265
in situ Nanomechanical Investigations of Bone Hang, Fei For additional information about this publication click this link. http://qmro.qmul.ac.uk/jspui/handle/123456789/2399 Information about this research object was correct at the time of download; we occasionally make corrections to records, please therefore check the published record when citing. For more information contact [email protected]

in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

in situ Nanomechanical Investigations of BoneHang, Fei

For additional information about this publication click this link.

http://qmro.qmul.ac.uk/jspui/handle/123456789/2399

Information about this research object was correct at the time of download; we occasionally

make corrections to records, please therefore check the published record when citing. For

more information contact [email protected]

Page 2: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

in situ Nanomechanical

Investigations of Bone

A THESIS SUBMITTED TO THE UNIVERSITY OF LONDON

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

SEPTEMBER 2011

BY

FEI HANG

SCHOOL OF ENGINEERING AND MATERIALS SCIENCE

QUEEN MARY, UNIVERSITY OF LONDON

MILE END ROAD, LONDON, E1 4NS

Page 3: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Declaration

2

Declaration

I declare that the work performed is entirely by myself during the course of my

PhD studies at the Queen Mary, University of London and has not been submitted

for a degree at this or any other University.

Fei Hang

Page 4: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Abstract

3

Abstract

Mineralized collagen fibrils (MCFs) are the fundamental building blocks that

contribute to the extraordinary mechanical behaviour of bone. Despite its

importance in defining bone mechanics, especially the high resistance to fracture

recorded in bone tissue, MCFs have yet to be mechanically tested and, thus, MCF

contributions to the global mechanical properties of bone is unclear. In this thesis,

a complete strategy for performing direct mechanical testing on nanosized

fibrous samples including MCFs from bone using a novel in situ atomic force

microscope (AFM) – scanning electron microscope (SEM) combination was

established. This technique was used to mechanically test MCFs from antler bone

tissue for the first time and resultant stress-strain behaviour was recorded to

highlight the inhomogeneous response of fibrils, which is associated with fibrillar

compositional heterogeneity. Mechanical properties of MCFs and bone tissue

were found to be controlled by biomineralization process using additional tensile

testing of MCFs and bulk samples from mouse limb bones at different ages.

Page 5: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Abstract

4

Extrafibrillar mineralization was found to have effects on the Young’s modulus of

bone tissue rather than fibrils, indicating the importance of fibrillar interfaces in

controlling overall mechanical behaviour of bone tissue. Interfaces between

fibrils in bone were examined by carrying out single fibril pullout tests. A weak

but reformable interface, dominated by ionic bonds between fibrils, was

recorded and the sacrificial bond reforming activity at the interface was found to

be dependent on pullout strain rate. Finally, considerations of bone as a fibrous

composite was used to evaluate nanomechanical testing data, with approximately

50 % of the bone fracture energy accounted for in failure of fibril interfaces.

Page 6: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Acknowledgements

5

Acknowledgements

I am grateful to my supervisor Dr. Asa Barber for his continuous support, patient

guidance, stimulating discussions, trust and encouragement throughout my

research. I also want to thank my second supervisor Dr. Himadri Gupta for his

inspiring suggestions and ideas. All the other researchers and technicians who

supported me are gratefully acknowledged. Particularly, I would like to thank

Prof. Ton Peijs for his valuable discussion on composite theories. Special thanks

to Dr. Zofia Luklinska and Mick Willis at the NanoVision Centre and Dr. Emiliano

Bilotti in Nanoforce for their help in experiments.

The work in this thesis is also supported by friends and cooperative partners.

Special thanks go to Dr. Dun Lu who is my close friend and our group member for

his help and discussion especially for the efforts we put together in developing

the testing methodology. Thanks are given to Ines Jimenez for her valuable

discussion on literature and results, Russell Bailey for his help in developing data

analysis, and all other current and former group members I have worked with: Dr.

Page 7: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Acknowledgements

6

Urszula Stachewicz, Congwei Wang, Beatriz Cortes, Dr. Shuangwu Li and Dr. Wei

Wang. I would like also express my appreciation to Dr. Martin Zech and Dr.

Christoph Bödefeld from Attocube Systems AG for working together on

developing the experimental technique used in this thesis and Angelo

Karunaratne for his help in testing mouse bone in Chapter 5. Thanks to Prof. John

Currey for his discussion on antler mechanics and providing antler samples for

this thesis. Thanks to Prof. Peter Fratzl at Max Planck Potsdam, Prof. Markus

Buehler from MIT and Prof. Robert Ritchie from UC Berkeley for their valuable

and inspiring discussions on MCF mechanics during MRS meetings.

Special acknowledgement to my dearest parents and my dearest wife Yaxue to

whom this work is dedicated and all my close friends who gave me support and

encouragement in the difficult times.

Page 8: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Publications Arising from this Work

7

Publications Arising from this

Work

1. Hang, F., & Barber, A. H. (2011) Nano-mechanical properties of individual

mineralized collagen fibrils from bone tissue. J. Royal Society Interface. 8(57):

500-505

2. Hang, F., Lu D., Bailey R., Palomar, I. J., Stachewicz U., Cortes-Ballesteros B.,

Davies M., Zech M., Bödefeld C., Barber, A.H. (2011) In situ tensile testing of

nanofibers by combining atomic force microscopy with scanning electron

microscopy. Nanotechnology. 22(36): 365708.

3. Hang, F., & Barber, A. H. (2010) Experimental investigations into the

mechanical properties of the collagen fibril-noncollagenous protein interface

in antler bone. MRS Proceedings, Biological Materials and Structures in

Physiologically Extreme Conditions and Diseases, 1274, 119-123.

4. Hang, F., Lu, D., & Barber, A. H. (2009). Combined AFM-SEM for mechanical

testing of fibrous biological materials. MRS Proceedings, Structure-Property

Relationships in Biomineralized and Biomimetic Composites, 1187, 135-140.

5. Hang, F., Lu, D., Li, S. W., & Barber, A. H. (2009). Stress-strain behavior of

individual electrospun polymer fibers using combination AFM and SEM. MRS

Proceedings, Probing Mechanics at Nanoscale Dimensions, 1185, 87-91.

Page 9: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Publications Arising from this Work

8

6. Hang, F., Barber, A. H. (2011) Composite toughness in bone from

nano-mechanical considerations. In preparation

7. Hang, F., Karunaratne, A., Gupta, H. & Barber, A. H. (2011) Insensitivity of

individual mineralized collagen fibrils to age. In preparation

8. Hang, F., Palomar, I.J., Gupta, H. & Barber, A. H. (2011) Reforming efficiency of

sacrificial ionic bonds at nano-interfaces in bone. In preparation

Page 10: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Table of Contents

9

Table of Contents

Abstract .................................................................................................................................. 3

Acknowledgements ........................................................................................................... 5

Publications Arising from this Work ........................................................................... 7

Table of Contents ................................................................................................................ 9

List of Tables ...................................................................................................................... 13

List of Figures .................................................................................................................... 15

List of Constants ................................................................................................................ 29

Chapter 1 Introduction ................................................................................................... 31

1.1 Background ..................................................................................................... 31

1.2 Current Studies on Mechanics of Bone ........................................................... 33

1.3 Project Aims and Thesis Outline ..................................................................... 36

1.3.1 Project Aims ......................................................................................... 36

1.3.2 Thesis Outline ...................................................................................... 38

Chapter 2 Literature Review ........................................................................................ 41

2.1 Introduction ..................................................................................................... 41

2.2 Bone Formation and Secondary Bone............................................................. 41

2.3 Hierarchy Structure of Bone ........................................................................... 46

2.3.1 Level 1: Major Components of Bone .................................................. 48

2.3.2 Level 2: Mineralized Collagen Fibrils (MCFs) .................................... 49

2.3.3 Level 3 Fibril Array .............................................................................. 52

Page 11: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Table of Contents

10

2.3.4 Level 4: Fibril Array Pattern ............................................................... 53

2.3.5 Level 5: Osteonal/Haversian System ................................................. 55

2.3.6 Level 6: Architecture Level – Compact and Cancellous Bone .......... 56

2.3.7 Level 7: Whole Bone ............................................................................ 57

2.4 Bone’s Structure and Mechanical Properties .................................................. 59

2.4.1 Level 7: Whole bone ............................................................................ 61

2.4.2 Level 6: Architecture Level ................................................................. 66

2.4.3 Level 5-3: From Tissue to Fibrils ........................................................ 92

2.4.4 Level 2: Mineralized Collagen Fibrils ................................................. 93

2.4.5 Sub-structural Components of Bone .................................................. 97

2.5 Conclusions ................................................................................................... 102

Chapter 3 Methodology ............................................................................................... 103

3.1 Materials Preparation .................................................................................... 103

3.1.1 Antler Bone Sample Preparation ..................................................... 104

3.1.2 Mouse Bone Sample Preparation ..................................................... 107

3.2 in situ Nanomechanical Testing .................................................................... 109

3.2.1 Introduction ....................................................................................... 109

3.2.2 AFM-SEM Setup and in situ Tensile Test .......................................... 115

3.2.3 Data Analysis Principle ..................................................................... 124

3.2.4 Calibration and Data Analysis .......................................................... 133

3.2.5 Accuracy and Errors .......................................................................... 137

3.2.6 Results of Tensile Test on Polymeric Nanofibres and Discussion.. 142

3.3 Conclusions ................................................................................................... 146

Chapter 4 Nanomechanical Properties of Individual Mineralized Collagen

Fibrils from Bone Tissue............................................................................................. 147

4.1 Introduction ................................................................................................... 147

4.2 Materials and Methods .................................................................................. 148

4.2.1 in situ Nanomechanical Testing ........................................................ 148

Page 12: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Table of Contents

11

4.2.2 Compositional Study Using EDS ....................................................... 152

4.2.3 Thermal Gravimetric Analysis .......................................................... 152

4.3 Results ........................................................................................................... 153

4.3.1 Mechanical Behaviour of Individual MCFs ...................................... 153

4.3.2 Compositional Analysis ..................................................................... 155

4.3.3 Water Content Change in Bone Samples Exposed to SEM Vacuum

Studied by TGA ............................................................................................ 156

4.4 Discussion ..................................................................................................... 159

4.5 Conclusions ................................................................................................... 164

Chapter 5 Mechanical Properties of MCFs from Differing Ages of Bone .... 165

5.1 Introduction ................................................................................................... 165

5.2 Materials and Methods .................................................................................. 167

5.2.1 Animals ............................................................................................... 167

5.2.2 Sample Preparation of Bone Tissue ................................................. 167

5.2.3 Tensile Testing of Bulk Bone Tissue ................................................. 169

5.2.4 in situ Nanomechanical Testing of MCFs .......................................... 171

5.3 Results ........................................................................................................... 171

5.4 Discussion ..................................................................................................... 174

5.5 Conclusions ................................................................................................... 185

Chapter 6 Nanoscale Interfacial Behaviour in Bone ......................................... 187

6.1 Introduction ................................................................................................... 187

6.2 Materials and Method ................................................................................... 190

6.2.1 Material Preparation ......................................................................... 190

6.2.2 in situ AFM-SEM Fibril Pullout Experiment .................................... 191

6.3 Results and Discussion .................................................................................. 193

6.4 Conclusions ................................................................................................... 205

Chapter 7 Mineralized Collagen Fibril Contributions to the Fracture

Behaviour of Bone ......................................................................................................... 207

Page 13: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Table of Contents

12

7.1 Introduction ................................................................................................... 207

7.1.1 Pullout Theory: Energy Balance Model ........................................... 208

7.2 Results ........................................................................................................... 213

7.3 Discussion ..................................................................................................... 220

7.4 Conclusions ................................................................................................... 226

Chapter 8 Conclusions and Future Work .............................................................. 228

8.1 Summary of the Thesis .................................................................................. 228

8.2 Future Work ................................................................................................... 232

8.3 Major Findings of the Thesis ........................................................................ 234

References ....................................................................................................................... 237

Page 14: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Tables

13

List of Tables

2.1 Young’s moduli of human and bovine bone specimens (in GPa).

Reproduced from data in [12, 109, 111]. Subscripts 1, 2 and 3

represent the radial, circumferential and longitudinal directions

relative to the long axis of the bone.

71

2.2 Strength of human and bovine bones (in MPa) summarizing

results found in previous studies [11, 12, 109, 112, 115].

74

2.3 Recorded Young’s moduli of cancellous bone material from Currey

[12]

87

3.1 Mechanical properties of PVA and nylon-6 nanofibres measured by

in situ AFM-SEM and their bulk equivalents. E is the Young’s

modulus measured for each sample in their linear region up to 4%

strain. UTS is the ultimate tensile strength and UTS is the ultimate

tensile strain of the sample. Comparable film and bulk properties

were measured from PVA films or taken from literature for nylon

films (*see [206]).

