of Objective 5 was to understand the effect of KOH treatment on the bone mineral-

collagen interface using Raman spectroscopy and AFM. The Raman spectra of bone samples

treated with KOH showed an increase in peak position and band width of the amide III band

for both sexes compared to untreated samples. AFM images revealed a ‘swollen’ appearance

in KOH treated samples compared to untreated samples. No differences were observed in

Raman spectroscopy parameters as well as AFM images between the untreated (0-day) and

control (14-day filled with saline) groups for both sexes. Below is a summary of this data.

Table A.19: Average female and male emu tibiae Raman spectroscopy parameters of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in Raman spectroscopy parameters between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Parameter

Female Male Female Male Mineralization (phosphate/amide I) 9.17 ± 0.55 10.18 ± 0.91 9.23 ± 0.66 9.67 ± 0.45

Mineralization (phosphate/amide III) 18.47 ± 0.71 18.39 ± 1.48 18.45 ± 1.21 18.21 ± 1.64

Mineralization (phosphate/CH2)

11.20 ± 0.70 12.25 ± 1.19 12.11 ± 0.88 11.75 ± 1.94

Carbonate Substitution (phosphate/carbonate) 4.76 ± 0.14 4.56 ± 0.29 4.42 ± 0.27 4.47 ± 0.34

Mineral Crystallinity (1/phosphate FWHMH) 0.057 ± 0.001 0.056 ± 0.001 0.059 ± 0.002 0.058 ± 0.002

Table A.20: Average Raman peak widths of the major bands of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in Raman peak widths between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) FWHMH (cm-1)

Female Male Female Male

Phosphate 17.60 ± 0.36 17.19 ± 0.33 17.45 ± 0.27 17.30 ± 0.41

Carbonate 53.00 ± 1.83 52.99 ± 1.36 53.01 ± 2.01 53.00 ± 1.18

Amide III 41.01 ± 1.36 40.23 ± 1.46 41.11 ± 1.26 40.55 ± 0.99

C-H bending 30.71 ± 0.81 29.90 ± 1.21 29.98 ± 1.01 30.52 ± 0.69

Amide I 50.65 ± 0.93 51.81 ± 0.85 50.85 ± 0.37 50.47 ± 0.91

229

Table A.21: Average Raman peak positions of the major bands of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in Raman peak positions between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Peak position (cm-1)

Female Male Female Male

Phosphate 964.8 ± 0.2 965.1 ± 0.3 964.9 ± 0.3 965.0 ± 0.1

Carbonate 1071.5 ± 1.0 1073.0 ± 1.1 1072.0 ± 1.2 1072.1 ± 0.8

Amide III 1260.6 ± 1.2 1263.5 ± 2.3 1262.7 ± 2.0 1263.0 ± 2.7

C-H bending 1457.4 ± 0.7 1457.0 ± 1.0 1457.2 ± 1.2 1457.0 ± 1.4

Amide I 1666.5 ± 0.8 1665.0 ± 0.8 1665.7 ± 0.6 1666.0 ± 1.1

Table A.22: Quantitative grey level distribution and surface roughness results for female and male emu tibiae of untreated (0-day) and control (14-day filled with saline) samples. No differences were observed in secondary surface roughness measurements between the untreated and control samples for both sexes.

Untreated (0-day) Control (14-day saline) Peak position (cm-1)

Female Male Female Male Peak grey level (pixels) 77 ± 3 78 ± 4 77 ± 6 75 ± 5

Width at half peak height (pixels) 53 ± 2 50 ± 1 53 ± 1 49 ± 2

Surface Roughness (Bio scope™: r, nm2) 124 ± 12 101 ± 10 111 ± 9 113 ± 11

Surface Roughness (Optical profiler: Ra, μm) 1.7 ± 0.2 1.4 ± 0.1 1.8 ± 0.1 1.4 ± 0.1