143

4.1 Mechanical properties of six individual MCFs tensile tested to

failure using combined AFM-SEM

155

4.2 TGA test on antler bone samples showing water content changes

with different exposure time to the vacuum in SEM chamber.

158

Page 15: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Tables

14

5.1 The mechanical behaviour of MCFs tensile tested to failure. 174

5.2 Quantitative analysis of qBSE images taken on bone at different

development stages (1 to 16 weeks)

180

6.1 Data showing the geometry, work done and calculated interfacial

behaviour in MCF pullout tests.

203

Page 16: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

15

List of Figures

1.1 Scanning electron micrographs showing the fracture surface of

equine bone (A) and bovine bone (B) at comparable

magnifications. Separated mineralized collagen fibrils are clearly

identified in each sample as hair-like structures protruding from

the bone surfaces.

32

1.2 Atomic force microscopy can be used to deform (but not fail)

individual collagen fibrils from bovine tendon [20]. (A) An AFM

probe with glue at its apex is used to pick up a collagen fibril from

a Teflon coated surface with the one end of the fibril fixed in a

drop of glue on the surface. (B) Schematic diagram indicating the

AFM configuration, after fibril pick-up, used for mechanical

testing.

34

1.3 Individual collagen fibrils from sea cucumber tensile tested using

(A) a MEMS device in ambient/humid environment. (B) Individual

collagen fibrils were isolated from the bulk and tensile tested

between the grips of the MEMS device [40].

35

2.1 Schematic diagram showing the stages of endochondral

ossification [53].

42

2.2 Optical microscope image showing ossification from epiphysial 44

Page 17: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

16

cartilage to mineralized collagen/woven bone [64].

2.3 The hierarchical levels of structure found in secondary osteonal

bone, as demonstrated by Weiner and Wagner [1]

47

2.4 Schematic diagram showing structure of a collage fibrils

composed of self-assembled tropocollagen molecules

50

2.5 (A) 3-D organization of apatite minerals inside a single collagen

fibril. Collagen molecules (represented by red rods) aggregate into

a quarter-staggered pattern described by Hodge and Petruska

[82]. Calcium phosphate (blue plates) nucleates in the gaps

between the collagen molecules. (B) TEM image of mineralized

collagen fibrils from mineralized turkey tendon showing the

characteristic quarter-stagger pattern [83].

51

2.6 Schematic illustration of MCFs containing mineral plates (not

drawn to scale, MCFs are shown in cylinders with black mineral

crystals embedded) showing (A) an arrangement of mineralized

collagen fibrils aligned both with respect to crystal layers and

fibril axes (orthotropic symmetry). (B) Arrangement of

mineralized collagen fibrils with only the fibril axes aligned

(transversal isotropy) [1].

52

2.7 Scanning electron micrographs and schematics highlighting the

four most common fibril array pattern organizations in bone. SEM

micrographs of fractured surfaces and schematic illustrations (not

drawn to scale) of the basic organizational motifs indicate: (A)

Array of parallel fibrils of mineralized turkey tendon (scale: 0.1

mm) and schematic illustration showing the localized orthotropic

symmetry of a fibril bundle. (B) Woven fibre structure of the outer

layer of a 19-week old human fetus femur [94]. The schematic

illustration shows fibril bundles with varying fibril diameters

54

Page 18: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

17

arranged in random orientations. (C) Plywood-like structure

presenting fibril organization in lamellar bone from the fracture

surface of a baboon tibia, with prominent fourth (large

arrowhead) and fifth (small arrowhead) sub-layers [91, 95]. The

corresponding schematic illustration shows the five sub-layer

model described in [91] with sub-layers one (right hand side),

two, and three arbitrarily composed of one fibril layer each,

whereas sub-layers four and five are composed of four fibril layers

each. Note that the fibrils in each layer are rotated relative to their

neighbours (depicted by the change in direction of the ellipsoid

cross-section), following the rotated plywood model [89]. (D)

Radial fibril arrays from human dentin fractured roughly parallel

to the pulp cavity surface. The tubules (holes) are surrounded by

collagen fibrils that are approximately in one plane. The schematic

illustration of the fibril bundles highlights the arranged in a plane

perpendicular to the tubule long axis with no obvious preferred

orientation within the plane [1, 2].

2.8 (A) Micro-anatomy structure of a human femur head with

compact and cancellous bone demonstrated. (B) Schematic

diagram showing the structure of cortical bone [97].

57

2.9 Schematic diagram showing a range of different bone shapes

found in the human skeleton [98]

58

2.10 A possible design criterion for a limb bone. The long bone is

initially straight with a length of L.

63

2.11 (A) A beam of original length L under an external loading

condition will (B) bend, causing a beam length change above and

below the neutral plane. (C) Cross section of the beam showing

the second moment of area will be different if loaded at different

65

Page 19: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

18

location on the cross section. δarea is the unit area at a distance y

from the neutral plane.

2.12 Loading bone in different directions. Brown cylinders represent

osteons while the interface region is shown in yellow (not to

scale).

67

2.13 Schematic diagram showing three loading directions in

mechanical test of bone material.

68

2.14 A typical stress-strain curve of cortical bone recorded in tensile

test on human bone along longitudinal axis. The black line shows

the linear response at small strain, with the slope defining the

bone’s Young’s modulus. Yield stress and failure strain of the bone

were also recorded as shown [107].

69

2.15 Three-point bending for toughness measurements. (A) Schematic

showing the three point bending test for measuring toughness. (B)

A typical load-displacement curve obtained from testing on mouse

tibia bone in the longitudinal direction [121]

78

2.16 Anisotropy in the fracture behaviour of bone. (A) and (B) show the

crack propagate along osteon interfaces and through osteons to

create longer crack paths in three point bending tests on bone at

the longitudinal direction (defined as the osteon parallel to

specimen long axis). (A) Scanning electron micrograph showing

crack tip and (B) schematically shows the crack growth. (C) and

(D) demonstrate the crack developing in bone in the three point

bending test on bone in the transversal direction (osteon vertical

to specimen long axis). In the schematic diagrams (B) and (D), the

osteons are represented in brown while the interface regions are

in yellow.

80

2.17 (A) to (D) Different toughening mechanisms in compact bone at 83

Page 20: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

19

the architectural level. (E) Example of an R-curve [136] (Fracture

toughness vs. crack length) for an equine femur bone specimen.

2.18 (A) Culmann’s schematic diagram showing principal stress line in

an engineering crane. The crane was under a distributed vertical

load from the top. (B) Wolff’s depiction of trabecular arrangement

in proximal femur bone. Cited from Wolff [143].

89

2.19 Stress-strain curve recorded in a cancellous bone compression

test from Currey [12] based on data derived from Fyhire and

Schaffler [150].

91

2.20 Schematic of direct mechanical testing on mineralized collagen

fibrils. (A) Force spectroscopy experiment on collagen fibrils from

Graham’s work. The fibrils are adsorbed between a clean glass

surface and an AFM cantilever. The movement of the cantilever

caused the resultant strain of collagen fibril and recorded by laser

optical sensor [39]. (B) A similar test improved by van der Rijt et

al using a drop of glue on the glass surface [20]. Both of the tests

recorded relatively low modulus and did not measure the fracture

properties of collagen fibrils.

96

2.21 Computational simulation results of the stress-strain response of a

mineralized fibril versus an unmineralized collagen fibril,

demonstrating the significant effect of the presence of mineral

crystals in collagen fibril on its mechanical behaviour [138].

100

3.1 Struers Accutom-5 water-cooled rotating diamond saw. 105

3.2 (A) Optical image of an antler cross section, highlighting the

compact cortical shell and the trabecular core. (B) Schematic

representation of the sample: S shows the sample location and the

orientation within the cortical bone C of an antler section; the

trabecular region is indicated by T; (C) Optical microscopy image

106

Page 21: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

20

of a polished and HCl etched cortical bone cross section surface,

showing the typical tissue composition used for the experiments.

(D) and (E) scanning electron micrographs on fractured antler

samples at different magnifications showing exposed MCFs.

3.3 (A) Schematic skeletal system of a mouse indicating the location of

femur and radius. (B) Fresh mouse radius bone from a wild type

10 week old male mouse with muscle and connective tissue

covering the bone surface. (C) Scanning electron micrograph

showing individual mineralized collagen fibrils at the bone

fracture surface.

108

3.4 (A) Schematic diagram showing the in situ configuration for

tensile testing of a nanofibrous sample using combined AFM-SEM.

The insert shows fixing of the nanofibre with glue between the

substrate and AFM cantilever. (B) Optical photograph showing

AFM fitted on SEM sample stage when door of SEM chamber is

open. Dashed rectangle indicates location of the AFM. Secondary

Electron axis (SE) and Focused Ion Beam axis are also labelled in

the photo. (C) Higher magnification optical side view image of the

AFM on the SEM sample stage.

117

3.5 Scanning electron micrographs showing a typical tensile test of

electrospun nanofibres (PVA) using in situ AFM-SEM. (A) A single

PVA nanofibre was attached to an AFM probe using epoxy glue at

the end of probe. (B) Sectioning of the PVA nanofibre from the

fibrous mat provides isolation at the end of AFM probe. (C)

Insertion of the free PVA nanofibre end into glue provides the

standard nanotensile testing configuration. (D) PVA nanofibre

tensile testing carried out after solidification of the glue was

achieved by translation of the AFM probe away from the glue

119

Page 22: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

21

surface until the nanofibre failed at its mid-length. The

deformation of epoxy glue at the anchor point with fibre and was

carefully examined by pixel analysis and excluded from original

tensile test result. The maximum deformation of epoxy glue is

relatively small (0.33 m) compared to the length of the tested

fibre recorded before and after testing (10.2 m and 12.1 m

respectively)

3.6 Schematic diagram of the laser interferometer used in the AFM

illustrating the interference signal measured. Approximately 4 %

of the laser light is reflected at the glass-air interface with an

intensity of I1 while ~96 % of the light is transmitted and partially

reflected at the AFM cantilever with an intensity of I2. The

intensity I2 depends on the reflectivity of the AFM cantilever.

123

3.7 Force spectroscopy plot from tensile testing of an individual

nanofibre. (A) Original sinusoidal laser intensity-distance curve

recorded in force spectroscopy. The Y-axis is reflected laser

intensity while the X-axis indicates the piezoscanner retraction

distance. Failure of the nanofibre is observed at a retraction

distance of 15 μm. (B) Resultant stress-strain curve converted

from the sinusoidal curve based on Equation 3.4

125

3.8 Schematic diagrams showing the deviation of laser light reflecting

off a deflecting AFM cantilever.

128

3.9 A typical Lorentzian fitting on original data obtained in a tensile

testing experiment on fibre sample. The black line is the original

data plot. Red line is the Lorentzian fitting

130

3.10 Schematic diagram of cantilever in different working modes. (A)

Z-Spectroscopy is the usual force spectroscopy tool used for

indentation and compression test. The cantilever bends up by

131

Page 23: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

22

applying compressive load. (B) Dither-Spectroscopy is the

maintenance mode in which the cantilever is driven by dither

piezo mounted behind the cantilever, which only changes the

distance between cantilever and fibre optics. The cantilever is not

loaded and therefore does not bend. (C) In nanofibre experiments,

the cantilever is under tensile load and bends down.

3.11 Schematic diagrams showing different relative positions of the

fibre optics along the AFM cantilever (positions d1 and d0). Orange

rectangles represent different locations of fibre optics after

installing AFM cantilever. When considering a constant cantilever

bending, the distance between cantilever and fibre optics at each

of these positions change due to different laser positions. The

insert shows the bottom view from the AFM probe with the

circular in orange representing the fibre optics.

132

3.12 An example of a laser intensity versus cantilever displacement

plot obtained from tensile testing on an individual electrospun

PVA fibre and subsequent calibration. (A) Part of the original laser

intensity with cantilever displacement plot recorded during

tensile test. (B) The same dataset after normalization. (C) Part of

the original laser intensity with cantilever displacement plot of

calibration. (D) Calibration data plot after normalization.

135

3.13 Plot of laser intensity reflected from the AFM cantilever against

time for the AFM integrated within SEM (1 kHz data sampling

bandwidth, cantilever spring constant is 0.02Nm-1).

140

3.14 Tensile stress-strain curves for different fibrous samples. (A) 5

PVA nanofibres exhibit a linear stress-strain relationship before

10 % strain followed by non-linear behaviour indicating ductile

yield in one of the samples. (B) One of the five nylon-6 nanofibres

144

Page 24: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

23

(N3) shows double yield point behaviour in the stress-strain plot,

suggesting crystal structure changes during deformation, while

other nanofibres exhibit clear yield behaviour.

4.1 (A) Scanning electron micrograph showing a typical testing

configuration for tensile testing of MCFs. The image shows a large

number of exposed collagen fibrils observed at the fracture

surface of antler bone. An individual collagen fibril protruding

from the fracture surface is attached to the glue at the end of the

AFM probe. Translation of the AFM probe away from the fibril

causes tensile deformation of the fibril until failure occurs which

is shown in the inserted image. (B) Schematic diagram showing

the combined SEM-AFM setup.

149

4.2 (A) The stress-strain plot for tensile testing of individual MCFs.

The MCFs show a linear stress-strain regime with a linear elastic

modulus of 2.4 ± 0.4 GPa. At higher strains above approximately

2-3 % strain, the MCFs exhibit either yield or an apparent increase

in the tangential modulus. The insert shadow area shows the

inhomogeneity of the strain response with increasing stress. (B)

Failure strength verse ultimate strain recorded by stretching

collagen fibrils until failure. The arrow shows increasing strength

along with decreasing fracture strain, indicating a ‘brittle like’

change.

154

4.3 Scanning backscattered electron micrograph of the antler bone

fracture surface. The light regions on the image indicate higher

mineralization. The corresponding Ca/P ratio in the image varied

from 1.48 to 3.12, indicating hydroxyapatite mineral is present

[28, 225] and the content varies throughout the antler bone.

156

4.4 Plot of water content in bone with time of SEM vacuum exposure. 157

Page 25: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

24

The water content in bone was measured from the weight loss

recorded during TGA tests by heating bone samples up to 200 C.

4.5 Tensile stress-strain curves for mineralized collagen fibrils

showing two distinct mechanical behaviours. Tropocollagen

uncoiling occurs initially in both types of fibrils within Region I. In

(A) the fibril shows an enhanced elastic modulus in Region II due

to the mineral increasing the stress transfer between

tropocollagen macromolecules. In (B) the low mineral density

within the fibrils allows sliding between tropocollagen molecules,

resulting in a plastic deformation.

161

5.1 Sample preparation for tensile testing (A) Schematic of mouse

skeleton showing the location of the femur (ellipse). (B) Optical

image of an isolated femur bone. (C) Custom made micro milling

setup used to remove bone leaving 0.1 mm thick of anterior

quadrants. (D) Bony ends of Femur included in the dental cement

(FiltekTM Supreme XT, 3M ESPE, USA) and milled mid diaphysis.

168

5.2 Schematic view of the micro tensile testing machine. The sample is

immersed in Hank’s buffered solution. Tensile load is applied

along the schematically prepared femur long axis as shown in the

inset.

170

5.3 Mechanical behaviour of femur samples from 4 and 10 week old

mouse bone recorded during tensile testing showing (A)

stress-strain plots of bone samples with linear regressions and (B)

summary plot showing the Young’s modulus calculated from

stress-strain plot linear regressions for the 4 and 10 week old

bone.

172

5.4 Stress-strain behaviour of individual mineralized collagen fibrils

from mouse radius bone at 4 and 10 weeks respectively.

173

Page 26: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

25

5.5 Linear regressions of average stress-strain plots for bulk bone and

MCFs samples from 4 week and 10 week old mouse bone. Error

bars are standard deviations (see Table 5.1).

175

5.6 BSE microscopy images of mid shaft cross sections of limb bones

from 1 week to 16 weeks for wild type mice.

176

5.7 Quantitative analytical diagrams of BSE microscopy images taken

from bone of different ages. (A) Frequency of Appearance against

calcium concentration with age. (B) Histogram of FWHM of BMDD

plotted with development which corresponds to the distribution

in local variation of Calcium content and (C) Mean calcium

concentration in form of weighted fraction (Ca wt %) calculated

from (A) was plotted as a function of development (1 week to 16

weeks).

179

5.8 Schematic sketches showing three stages of biomineralization on

collagen fibrils in mouse bone. Tropocollagen molecules are

denoted as blues cylinders and the hydroxyapatite mineral

crystals are red particles with random shapes. Negatively charged

molecules (most of which are Non-Collagenous Proteins as

discussed in Chapter 2) existing in the extrafibrillar space are

shown in green and bonded via positive calcium ions. (A) Calcium

and phosphate ions migrate to the gap regions (initially filled with

water) within collagen fibrils and initiate nucleation of HA

crystals. (B) All the gap regions within collagen fibrils are

completely filled with mineral crystals. (C) Mineral continues to

grow within the extrafibrillar space and cover the fibril surfaces.

181

5.9 Young’s modulus measured by different methods. (A) Tangent

slope of stress-strain curve at small strain is recorded as Young’s

modulus. (B) Young’s modulus is measured as the linear

185

Page 27: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

26

regression of data points on a stress-strain curve.

6.1 (A) Scanning electron micrographs showing an AFM probe

containing glue at its apex attached to an individual MCF partially

embedded in a fibril bundle at the fracture surface of antler bone.

(B) The exposing fibril was pulled out by the AFM with images at

high magnification used to measuring the fibril embedded length

le.

192

6.2 Plot showing the force applied to the partially exposed MCF

during pullout against progression time for the pullout

experiment. The force increases linearly with progression time

until a maximum force Fp is reached, which causes failure of the

interface and rapid separation of the MCF from the bulk bone

sample.

194

6.3 Schematic of the sacrificial ionic bonding system in NCPs region

with pink indicating MCFs, the blue region defining the

extrafibrillar NCPs glue layer and red denoting mineral crystals.

Each short black dash line in NCPs layer represents an activation

volume. (A) NCPs are shown anchored to the mineral sites of the

MCF. (B) The negatively charged NCP molecules are bonded

together by positively charged divalent calcium ions. (C) MCF

pullout causes shear in the NCP layer where ionic bonds are

assumed to reversibly form process of MCFs pullout. Thus, during

plastic deformation of the interface glue region, ionic bonds fail as

negatively charged NCPs molecules are separated but reform

when the NCP molecules contact neighbouring NCP molecules

during the pullout process. The NCP bonding will reform and

break continually with the fibril pullout movement, resulting in a

large number of bonds breaking relative to a non-reforming

198

Page 28: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

27

pullout process as show in (D).

6.4 Plot of the sacrificial ionic bond reforming efficiency against

individual MCF pullout velocity speed for separation of MCFs from

the surrounding NCP region. The dashed line shows the trend of

the reforming efficiency increasing with decreasing pullout speed.

204

7.1 Schematic diagrams showing stress distribution in the pullout

fibril and surrounding bone tissue. MCFs are represented by

orange cylinders with blue extrafibrillar NCP regions existing

between the MCFs. (A) Adjacent MCFs are deformed by the shear

load transferred from NCPs whereas (B) Surrounding bone tissue

is not affected if assume NCP do not transfer load which is

contradictory to the phenomenon observed in Chapter 6.

212

7.2 Maximum pullout force required for pulling out of individual MCFs

from bone matrix plotted against fibril embedded length. Red dash

line shows the linear regression.

214

7.3 Calculated interfacial fracture energy from pullout of individual

MCFs from bone matrix is plotted with different stress transfer

parameters R/r. Error bars indicate the standard deviation of the

interfacial fracture energy values calculated from five individual

fibril pullout tests.

215

7.4 Schematic showing the propagation of a crack perpendicular to

the osteon orientation in bone leading to fibril fracture and

pullout. MCFs are represented by orange cylinders with blue

extrafibrillar NCP regions existing between the MCFs.

216

7.5 Scanning electron micrographs of fracture surfaces in antler bone

from different viewing angles. (A) Image taken with sample stage

at 0 degree tilt. (B) Image taken when sample stage was tilted by

30 degrees.

219

Page 29: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Figures

28

7.6 Distribution of fibril pullout length at antler bone fracture

surfaces.

220

7.7 (A) and (B) Scanning electron micrographs of crack growth path at

a fracture tip in antler bone [283]. Transverse crack propagates

against osteon orientation in (A) with red arrows show the crack

‘twisting’ during crack growth. (B) Cross section image of osteons

showing how ‘micro cracking’ occurs away from a crack tip. (C)

and (D) Synchrotron X-ray computed micro-tomography images of

cracks occurring in both the longitudinal and transverse

orientations of antler bone [283] with cracks shown in purple.

Central canals in osteons (hollow volume in osteons) and vascular

channels are represented in brown. Both longitudinal and

transverse directions exhibit significant cracking. However,

cracking in the transverse orientation (D) shows more failure

volume due to crack deflection and microcracking generated in the

neighbouring space near the principal crack tip [283].

225

Page 30: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Constants

29

List of Constants

Z bone beam deflection

M bending moment

L length of the limb

E Young’s modulus

I second moment of inertia

ρ density

A cross-sectional area

v velocity of sound in bone

ν Poisson’s ratio

Gc critical energy release rate or fracture energy

Kc critical stress intensity factor or fracture toughness

F tensile force

k cantilever spring constant

d cantilever deflection

r nanofibre radius

L original fibre length

ΔL elongation of the fibre

X retraction distance of the cantilever

I0 collected laser intensity

I1 and I2 reflected laser intensity

wavelength of the laser

Page 31: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

List of Constants

30

Im geometric moment of inertia

a frequency of the cosine function

b scaling constant

c decay rate of the laser intensity

I0’ normalized laser intensity

x’ cantilever displacement

ε strain

σ stress

Df fibril diameter

le fibril embedded length

Va activation volume

H activation energy

W work done to pullout the fibril

H100% and H0% activation energies for the boundary conditions of complete

bond reforming (100 %) and complete bond failure (0 %)

α sacrificial bond reforming efficiency

Fp maximum tensile force applied to the fibre before pullout

Uf, Ud and Ub, strain energy stored in fibre free length region, debonded

region and bonded region respectively

Gf fibre interfacial fracture energy

Ef and Em Young’s modulus of fibre and matrix

Af and Am Cross-sectional area of fibre and matrix

νm Poisson ratio of matrix

Wf work of friction at the debonded interface

a crack length

f MCF volume fraction

lp pullout length of the fibrils

Gc fracture energy from fibril fracture and interface failure

Page 32: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

31

Chapter 1

Introduction

1.1 Background

Many biological materials incorporate biominerals within a polymeric

framework in order to achieve a specific mechanical function. Bone is a prevalent

example of a mineralized biological material with defined mechanical functions

including transmitting the force of muscular contraction from one part of the

body to another during movement, to protect vital organs and bone marrow as

well as to enable the skeleton to maintain the shape of the body.

The material composition of bone is comprehensively understood, with soft

materials predominantly of polymeric collagen and non-collagenous proteins

acting as a framework for the formation of a harder mineral phase of

plate-shaped crystals of calcium carbonated apatite. The organization of these

elements of bone is complex but essentially based around nanofibres of collagen

Page 33: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

32

containing the mineral phase. Although bone exists in many shapes and different

anatomy locations with various internal structures, all bone consists of a certain

constituent called mineralized collagen fibrils (MCFs) which can be considered as

the building blocks of bone [1-5]. MCFs can be both clearly identifiable using high

resolution microscopy and used to form higher length scale structures in bone as

shown in Figure 1.1.

Figure 1.1 Scanning electron micrographs showing the fracture surface of equine

bone (A) and bovine bone (B) at comparable magnifications. Separated

mineralized collagen fibrils are clearly identified in each sample as hair-like

structures protruding from the bone surfaces.

The origin of the mechanical properties of bone is currently unclear, although

numerous reports have indicated a strong relationship between bone

hierarchical structure, especially the constituents of bone across different length

(A) (B)

Page 34: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

33

scales, with overall mechanical behaviour [1, 2, 4-9]. Of fundamental interest is

the toughness of bone, especially as bone material is considerably tougher than

many other natural materials despite the relatively poor mechanical properties

of constituents [3, 10-15]. However, fundamental understanding of the nanoscale

mechanisms operating in bone that provide this toughness still remain poorly

understood [16]. The composition of bone is expected to define overall

mechanical performance and over 80 % by volume of bone is occupied by MCFs

[4]. Therefore, understanding how these fibrils organize and interact with each

other in bone and, more importantly, how they behave mechanically is crucial in

studying mechanics of whole bone and its toughening mechanism.

1.2 Current Studies on Mechanics of Bone

Common equipment employed to assess the structural properties of bone are a

scanning electron microscope (SEM) and an atomic force microscope (AFM),

used particularly for their high resolution imaging capability [15, 17-22]. AFM

was initially introduced as a powerful image tool with atomic resolution [22-24]

but was later used for mechanical testing including nanoindentation and direct

mechanical testing on bone material. For example, many previous investigations

examine the mechanical properties of bone at the nanoscale using AFM

indentation techniques [25-38]. However, indentation testing is limited by the

complex stress analysis required to interpret experimental data, so can be

Page 35: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

34

considered as an indirect mechanical testing method. The indentation technique

is also suitable for measuring surface mechanics rather than fibrous samples and

is therefore not eligible for MCF mechanical testing. Only a limited number of

studies [20, 39] have used AFM for direct mechanical testing of collagen fibrils, as

shown in Figure 1.2, but no test considers mineralized fibrils from bone tissue.

Figure 1.2 Atomic force microscopy can be used to deform (but not fail)

individual collagen fibrils from bovine tendon [20]. (A) An AFM probe with glue

at its apex is used to pick up a collagen fibril from a Teflon coated surface with

the one end of the fibril fixed in a drop of glue on the surface. (B) Schematic

diagram indicating the AFM configuration, after fibril pick-up, used for

mechanical testing.

Figure 1.2 highlights how isolated collagen fibrils from bovine tendon were

mechanical tested using AFM. The stress-strain behaviour of collagen fibrils

obtained in this test was used to define fibril elastic properties but the fibril

(A) (B)

AFM probe

Glue

Collagen

fibril

AFM probe

Collagen

fibril Glue

Page 36: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

35

fracture properties including strength and ultimate strain were not measured

due to the limited strain applied to the fibril by the experimental method.

Additional research was carried out on mechanically testing collagen fibrils from

sea cucumber using a micro electro mechanical (MEMS) device [40]. The

individual collagen fibrils were tensile tested to failure with failure behaviour

recorded. However, such a test is suitable for the selection of fibrils with

relatively large diameters (usually up to microns) and is therefore not suitable

for testing MCFs from bone tissue.

Figure 1.3 Individual collagen fibrils from sea cucumber tensile tested using (A) a

MEMS device in ambient/humid environment. (B) Individual collagen fibrils

were isolated from the bulk and tensile tested between the grips of the MEMS

device [40].

(A)

(B) Collagen fibril

MEMS grips

Page 37: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

36

These two examples of direct tensile tests on collagen fibrils using AFM and

MEMS are of importance to soft tissue mechanics because it was first time the

mechanical behaviour of collagen fibrils was recorded using reliable techniques.

Critically, separation of the collagen fibrils from the parent material is required in

both experiments. Unmineralized collagen fibrils from tendon and sea cucumber

are usually packed in parallel arrays and can be easily separated by high

frequency vibration ultrasound and chemical agents as used in these studies.

However, mineralized collagen fibrils in bone are attached together with

extrafibrillar matrix and mineral that become damaged when exposed to

chemical or ultrasound techniques [19, 32, 39, 41-47]. A novel testing

methodology is therefore desired for mechanically testing MCFs without using

previous preparation techniques for unmineralized collagen fibrils, thus avoiding

potential MCF damage, to allow understanding of MCF mechanical behaviour and

their influence on whole bone mechanics.

1.3 Project Aims and Thesis Outline

1.3.1 Project Aims

This thesis details nanomechanical testing of MCF fibrous structures in bone

materials and evaluates how the MCF unit contributes to the overall fracture

behaviour of bone. In particular, little is known about how the MCFs

Page 38: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

37

mechanically contribute to bone’s surprising toughness, especially as mechanical

properties of bone constituents are considered to be relatively poor [2, 7, 9, 12,

15, 16, 36, 48-50]. The overall objective of this thesis is to address this question

based on experimentally studying the mechanical properties of bone tissue at

small scales appropriate to the MCFs in bone.

The main objectives of this project are therefore:

1. To develop a novel mechanical testing technique suitable for measuring

individual nanofibres. Such a technique requires a suitable sample

preparation method to separate individual MCFs from bone while avoiding

induced damage from the preparation. The testing technique and associated

sample preparation will form a universal mechanical testing strategy for

MCFs.

2. Apply this new technique to bone materials and measure the mechanical

properties of the individual MCFs.

3. Test the flexibility of the nanomechanical testing by studying MCFs from

different ages of bone to obtain information on aging effects on the mechanical

properties of MCFs.

Page 39: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

38

4. Evaluate the interfacial properties of the interphase region between MCFs in

bone tissue in order to understand the mechanical coupling between MCF

assemblies.

5. Examine the nanoscale fibril related fracture behaviour and its contribution to

the global fracture properties of bone.

1.3.2 Thesis Outline

In this study, a novel in situ nanomechanical testing methodology was developed

for investigating nanofibrous samples including MCFs of bone. Individual MCFs

from antler bone and mouse bones at different development stages have been

successfully tensile tested. Furthermore, the interface between MCFs and

surrounding non-collagenous proteins (NCPs) has also been mechanically

evaluated by pulling out single collagen fibrils from the bone matrix. An

analytical composite model was used to describe the overall fracture behaviour

of bone material at the fibrillar level.

The second chapter of the thesis reviews existing bone literature, with emphasis

on bone structural formation and theories developed from experimental results

describing bone’s mechanical properties at different length scales. This literature

Page 40: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

39

review will justify the importance of the constituent mechanical properties on

overall bone mechanical behaviour.

In Chapter 3, detailed sample preparation methods for isolating individual

mineralized collagen fibrils from bone are described. A novel in situ mechanical

test method using a custom built atomic force microscope (AFM) combined with

a scanning electron microscope (SEM) dual beam system is reported for

performing tensile testing individual collagen fibrils from antler bone.

The fourth chapter details measurements of the mechanical properties of

individual mineralized collagen fibrils from antler bone using the novel testing

setup described in Chapter 3. The response of the mineralized collagen fibrils are

examined in terms of their role in defining the toughness of antler through

testing collagen fibrils from the surface of fractured antler specimens.

Unprecedented insight into fracture behaviour of antler at the nanoscale is

determined with novel fracture processes found to be intimately linked to

inhomogeneous deformation at the fibrillar level.

The fifth chapter extends the mechanical testing of Chapter 4 to tensile testing of

MCFs from different ages of healthy mouse bone. Mechanical properties of MCFs

from different developmental stages of bone, where compositional changes in

Page 41: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 1. Introduction

40

whole bone are widely reported, are compared to bulk bone mechanical

behaviour with implications of MCF mechanics on different ages of bone

examined.

Chapter 6 uses in situ nanomechanical testing for pullout of MCFs from bone

material in order to evaluate the interfacial properties between MCFs and NCPs.

Interfacial behaviour in composite materials is critical in defining both

deformation and failure, and the evaluations of interfacial failure in bone at

nanometre length scales are expected to be instructive in defining whole bone

mechanics.

Chapter 7 combines the mechanical properties of MCFs and their interfaces to

describe, using analytical composite theories, the contribution of nanoscale

mechanical behaviour towards whole bone mechanical performance. Finally, the

eighth chapter summarizes the main contributions of thesis. Additional studies

beyond the thesis and future work are also examined.

Page 42: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

41

Chapter 2

Literature Review

2.1 Introduction

This chapter reviews the existing literature on bone structure and resultant

mechanical behaviour, with emphasis on the prevailing results and theories

regarding bone formation, hierarchy structure and composition of bone. The

literature review is particular important for justifying the experimental work of

exploring mechanical properties of bone at the nanoscale.

2.2 Bone Formation and Secondary Bone

Bone is a composite material consisting of various different components, as

described in Section 2.3 below, organized across different length scales to form

whole bone incorporated within a skeletal structure. Bone material is distinct

from other synthetic materials because of its ability to change and be in a

Page 43: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

42

constantly changing structural state. While bones can be found from many

different locations within a skeletal structure, long bones such as femur have

been perhaps the most extensively investigated and provide the standard models

used to describe bone structure. Before discussing the composition and structure

of bone, one first needs to understand how the bone tissue is constructed

hierarchically across different length scales during the bone formation process.

There are two main stages of the bone formation: primary and secondary

ossification (known synonymously as osteogenesis) [12, 51, 52] which mainly

involves three cellular activities i) the chondrocytes which produces cartilage for

ossification in early stage ii) the osteoblasts which secrets collagen and form

bone matrix that then differentiate into osteocytes embedded in bone tissue and

iii) the osteoclasts which dissolve and absorb old bone tissue for remodelling. An

example of growth process of long bone was demonstrated in Figure 2.1.

Figure 2.1 Schematic diagram showing the stages of endochondral ossification

[53].

Page 44: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

43

Bone formation is initiated by mesenchymal condensation where intercellular

substances aggregate and condensed to form a cartilage model of the bone to be

formed [54, 55]. The cartilage model grows in length and thickness by the

continued cell division of chondrocytes, which is accompanied by further

secretion of extracellular matrix to form cartilage for ossification [51, 56].

Following the appearance of cartilage model, a primary centre of ossification is

formed by cartilage matrix mineralization, osteoclast activities and

vascularization. The formation of bone then occurs at the epiphysial cartilage in

primary ossification centre, which consists of ground substance (containing

chondroitin-4-sulphate, chondroitin-6-sulphatc, and keratosulphate [57]) and

loosely packed collagen fibrils (having diameter of 10-20 nm) [52, 58]. The

matrix vesicles embedded in the epiphysial cartilage deliver mineral ions to the

mineralization front [59-63] as shown in Figure 2.2.

A “woven” bone microstructure is formed during this rapid and unorganized

mineralization in which the mineral component does not form in close

association with the collagen and no regulated structures are formed. The

collagen fibrils in this process are found to be too narrow for the mineral to

deposit within them [58]. Instead, an extrafibrillar mineralization can be

observed during this initial bone formation process, which indicates that the

collagen does not play an appreciable role in directing the mineralization. The

bone material produced during this process is called primary bone, which is

usually found to exist shortly in the early development stage. The primary bone

Page 45: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

44

can be also found in later development stages when skeleton is healing from

trauma. During healing of bone fracture, the woven bone forms after collagen is

secreted by osteoblasts. The mineralization process is similar to endochondral

ossification. Mineral ions are delivered from matrix vesicles in sounding

extracellular matrix to the newly formed collagen fibrils.

Figure 2.2 Optical microscope image showing ossification from epiphysial

cartilage to mineralized collagen/woven bone [64].

Primary bone is produced quickly to define the shape of whole bone that

provides initial mechanical support, which is crucial in the early stage of bone

development and in the recovery of bone fracture [53]. However, the collagen

Chondrocytes

Mineralization

front

Epiphysial

cartilage

Woven Bone

Page 46: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

45

fibrils constituents in primary bone contain a relatively low mineral content and

are not organized in an efficient way to provide effective mechanical properties.

The relatively poor mechanical properties of primary bone therefore require

changes in this bone material in order to provide more effective mechanical

function. Thus, secondary ossification occurs in order to provide constituent

ordering and improved mechanical behaviour in bone at specific centres as

shown in Figure 2.1.

The initiation of secondary ossification occurs at each end of long bone where the

mesenchyme and blood vessels are developed in a similar process to primary

ossification. The cartilage between the primary and secondary ossification

centres is called the epiphyseal plate, and it continues to form new cartilage that

is replaced by bone, a process that results in an increase in length of the bone as

shown in Figure 2.1. Growth continues until the cartilage in the plate is complete

replaced by bone and the bone growth stops. Meanwhile, secondary ossification

proceeds by the existing primary bone tissue becoming dissolved by osteoclasts

and absorbed into the extracellular matrix, which is then used by osteoblast to

form a more organized and optimal structure. The changes in the primary bone

material structure, referred to as bone remodelling, form bone material called

secondary bone [65]. The shape of whole bone is therefore defined by the

activities of osteoclasts and osteoblasts.

The collagen fibrils formed in secondary ossification by osteoblasts have a much

larger diameter (70-100 nm) compared to those produced in primary woven

Page 47: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

46

bone by chondrocytes [52]. Sufficient internal space is available in collagen fibrils

from secondary bone to allow mineralization within fibrils, defined as

intrafibrillar mineralization. The structure of collagen fibrils in primary and

secondary bone is therefore different due to the mineralization process. While

collagen fibrils in primary bone contain extrafibrillar mineral between the

collagen fibrils, mineralization of fibrils in secondary bone occurs initially in the

intrafibrillar spaces before extending to extrafibrillar mineralization.

2.3 Hierarchy Structure of Bone

During the secondary bone formation, the primary woven bone tissue is

remodelled into a more optimal and organised structure [65]. The collagen fibrils

produced in secondary ossification are highly organized and assembled into

close packed and parallel lamellar structures used to form larger structures and

eventually a complex hierarchy. Thus, discussions on the structural hierarchy of

bone typically refer to the complex architectures found in secondary bone [66].

In a seminal review of bone structure provided by Weiner et al. [1], the whole

structure of bone has been considered as seven levels of hierarchy (Figure 2.3)

starting with nano-sized platelets of hydroxyapatite (HA) that are oriented and

aligned within self-assembled collagen fibrils. These fibrils are layered in parallel

arrangement referred to as lamellae which are arranged concentrically around

blood vessels to form osteons. Finally, the osteons are either packed densely into

compact bone or comprise a trabecular network of microporous bone, referred

Page 48: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

47

to as spongy or cancellous bone. Skeletal whole bone therefore contains both

compact and cancellous bone tissue.

Figure 2.3. The hierarchical levels of structure found in secondary osteonal bone,

as demonstrated by Weiner and Wagner [1].

Level 1: Major Components

Level 2: Mineralized Collagen Fibrils

Level 3: Fibril Array

Level 4: Fibril Array Pattern

Level 5: Haversian System

Level 6: Cancellous & Cortical Bone

Level 7: Whole Bone

Page 49: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

48

2.3.1 Level 1: Major Components of Bone

Level 1 of the seven level hierarchy of bone describes the major components of

bone materials: the dahllite (carbonated apatite) crystals, type I collagen fibrils,

and water. The carbonated apatite (Ca5(PO4,CO3)3(OH)) is the only inorganic

component found in mature bone [1] and form crystallites with a flatted

plate-like shape, with the average length and width of 50 25 nm (Figure 2.3;

Level 1) [51, 67] which are believed to be the smallest biologically formed

crystals known [51].

The apatite crystal platelets are extremely thin with a remarkably uniform

thickness of only few nanometres. Transmission electron microscopy (TEM) and

small angle X-ray scattering (SAXS) studies show that the thickness of the apatite

crystals varies from 1.5 nm for mineralized tendon up to about 4.0 nm for some

mature bone types [68, 69]. Due to the small size of the apatite crystals found

in bone, the surface atomic structure as well as the mechanical properties of

single crystals has not been experimentally measured using reliable methods.

Atomic force microscopy (AFM) studs of synthetic apatite show that the surface

is highly ordered and matches the structure of bulk equivalents [70].

Interestingly, bulk dahllite forms symmetrical hexagonal crystal structures

reflecting unit cell symmetry whereas dahllite in bone has a unique plate shape.

Page 50: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

49

The organic constituent of bone consists of type I collagen used to construct

collagen fibrils but also other diverse non-collagenous proteins (NCPs) [71]. In

bone formation, NCPs are found to exist near the mineralization front between

collagen fibrils and are also highly charged from an abundance of carboxylate

groups from aspartic and glutamic acid residues, as well as phosphate from

phosphoserine [51, 56, 72-74]. Although in low concentration, the NCPs are

believed to perform an important function in the mineralization process of bone

[51, 52, 75].

2.3.2 Level 2: Mineralized Collagen Fibrils (MCFs)

Type I collagen accounts for 90 % of the organic/protein components by volume

within bone and is therefore expected to be important in defining mechanical

properties of the bulk bone material. Type I collagen is formed as a fibrous

structure with the fibrils in bone having diameter of 80–100 nm as measured by

TEM (Figure 2.3; Level 1). The lengths of individual MCFs remain unknown as the

collagen fibrils tend to merge with neighbouring fibrils [76, 77]. Level 2 describes

the mineralized collagen fibril (MCF) building block, which is defined by Weiner

and Wagner [1, 2] as the smallest discrete basic unit used to construct bone

materials (Figure 2.3; Level 2). The MCFs are therefore critical in defining the

mechanical properties of bone, which is also the objective of this thesis.

The basic collagen fibril in bone is secreted by osteoblasts in the growth of bone.

As shown in Figure 2.4, every individual collagen molecule is made up of three

polypeptide chains about 1000 amino acids long [1] wound together in a triple

Page 51: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

50

helix. A triple-helical molecule is thus cylindrically shaped, with an average

diameter of about 1.5 nm, and lengths of 300 nm [51, 78, 79].

Figure 2.4 Schematic diagram showing structure of a collage fibrils composed of

self-assembled tropocollagen molecules [1].

Collagen fibrils are formed from the self-assembled tropocollagen molecules. In a

single collagen fibril, tropocollagen molecules are parallel to long axis of fibril,

but their ends are separated by holes of about 35 nm as shown in Figure 2.5 (A)

and neighbouring triple-helical molecules are staggered by 67 nm [80]. This

quarter-stagger arrangement [31] shown in Figure 2.5 produces a characteristic

and repeating gap region between the tropocollagen molecules that is occupied

by water molecules. The mineralization process of bone replaces the water in gap

regions within the collagen fibrils with plate-shaped mineral crystals [30] as

shown in Figure 2.5. The formation of the mineral within the collagen fibrils is

directed by the internal gap regions produced from the quarter-stagger

arrangement [80, 81]. Single mineralized collagen fibrils can be therefore

considered as a ceramic phase reinforced polymer system.

Collagen fibril

Collagen molecule

Polypeptide chains

Page 52: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

51

Figure 2.5 (A) 3-D organization of apatite minerals inside a single collagen fibril.

Collagen molecules (represented by red rods) aggregate into a quarter-staggered

pattern described by Hodge and Petruska [82]. Calcium phosphate (blue plates)

nucleates in the gaps between the collagen molecules. (B) TEM image of

mineralized collagen fibrils from mineralized turkey tendon showing the

characteristic quarter-stagger pattern [83].

The mineral formed within the collagen fibril is typically of relative small

dimensions and is not normally expected to be thermodynamically stable if

existing outside of the organic matrix of the collagen fibril [44, 84-86]. The

distribution of mineral components at different sites within the collagen network

has been studied by Martin et al. [87] for canine whole bone and indicated 58%

of mineral is intrafibrillar, 14% extrafibrillar, and 28% found at the gap regions

1.5 nm

300 nm

35 nm

67 nm

Page 53: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

52

between the ends of the tropocollagen molecules.

2.3.3 Level 3 Fibril Array

The third structural hierarchy level in bone denotes the fibril array incorporating

packing of mineralized collagen fibrils. In most of the bone materials, MCFs are

close packed into bundles or arrays along the long direction of their lengths [88]

(Figure 2.3; Level 3). These fibril bundles are mostly composed of MCFs that

sometimes merge with neighbouring fibril bundles [76]. Detailed information on

how the MCFs are organized within fibril arrays still remains poorly understood

and is controversial. Generally, two different fibril structural organizations are

proposed in the literature [2, 89] as shown in Figure 2.6. Literature proposes that

the two different arrays are formed in different bones to optimise the stress

transfer in bones under different loading conditions [89-91]. mise the stress

transfer in bones under different loading conditions [89-91].

Figure 2.6 Schematic illustration of MCFs containing mineral plates (not drawn to

scale, MCFs are shown in cylinders with black mineral crystals embedded)

showing (A) an arrangement of mineralized collagen fibrils aligned both with

(A) (B)

Page 54: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

53

respect to crystal layers and fibril axes (orthotropic symmetry). (B) Arrangement

of mineralized collagen fibrils with only the fibril axes aligned (transversal

isotropy) [1].

2.3.4 Level 4: Fibril Array Pattern

At level 4, the fibril array organizational patterns vary to produce a diverse range

of structural types in bone tissues. As shown in Figure 2.7, organization of fibril

arrays into a higher structure are generally divided into four categories: arrays of

parallel fibrils, woven fibre structure, plywood-like structures and radial fibril

arrays in different bone tissues [2, 49, 91].

In mature secondary lamellar bone, the plywood-like structure of Figure 2.7 (C)

is the most common structure used to describe the organization of MCFs. Weiner

et al made a detailed study of collagen organization in the lamellar bone from

mouse limb bones [91, 92] and observed parallel layers of fibrils, each with a

successive tilted angle of around 30° from layer to layer. A model was therefore

proposed in which a lamellar unit is composed of five such sub-layers as shown

in Figure 2.7 (C). Because only five and not six sub-layers oriented progressively

every ± 30°, the lamellar unit structure is not symmetrical [93]. Variations in the

thicknesses of the sub-layers occur for bones from different species, rather than

the tilted angle between layers.

Page 55: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

54

Figure 2.7 Electron micrographs and schematics highlighting the four most

common fibril array pattern organizations in bone. SEM micrographs of fractured

surfaces and schematic illustrations (not drawn to scale) of the basic

organizational motifs indicate: (A) Array of parallel fibrils of mineralized turkey

(A)

(B)

(C)

(D)

Page 56: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

55

tendon (scale: 0.1 mm) and schematic illustration showing the localized

orthotropic symmetry of a fibril bundle. (B) Woven fibre structure of the outer

layer of a 19 week old human fetus femur [94].

The schematic illustration shows fibril bundles with varying fibril diameters

arranged in random orientations. (C) Plywood-like structure presenting fibril

organization in lamellar bone from the fracture surface of a baboon tibia, with

prominent fourth (large arrowhead) and fifth (small arrowhead) sub-layers [91,

95]. The corresponding schematic illustration shows the five sub-layer model

described in [91] with sub-layers one (right hand side), two, and three arbitrarily

composed of one fibril layer each, whereas sub-layers four and five are composed

of four fibril layers each. Note that the fibrils in each layer are rotated relative to

their neighbours (depicted by the change in direction of the ellipsoid

cross-section), following the rotated plywood model [89]. (D) Radial fibril arrays

from human dentin fractured roughly parallel to the pulp cavity surface. The

tubules (holes) are surrounded by collagen fibrils that are approximately in one

plane. The schematic illustration of the fibril bundles highlights the arranged in a

plane perpendicular to the tubule long axis with no obvious preferred

orientation within the plane [1, 2].

2.3.5 Level 5: Osteonal/Haversian System

At level 5, the bundles of mineralized collagen fibrils are organized into a higher

length scale cylindrical structure unit known as the osteon as shown in Figure 2.3.

Osteons are formed during remodelling of primary bone into secondary bone and

Page 57: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

56

are initiated by osteoclasts removing primary bone material in a dissolution

process as they migrate through the material. Such dissolution forms channels

that are subsequently filled by osteoblasts. The osteoblasts excrete the more

regularly organized layers of lamellar bone within the channel until only a

narrow channel in the middle of the osteon remains to allow movement of blood

vessels within the secondary bone [48] (Figure 2.3; Level 5). Other smaller

capillary-like features (canaliculi) are also built into the structures and connect

trapped osteoblasts, referred to as osteocytes. These osteonal structures are also

called Haversian systems named after John Havers who discovered the lamellae

structure in 1691 [96]. Primary bone is sometimes formed of osteons (primary

osteon) but the slower formation of osteons relative to woven bone typically

requires remodelling of woven bone to give secondary osteons, the most

common structures found in mature secondary bones and described in the bone

formation section 2.2 in this thesis.

2.3.6 Level 6: Architecture Level – Compact and

Cancellous Bone

The level 6 architectural level identifies structural features over length scales of

hundreds of microns and essentially describes the distinct regions of compact

and cancellous bone. Osteons are formed into dense, solid bone tissue called

compact bone or lighter, porous bone tissue called spongy bone or cancellous

bone, which both can be found in secondary lamellar bones as shown in Figure

2.8.

Page 58: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

57

Figure 2.8 (A) Micro-anatomy structure of a human femur head with compact

and cancellous bone demonstrated. (B) Schematic diagram showing the structure

of cortical bone [97].

2.3.7 Level 7: Whole Bone

At the highest scale level, whole bone represents the overall shape of the bone. In

most of mammal animal’s skeletal system, bones exist in different shapes and

sizes but are classified into three common types: long bone, flat bone, irregular

bone and short bone as shown in Figure 2.9.

The long bones have a greater length comparing to width and usually exist in

arms and legs to support motion. Flat bones are flattened and usually have a

protective function such as the bones of the skull and sternum, which protect the

internal organs from external impact. Short bones have relatively equal length

and width and usually exist at joints such as carpal bones to facilitate motion.

(A) (B)

Page 59: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

58

Irregular bones possess little symmetry such as sphenoid bone and vertebrae

and typically function as connective parts in the skeletal system.

Figure 2.9 Schematic diagram showing a range of different bone shapes found in

the human skeleton [98].

The spatial organization in whole bone is defined by the microstructural units

from the architectural level 6. These two levels are far larger than the

mineralized collagen fibrils studied in this report and therefore are not discussed

in detail. There is a considerable body of literature studying bone at large scales

over decades such as the comprehensive reviews provided by Currey [99] and

Thompson [55].

Long Bone

(femur or thighbone)

Irregular Bone

(sphenoid bone

from skull)

Short Bone

(carpal or wrist bone)

Flat Bone

(parietal bone from

roof of skull)

Page 60: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

59

To conclude, all these 7 levels of bone are composed of structural differences in

one magnitude between the subsequent levels, bridging from the ultrastructural

component level to the whole bone [27]. In order to expedite a complete

understanding of bone material, individual structural components and features

at different levels contribute to the overall mechanical behaviour of bone. By

examining bone at different hierarchical levels, a comparison between structure

and material properties between levels can be established. However, each

structural hierarchy level depends on the structural features of the levels below

to provide function and structural support. Thus, the lowest discrete unit of MCFs

must be critical in defining higher order mechanical behaviour and the overall

mechanical response of bone, as will be discussed in the next section.

2.4 Bone’s Structure and Mechanical Properties

The earliest study on mechanical properties of bone can be perhaps traced back

to the beginning of anatomy and orthopaedic surgery. However, initial studies

highlighted the lack of interaction between biologists and mechanics researchers.

In a symposium “The Mechanical Properties of Biological Materials”, Vincent and

Currey noted: “…it was clear that the biologists who work on the mechanical

properties of materials need much more to call upon the materials scientists for

guidance on what is known already about the phenomena they are studying; on the

other hand it was equally clear that the materials scientists had little idea of the

Page 61: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

60

richness of biological diversity and of the pervasive ability of natural selection to

provide the optimum solutions to very complex problems.”[36] Since the initiation

of the symposium in 1980, a vast number of studies on mechanical behaviour of

bone have been published and significant understanding achieved. This

understanding was driven not by deep theoretical insights, but by the

development of multidisciplinary subjects and the advances in materials testing

techniques.

The mechanical behaviour of bone at macro scales is highly affected by the

mechanical properties of its constituents [15]. However, the material properties

of bone structure are strongly dependent on the scale of observation. Recent

experimental and theoretical studies have provided new insight into the

mechanics of bone, most notably through the use of advanced instrumentation

that has permitted the examination of bone properties at ever-decreasing length

scales, e.g. transmission electron [56, 100] and X-ray microscopy [69, 101],

atomic force microscopy [17, 20, 102], and Raman spectroscopy [103, 104], as

well as bottom-up multiscale simulation modelling [41, 105, 106].

In this section, the previous studies on mechanical properties of bone materials

at different length levels using various methods are reviewed. As discussed in

Section 2.3, bone is a composite material that exists on several hierarchical levels.

Page 62: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

61

In the discussion of the structure of bone, hierarchy was reviewed from lowest

length scale to whole bone, which roughly coincides with the bone development

process i.e. bone is produced at small lengths scales initially but organization

builds up larger length scale features. In this section, the review of studies on

mechanical properties of bone starts from the highest level: the whole bone,

because organizing the structural constitutes in bone tissue at different length

scales is adapted due to the overall requirement of the mechanical functions of

whole bone.

2.4.1 Level 7: Whole bone

The whole bone level is the largest length scale level of bone and represents the

summation of the structural and material properties of all length scales of bone.

At this whole bone level, the bone or skeletal system provides the mechanical

functions of structural support and protection to internal organs as well as aiding

motions. The structural geometry and organization of bone are adapted or

‘coincident’ with bone’s mechanical functions. Interaction between whole bone

and other constituents of the body may include tendons, ligaments, muscles and

other bones.

The shapes of whole bone are various according to the bone’s mechanical

functions. While it is almost impossible to examine the physiological loading

Page 63: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

62

conditions for all types of bones with different shapes, identifying the complete

loading conditions for even a single type of bone with a certain shape is still

difficult. This section will mainly discuss the mechanics of the most common

bone we can find in skeletal system, which is long bone. The long bones are

relatively simple in shape and classical engineering principles have been applied

to understand the loading on such particular bone types.

Long bones provide common features for other bone types including a hollow

internal structure with a relatively thick wall defining the whole bone’s external

surface. Long bone is specifically expanded at each of its ends, with a cancellous

bone structure found within these ends. The hollow structure of long bone is

derived from mechanical function and the so called ‘minimum mass’ requirement

that exists widely in nature [12]. In minimum mass analysis, the aim is to build a

structure to perform a certain mechanical function with minimum mass.

A simple example is a limb bone loaded as a beam as shown in Figure 2.10. The

limb bone acts primarily to exert loads from the environment and therefore

needs to be stiff to withstand large bending moments. The distortion caused by

bending on the bone must not exceed a limit. In evolution theory, natural

selection will favour animals that function with the highest efficiency. This

efficiency is when the function can be performed by such a structure produced

Page 64: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

63

with the minimized cost to animals in terms of metabolic energy consumed and

the time spent. Therefore, a mechanically optimized hollow structure was

developed in most of the bone structures such as long bone.

Figure 2.10 A possible design criterion for a limb bone. The long bone is initially

straight with a length of L.

A mechanical analysis can be presented to highlight how the structure of bones is

optimized towards a mechanical function. For the limb bone shown in Figure

2.10, the maximum deflection can be calculated as:

𝑍 =𝑀𝐿2

8𝐸𝐼 Equation 2.1

where Z is the maximum deflection of bone, M the bending moment applied, L is

the length of the limb, E is the Young’s modulus of the bone and I is the second

moment of inertia.

L

M M

Page 65: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

64

In Equation 2.1, the geometry and composition of the bone provide fixed values

of L and E respectively. An external load is applied to the limb generates a

bending moment M. To make sure the deflection of bone Z does not exceed the

criteria value to cause failure, a larger second moment of area I is preferred. The

parameter I is typically defined for a beam with a round cross section in pure

bending, as shown in Figure 2.11, as:

𝐼 =∑𝑦2𝛿𝑎𝑟𝑒𝑎 Equation 2.2

It is noticed that a larger I can be achieved by increasing y, indicating that the

simplest way of controlling deflection of beam under bending is to increase the

beam cross section. However this will cause a corresponding increase in the total

mass of the beam, which would therefore require more metabolic energy to

produce the limb and is not preferable.

By assuming the beam density is ρ with a cross area of A, the ratio of the mass to

second moment of area is:

𝜌𝐿𝐴

𝐼=𝜌𝐿∑𝛿𝑎𝑟𝑒𝑎∑𝑦2𝛿𝑎𝑟𝑒𝑎

Equation 2.3

Equation 2.3 suggests that a minimization of the beam mass and maximization of

Page 66: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

65

the second moment of area of the beam for a fixed I, L and ρ can be achieved if the

ratio ∑𝛿𝑎𝑟𝑒𝑎

∑𝑦2𝛿𝑎𝑟𝑒𝑎⁄ is minimized. This condition is achievable if the beam

mass is distributed as far as possible from the neutral axis

Figure 2.11 (A) A beam of original length L under an external loading condition

will (B) bend, causing a beam length change above and below the neutral plane.

(C) Cross section of the beam showing the second moment of area will be

different if loaded at different location on the cross section. δarea is the unit area

at a distance y from the neutral plane.

L

L+∆L

L-∆L Neutral plane

y

δarea

Neutral axis

(A)

(B)

(C)

Page 67: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

66

To minimize the mass required to limit the deflection, bone uses a small cross

section area and this area should be as far from the central neutral plane as

possible. A hollow structure is therefore formed in bone materials that are

expected to resist bending where the mass is far from the centre of mass. Thus

the shape of whole bones can be insightful in illustrating the optimization of

bone material occurring for mechanical function.

2.4.2 Level 6: Architecture Level

a. Compact Bone

As the compact bone carries most of the mechanical loading form environment,

many researches on mechanical properties of bone focus on the compact bone

tissue. Several mechanical properties of compact bone have been studied

including elastic properties (mainly Young’s modulus), strength (in different

loading directions), fracture behaviour (including mechanisms and fracture

toughness), creep rupture, and fatigue.

Bone is a type of anisotropic composite due to the orientation of various

components in structure. Different loading conditions in mechanical testing have

a considerable effect on the measured mechanical properties of bone. The highly

organized structure of bone has been discussed in Section 2.3 and highlights the

Page 68: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

67

orientation of constituents at different hierarchy levels in bone. A central feature

of compact bone is the cylindrical shape of the osteons parallel to the bone long

direction. The osteons containing calcium phosphate crystals are stiffer and

stronger than the extracellular matrix existing between osteons [3, 99]. Hence if

load is applied from different directions as shown in Figure 2.12, the stress will

be distributed in osteons and interface regions in different ways.

When load is applied along the longitudinal axis, the stiffer osteons carry the load,

resulting in bone exhibiting a maximum Young’s modulus and strength in this

loading direction. Conversely, applying load on transversal direction will make

osteon interface carry load directly which leads to a lower modulus and strength.

Figure 2.12 Loading bone in different directions. Brown cylinders represent

osteons while the interface region is shown in yellow (not to scale).

Longitudinal

Transversal

Lo

ng

dir

ecti

on

of

bo

ne

Page 69: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

68

In addition, the mechanical behaviour of bone, and indeed other biological

materials, is affected by the testing environment. Therefore, the loading direction

(such as longitudinal, circumferential and radial directions to the long axis of

long bone as shown in Figure 2.13), the humidity of the specimen and the strain

rate will influence the mechanical properties of bone and care must be taken in

testing under physiological conditions. The testing condition will be carefully

examined in the following discussion on different mechanical properties of bone.

Figure 2.13 Schematic diagram showing three loading directions in mechanical

test of bone material.

Elastic properties

The elastic properties of bone material are somewhat ambiguous due to its

inherent viscoelastic behaviour. However, two common methods are used to

measure the Young’s modulus of bone. The first is direct loading of a bone sample

Bo

ne

lon

g ax

is

Longitudinal Circumferential Radial

Page 70: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

69

to produce a stress-strain plot as shown in Figure 2.14. The apparent Young’s

modulus of the bone sample is found by taking the tangential value to the

recorded data in the linear region of the stress strain curve at low strain values.

Figure 2.14 A typical stress-strain curve of cortical bone recorded in tensile test

on human bone along longitudinal axis. The black line shows the linear response

at small strain, with the slope defining the bone’s Young’s modulus. Yield stress

and failure strain of the bone were also recorded as shown [107].

The second method of measuring elastic behaviour is through recording the

velocity of a sound wave (usually using ultrasonic sound wave) in bone [12]. The

sound velocity 𝑣 = √𝐸 𝜌⁄ , where E is the Young’s modulus of bone and is the

density of the media (which is bone in this case). However, the simple calculation

Strength

Failure

strain

Strain (%)

Stre

ss (

MP

a)

Page 71: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

70

of Young’s modulus from recording the velocity of sound is only valid for

isotropic materials, with the more complex structural anisotropy found in bone

[12] making such calculations difficult.

Thus, direct mechanical testing of bone samples is a relatively straight-forward

method with several advantages. In particular, the Young’s modulus of bone can

be easily tested under different environmental situations. Both compact and

cancellous bone can be tested without considering their structural

characteristics i.e. consideration of bone as a continuum. Calculation of the

Young’s modulus from velocity of sound measurements is indirect but can

provide the information on derivation of all the stiffness coefficients and applied

to specimens with complexe shapes or even in vivo [11, 108, 109]. A previous

study compared elastic properties measured using ultrasonic and mechanical

testing using applied loading methods on bovine bone specimens [110]. Results

indicated ultrasonic testing values correlated with indirect mechanical testing of

bovine bone selected from different anatomy locations. However, ultrasonic

testing showed significantly higher elastic property values than the direct

mechanical testing.

A considerable amount of data has been recorded for the mechanical properties

of bone. Table 2.1 gives the values of Young’s moduli from different specimens of

Page 72: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

71

human, canine and bovine bones by direct mechanical testing and the ultrasonic

methods [109, 111]. In later research on bone specimens from bovine femur long

bone, the mechanical properties of specimens were found to be higher along the

long axis of the bone, which was believed to be associated with the bone function

as a whole [110]. The Young’s modulus measured along the bone’s long axis is

about 1.6 to 2.4 times higher than the modulus measured in the transverse

direction. The variation in Young’s modulus with orientation has been validated

by numerous researchers [11, 112-114].

Table 2.1 Young’s moduli of human and bovine bone specimens (in GPa).

Reproduced from data in [12, 109, 111]. Subscripts 1, 2 and 3 represent the

radial, circumferential and longitudinal directions relative to the long axis of the

bone.

Asbman et al.[111] Reilly and Burstein [109]

Bovine

Canine Human Human Haversian Fibrolamellar

Ultrasound Direct mechanical test

Tension Compression Tension Compression Tension

E1 12.8 12.0 12.8 11.7 10.4 10.1 11.0

3.0(25) 1.01(5) 1.6(5) 1.8(8) 0.17(25)

E2 15.6 13.4 12.8 11.7 10.4 10.1 11.0

3.0(25) 1.01(5) 1.6(5) 1.8(8) 0.17(25)

E3 20.1 20.0 17.7 18.2 23.1 22.3 26.5

3.6(38) 0.85(4) 3.2(3) 4.6(5) 5.4(6)

G12 4.7 4.5

G13 5.7 5.6 3.3 3.6 5.1

Page 73: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

72

0.42(10) 0.25(22) 0.39(6)

G23 6.7 6.2 3.3 3.6 5.1

0.42(10) 0.25(22) 0.39(6)

12 0.28 0.38 0.53 0.63 0.51 0.51 0.63

0.25(24) 0.20(5) 0.24(5) 0.12(8) 0.23(6)

13 0.29 0.22 0.41 0.38 0.29 0.40 0.41

0.15(26) 0.15(4) 0.08(3) 0.21(5) 0.23(10)

23 0.26 0.24

21 0.37 0.42 0.53 0.63 0.51 0.51 0.63

0.25(24) 0.20(5) 0.24(5) 0.12(8) 0.23(6)

31 0.45 0.37 0.41 0.38 0.29 0.40 0.41

0.15(26) 0.15(4) 0.08(3) 0.21(5) 0.23(10)

32 0.34 0.35

As bone has some degree of viscoelasticity from its polymeric component

collagen, the strain rate can influence the elastic properties but the effect is

relatively small as the Young’s modulus is mainly affected by the presence of the

hard mineral phase [11, 79, 115, 116]. Research comparing bone mechanically

tested in tension under different strain rates highlighted a relatively poor

dependence on Young’s modulus [108, 115, 116]. Specifically, a thousand times

increase in loading strain rate caused a 40% increase the Young’s modulus of

bone [115]. However, the strain rate used in these previous investigations is

much higher that the loading frequency on bone in physiological condition.

Ultrasonic testing essentially provides deformation of bone at loading rates equal

to the velocity of sound. Therefore, bone testing using the relatively high loading

rates of ultrasonic methods will give higher elastic modulus values recorded in

Page 74: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

73

Table 2.1 compared to direct mechanical testing of bone under typically

quasi-static loading rates.

Strength

The strength of bone is typically measured using direct mechanical testing. The

stress-strain plots shown in Figure 2.14 used to define the elastic modulus can

give the maximum load required to fail the bone material. As with elastic

properties, bone strength is influenced by the direction of the testing relative to

bone structure and loading rate. The strength of bone material recorded in the

literature is summarized in Table 2.2. The results in Table 2.2 indicate significant

differences in bone strength depending on the orientation of the test i.e.

longitudinal, circumferential and radial as shown in Figure 2.13. In particular, the

tensile strength and failure strain of bone material along the longitudinal

orientation is significantly higher than in the circumferential direction. These

observations correlate with mechanical testing on baboon bones indicating four

times higher strength (289 MPa vs. 71 MPa) in the longitude direction compared

to the circumferential [113]. Baboon bone also exhibited a 3 fold and 16 fold

increase in failure strain and work to fracture in the longitudinal direction

loading when compared to the transverse direction [61].

Page 75: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

74

Table 2.2 Strength of human and bovine bones (in MPa) summarizing results

found in previous studies [11, 12, 109, 112, 115].

Bov

ine

Fib

rola

mel

lar

Ra

dia

l 3

0

3.2

(6)

0.2

0.1

(6)

64

7

(1

2)

Cir

cum

fere

nti

al

55

9(3

1)

0.7

0.1

3(3

1)

Lo

ng

itu

din

al

16

7

8.8

(6)

15

6

7.9

(6)

3.3

0.4

9(6

)

Ha

vers

ian

Ra

dia

l

39

4.7

(6)

0.7

0.2

(6)

19

0

18

(5)

7.2

1.4

(5)

70

9

(

7)

Cir

cum

fere

nti

al

54

5.8

(4)

0.7

0.4

(4)

17

1

25

(8)

42

1(8

)

Lo

ng

itu

din

al

15

0

11

(10

)

14

1

12

(10

)

2.0

0.5

(10

)

27

2

3.3

(3)

1.6

0.1

5(3

)

Hu

ma

n

Ha

vers

ian

Cir

cum

fere

nti

al

53

10

.7(2

0)

0.7

0.1

4(2

0)

13

1

20

.7(8

)

5.0

1.1

(8)

67

3

.5

(12

)

Lo

ng

itu

din

al

13

3

15

.6(2

1)

11

4

7.1

(21

)

0.0

31

0.6

(21

)

20

5

17

.3(2

0)

1.9

0.3

(20

)

Te

nsi

on

Str

en

gth

Yie

ld S

tre

ss

Ult

ima

te

Str

ain

%

Co

mp

ress

ion

Str

en

gth

Ult

ima

te

Str

ain

Sh

ea

r

Str

en

gth

Page 76: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

75

Increases in tensile strength and failure strain along the longitudinal axis have

been recorded in a number of additional studies [87, 109, 112]. The energy

required to fracture the bone sample in the longitudinal direction was also

shown to be 16 times higher than transverse direction, which coincided with

physiological loading conditions of bone as most of the bones carry loading along

the longitudinal axis [16, 117, 118]. For example, limb bone discussed in Section

2.4.1 has a main function of bearing a bending moment applied either in tension

or compression along the longitudinal axis. Strain dependent strength

additionally exists, with a higher tensile strength and higher strain rate recorded

in some studies [10, 108, 117-119].

Fracture Mechanics and Toughening Mechanisms

The fracture behaviour of bone and the resultant toughness requires

consideration of fracture mechanics. The first concept of fracture mechanics was

developed in 1920 by Griffith [120] to describe how a crack would propagate in a

material under external loading. The description of crack propagation in the

material used energy considerations to determine if the stored strain energy

could be released to propagate the crack by the creation of new surface area.

Load was applied to a sample containing an artificial pre-crack until failure. A

crack will extend in a solid material when the energy released from the material

Page 77: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

76

as the crack grows exceeds the energy required to propagate the crack. The

critical strain energy rate describes the energy dissipated during fracture per

unit of newly created fracture surface area. The crack will grow when the stored

energy release rate G is greater than or equal to a critical value Gc which is an

inherent mechanical property of material. The critical energy release rate Gc is

also called fracture energy and will be used in this thesis. Tough materials that

are characterized by a resistance to crack propagation have large Gc values. The

shape of the crack has provoked further analysis to consider crack geometry and

dimensions, and how these parameters influence crack propagation within the

material using stress analysis. Specifically, a crack will propagate when the stress

at the crack tip reaches a critical value that overcomes the cohesive strength of

the atoms just ahead of the crack. This critical value is also an inherent property

of materials, which is defined as the critical stress intensity factor Kc. For

materials with linear elastic behaviour, the critical stress intensity factor is

related to the critical strain energy release rate by:

𝐺c =𝐾c2(1 − 𝜈2)

𝐸 Equation 2.3

where is Poisson’s ratio and E is Young’s modulus. The application of fracture

mechanics to solid materials has been extensively established but is less

developed for bone due to the structural complexity where cracks may propagate

Page 78: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

77

at different length scales. The applicability of fracture mechanics to bone is

therefore limited due to the linear elastic behaviour assumed by fracture

mechanics that is not applicable in the rate-dependent viscoelastic behaviour of

bone [16].

The fracture behaviour of bone can be characterized by Gc or Kc and these values

are often calculated experimentally. However, a more direct method for

calculating parameters to describe the failure of bone uses a ‘fracture toughness’

evaluation. The method for determining fracture toughness involves deformation

of bone in order to record a load-displacement curve until failure. Deformation of

bone is typically achieved using a three-point bending method as shown in

Figure 2.15 (A). A notch is introduced at the mid-point of the bone and a crack

propagates during loading. The area under the load-displacement curve indicates

how much work is consumed in breaking the specimen into two as shown in

Figure 2.15 (B). The energy required to break the specimen is considerably larger

in bone that most other single phase materials [15], which can be explained

mechanistically in terms of the nature of the crack path. A central feature of this

bone fracture behaviour is that weak interphase regions between constituents at

different levels provide multiple preferred paths for cracking, as opposed to a

single crack pathway for many homogeneous materials. Interphase regions exist

between the constituents in bone that have specific orientations, and provide the

Page 79: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

78

basis for the marked anisotropy of the fracture properties of bone (bone is easier

to split and to break)[6, 15, 16, 118] and for the fact that the toughness is actually

lower in shear than in tension [14-16].

Figure 2.15 Three-point bending for toughness measurements. (A) Schematic

showing the three point bending test for measuring toughness. (B) A typical

load-displacement curve obtained from testing on mouse tibia bone in the

Load

Notch

Support

Lo

ad (

N)

Displacement (MM)

Slope of line = Young’s modulus

Work done to fracture

(A)

(B)

Page 80: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

79

longitudinal direction [121].

The toughening of bone materials can be explained by both deformation or

shielding mechanisms as reviewed by Launey et al. [15]. The strength and

plasticity of bone are highly affected by the mechanical properties of its

constituents which are strongly dependent on the scale of examination. Many

studies have indicated the toughening mechanisms of bone are active

predominantly at submicrometer levels [2, 9, 16, 122-126]. However, a smaller

number of studies have suggested crack propagation at larger scales absorb the

most energy [13]. While it is difficult to accept that small scale constituents such

as MCFs do not contribute to crack propagation at any length scale, a model

describing the relative contributions of the different length scales to toughness

has yet to be provided.

Previous work has indicated that osteons are the main structural feature that

appears to control toughening at large length scales of the order of several

hundred micrometres [104]. The osteons in bone fracture provide a source of

toughening that arises during crack growth rather than during crack initiation.

Hence, osteons are not inherently tough but control the propagation of cracks in

bone, known as shield toughening [115, 127]. As shown in Figure 2.16, the

existing osteons in bone provide a longer crack propagation pathway, which

Page 81: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

80

requires an increase in the overall work done to fracture of whole bone.

Figure 2.16 Anisotropy in the fracture behaviour of bone. (A) and (B) show the

crack propagate along osteon interfaces and through osteons to create longer

crack paths in three point bending tests on bone at the longitudinal direction

(defined as the osteon parallel to specimen long axis). (A) Scanning electron

micrograph showing crack tip and (B) schematically shows the crack growth. (C)

and (D) demonstrate the crack developing in bone in the three point bending test

on bone in the transversal direction (osteon vertical to specimen long axis). In

the schematic diagrams (B) and (D), the osteons are represented in brown while

the interface regions are in yellow.

(A) (C)

(B) (D)

Longitudinal Transversal

Preferable

crack path

Driving force

Driving force

Preferable

crack path

Crack tip

Uncracked-ligament

bridging

Uncracked-ligament

bridging

Page 82: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

81

In compact bone, the crack path of least resistance is invariably along the brittle

hypermineralized interfaces called cement lines [14, 126, 128] between osteons

and bone matrix. The cement lines are the preferential sites for major

microcracks to form. When a crack propagates, the microcracks form near the

crack tip area as shown in Figure 2.17 (A) where highest stresses are distributed.

These microcracks have a typical spacing in the tens to hundreds of micrometres

and are aligned primarily along the long axis of the bone, an orientation that

directly results in the strong toughness anisotropy in bone. As these microcracks

are mostly limited in the cement lines spaces, they are also called ‘monostrained

microcracking’ [12, 115, 127]. Some recent simulations [9, 126] have clearly

shown that, although microcracking is important intrinsically for toughness, its

direct contribution to toughness is minimal and has been estimated to be less

than 10 % [9, 126]. However, the importance of microcracking in toughening

mechanisms is that it is essential for developing other significant toughening

mechanisms at micron length scales and above [6, 129], including crack bridging

and crack deflection [128, 130-133], which are the most potent toughening

mechanisms in bone at large length scales. Microcracking may also important in

providing signalling for bone remodelling which occurs in bone multicellular

units (BMUs) including both osteoclasts and osteoblasts [134].

Page 83: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

82

Fracture mechanics descriptions of crack growth highlight the competition

between the direction of maximum mechanical driving force and the path of

weakest structural resistance [14, 135]. High toughness is usually developed

when these two paths exist orthogonally. For example, in three point bending

testing on longitudinal oriented bone samples, the crack driving force is vertical

to the preferable crack path (starting from microcracks in cement lines) which is

along the osteons generating high toughness as seen in Figure 2.16 (B). In

contrast, in the transversal direction, preferred mechanical and microstructural

crack paths are nominally in the same direction as shown in Figure 2.16 (D). The

orientation of the osteons can lead to significant (macroscopic) deflection of

cracks that are attempting to propagate in the longitudinal oriented sample,

which makes the longitudinal orientation much tougher than the transversal

direction. Recent fracture mechanics measurements show that after only 500 μm

of cracking, the fracture toughness, specifically the driving force for crack

propagation, is more than five times higher in the longitudinal direction than in

the transversal direction [128]. This toughening mechanism is named crack

deflection [15] as shown in Figure 2.17 (B).

Page 84: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

83

Figure 2.17 (A) to (D) Different toughening mechanisms in compact bone at the

architectural level. (E) Example of an R-curve [136] (Fracture toughness vs. crack

length) for an equine femur bone specimen.

(A) (B)

(C) (D)

Constrained

microcracking

Microcracks

Collagen fibril

bridging

MCFs

Uncracked ligament

bridging

Bridge

Crack deflection

Osteons F

ract

ure

To

ugh

nes

s (M

Pa

m1

/2)

Crack Length (mm)

(E)

Page 85: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

84

The crack path and toughening mechanisms in the transversal oriented sample

are quite different to the longitudinal direction. Preferable crack propagation

paths are now parallel to the driving force which leads to the microcracks formed

in cement lines in front of the crack tip and have the same direction as the crack

path. The aggregation of these microcracks leads to the formation of the

‘secondary’ cracks or ligaments in the uncracked region ahead of crack tip as

shown in Figure 2.17 (C). These secondary cracks act to bridge the crack and

carry load which would otherwise be used to further crack propagation [9, 104,

126, 128, 137]. This mechanism called uncracked ligament bridging results in

toughening but is a less significant mechanism than crack deflection [15, 128].

Another crack bridging mechanism can be found in tests on both longitudinal

and transversal bone samples at smaller scale. Collagen fibrils bridging in many

of the microcracks formed during in fracture affected area in bone [Figure 2.17

(D)] could help to stop crack propagation [133]. Researches showed that fibril

bridging does not contribute significantly to the overall toughness (~1 MPa.m1/2)

for a single propagating crack [126, 128]. However, it is still important to carry

out further study for its cumulative, integrated effect throughout the numerous

microcracks that are generated in bone, which could lead to a novel toughening

mechanism. The effect of collagen fibrils on bone toughness has been ignored in

these previous studies due to the difficulty of examining samples at low

Page 86: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

85

dimension but the influence of collagen fibril structural features on bone

toughness has been highlighted as important in some simulation works [138].

Specifically, numerous fibrils pulled out from the bulk bone material around

microcracks will create a relatively large new surface. The energy absorbed

during the creation of this new surface area could contribute dramatically to the

toughening mechanisms of bone. In addition, fracture between extrafibrillar

mineral crystals and MCFs in the plastic deformation of bone initiates the

formation of microcracks [139], which is the basic phenomenon involved in

toughening mechanisms including crack deflection and uncracked-ligament

bridging.

The multiple length-scale toughening acting in cortical bone leads to a

characteristic resistance-curve (R-curve) behaviour of bone material where

fracture toughness increases with crack extension [13, 136, 140], which is shown

in Figure 2.17 (D). It is believed that R-curve behaviour should be sensitive to

regional microstructural differences in bone [136]. With crack propagation, there

is an increasing number of crack deflection and microcracks appearing in a

neighbouring space away from the crack tip to enhance the fracture resistance.

Further, crack propagation resistance in longitudinal samples is more sensitive to

the regional structural differences comparing to transversal samples [127].

Page 87: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

86

b. Cancellous bone

Mechanical properties of cancellous bone

The lighter, porous tissue in bones, called cancellous or spongy bone, also plays

an important role in whole bone’s mechanical functions. The mechanical

properties of cancellous bone are more complex than cortical bone as its

mechanical properties are defined by both the mechanical properties of the

cancellous bone material (which is the solid bone making up the trabeculae) and

its porous structure. To isolate the material properties of cancellous bone from

its structural effect, single trabeculae have been machined away from the

cancellous bone matrix and mechanically tested using different testing methods.

Table 2.3 summarizes a range of methods used for measuring mechanical

properties of cancellous bone material including buckling testing, direct

mechanical testing, nanoindentation, ultrasonic testing and computational

simulation. In bulking tests, a single trabeculum was applied with compression

from both ends and the load required for buckling the trabeculae was recorded to

evaluate the Young’s modulus by Euler is buckling formula [141, 142].

Page 88: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

87

Table 2.3. Recorded Young’s moduli of cancellous bone material taken from

Currey [12].

Authors Year Bone type Method

Young’s

Modulus

(GPa)

Wet

/Dry

Runkle and Pugh 1975 Human femur Buckling 8.7 Dry

Townsend et al. 1975 Human femur Buckling 14.1 Dry

11.4 wet

Ashman and Rho 1988 Bovine femur Ultrasonic 10.9 wet

Human femur 12.7 wet

Choi et al. 1990 Human tibia 3-point bending 4.6 Wet

Cortical 3-point bending 4.4 wet

Kuhn et al. 1989 Human ilium 3-point bending 3.7 Wet

Cortical 4.8 Wet

Mente and Lewis 1989 Human femur Cantilever + FEA 6.2 Wet

Ryan and Williams 1989 Bovine femur Tension 0.8 Drying

Jensen et al. 1990 Human vertebra Structural analysis 3.8

Rho et al. 1993 Human tibia Tension 10.4 Dry

Cortical 18.6 Dry

Rho et al. 1993 Human tibia Ultrasound 14.8 Wet

Cortical 20.7 Wet

van Rietbergen et al. 1995 Human tibia 3-D FEA 6 Wet

Turner et al. 1999 Human femur Ultrasound 17.5 Wet

Cortical 17.7 Wet

Human femur Nanoindentation 18.1 Wet

Cortical 20.0 Wet

Zysset et al. 1999 Human femur Nanoindentation 11.4 Wet

Cortical 16.7 Wet

van Lenthe et al. 2001 Bovine femur FEA & ultrasound 4.5 Wet

Young’s modulus of trabeculae listed in Table 2.3 is noted as being slightly lower

than the values of compact bone in Table 2.1. In a following compression study

between the individual trabeculae and cortical bone specimens with the same

geometry and dimensions, it was found that the cancellous bone material has a

Page 89: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

88

lower Young’s modulus comparing to compact bone [25, 26]. The

nanoindentation tests on both individual trabeculae and cortical bone also

suggest a similar variation [25, 26] which was further confirmed by ultrasound

investigations on cancellous bone and was explained by the less organized

structure in cancellous bone material [26].

Mechanical function of cancellous bone and mechanically mediated bone

adaptation

Large volumes of whole bone are occupied by the cancellous structure to

minimize mass [1, 3]. Two features can be easily observed from bone’s anatomy

structure shown in Figure 2.8 (A). Firstly, cortical and cancellous bone can be

clearly distinguished as two distinct structures. Secondly, trabeculae are

arranged following a certain pattern in an angle associated with the direction of

applied loads. This organization of trabeculae indicates bone possesses a degree

of structural adaptation. Early theories first discussed in 1867 [143] presented

drawings of trabecular arrangements in human’s proximal femur that were

identical to the principal stress lines of a crane as shown in Figure 2.18. Wolff

notably developed theory based on similar observations and defined the classical

‘Wolff ’s law’ [143]:

Page 90: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

89

"Every change in the form and the function of a bone or of their function alone is

followed by certain definite changes in their internal architecture and equally

definite secondary alterations in their external confirmation, in accordance with

mathematical laws"

Figure 2.18 (A) Culmann’s schematic diagram showing principal stress line in an

engineering crane. The crane was under a distributed vertical load from the top.

(B) Wolff’s depiction of trabecular arrangement in proximal femur bone. Cited

from Wolff [143].

Further bone adaptation theories provided a summary of the common

mechanisms used to optimize bone when under external loading including:

(A) (B)

Page 91: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

90

Trabeculae are oriented along axes of principal stress.

The highest bone density can be found in areas of highest shear stress.

Bone tissue adapts to external mechanical stimulation in an optimal

manner.

Bone cells respond to local force and adapt bone tissue correspondingly.

The theories of bone adaptation have been developed dramatically afterwards

especially with the help of numerical simulations [144-148]. At present, the

current state of bone adaption theories can be summarized as:

The structural adaptation of bone is stimulated by mechanical strain

[147].

The development stage of bone is critical in defining how bone tissue

responds to external load [147].

The tolerance of mature bone to mechanical strain is reduced when

compared to more immature bone and mature bone operates within a

limited range of strain [147, 149].

Bone structure developed due to mechanical adaptation show attributes

of efficient or optimized structure [147, 148].

Bone cells (mainly osteoblast and osteoclast) can directly respond to

Page 92: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

91

applied mechanical strain [146, 148].

From discussion above, we can see how cancellous bone adapts to load and forms

its unique architecture with trabeculae oriented along the principle stresses

transferring applied loads effectively. An additional important mechanical

function of cancellous bone is the ability to absorb energy from impact [1, 3].

Figure 2.19 shows the stress-strain curve of cancellous bone and indicates a

characteristic shape optimized for energy absorption [150].

Figure 2.19 Stress-strain curve recorded in a cancellous bone compression test

from Currey [12] based on data derived from Fyhire and Schaffler [150].

1st trabeculum yield

Trabeculae yield and buckle

All trabeculae in one cross

section buckle

Plateau region

Strain

Str

ess

(M

Pa

)

Page 93: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

92

In the stress-strain curve of cancellous bone under compression testing, a clear

linear response can be observed at the start of the compression test which is due

to the elastic loading of bone material. Applied compressive strains further

causes a drop in the stress until a plateau stress region is achieved as shown in

Figure 2.19. The initial stress drop with strain is due to buckling of trabeculae.

The ‘stepping’ indicates increasing numbers of trabeculae involved in the

buckling. Finally, all trabeculae fail in bucking and the bone becomes compact to

prevent from further collapse. These trabeculae buckling processes occur over a

relatively large strain range to ensure that the work done is maximized during

compressive testing of cancellous bone [12, 150].

2.4.3 Level 5-3: From Tissue to Fibrils

Several levels exist in bone’s structural hierarchy that are below the architecture

level (macro level) and above fibrillar level (nano level). These include Haversian

systems, lamellar structure and fibril arrays as discussed in Section 2.3. Although

unique features existing at these levels, such as cement line or lacunae sites, are

believed to have different mechanical properties comparing to bone tissue, most

of the structural components at these levels have similar mechanical properties

and compositional proportions comparable bone tissue at the architectural level

[18, 37, 38].

Page 94: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

93

2.4.4 Level 2: Mineralized Collagen Fibrils

The origin of the mechanical properties of bone is contentious, with structural

organization over different length scales and constituent properties all expected

to contribute [1, 2, 7, 49]. To isolate these constituents and their potentially

synergistic effects is crucial in understanding the mechanism for bone

deformation and fracture. MCFs are considered as the basic building blocks of

bone as they are the smallest structural units containing all major materials in

bone and all MCFs from different sources are expected to be the same in

composition and material properties [2, 7, 49]. The collagen fibrils act as an

organic framework in bone and determine, at least to some extent, the

mechanical behaviour of bone. Bone is structurally a nanofibres composite of

collagen fibrils acting as a template for the mineralization of carbonated apatite

crystals. These mineral crystals exist both within the collagen fibril at the gap

region and also between fibrils.

As the basic building blocks of bone, MCFs take more than 80% of the total

volume of bone making it the most important constituent in defining mechanical

properties of bone. Some studies suggest the spatial organization of collagen

fibrils at higher levels determines the materials properties of bone at macro level.

However, the mechanical properties of these structural constituents at higher

levels are essentially defined by the collagen fibrils [2, 7, 49]. In addition, in bone

Page 95: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

94

fracture, the MCFs bridging in cracks provides resistance to crack propagation

[133] which contributes the high toughness of bone. Further study showed the

fracture of extrafibrillar mineral crystals surrounding fibrils or delamination at

the crystal/MCFs interfaces has been suggested as the cause of microcracks [139]

which is the essential phenomenon of most of the potent toughening

mechanisms.

Measurements and modelling of the mechanical properties of collagen and bone

have been performed at molecular, micron, and macro scales. However, the scale

around hundreds of nanometres, corresponding to the diameter of one fibril, is

relatively unexplored. Despite the fact that we have a relatively complete

understanding of structural information of MCFs in bone, their mechanical

properties and the influence on overall bone mechanics, such as toughness in

bone, has yet to be determined.

Several methods have been used to evaluate the mechanical behaviour of

mineralized collagen fibrils in bone. With the development of material testing

instrumentation, there are currently more techniques can be used to provide

capabilities on low scale mechanical testing such as MEMS and AFM. However,

the direct testing on isolated single mineralized collagen fibrils from bone tissue

has yet to be obtainable using current methodology [20, 40, 105, 151]. In this

Page 96: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

95

thesis, individual mineralized collagen fibrils and its system from bone are tested

using a combination of atomic force microscopy (AFM) to manipulate and

mechanically test individual mineralized collagen fibrils while scanning electron

microscopy (SEM) provides high resolution in situ imaging.

Previous work has examined the mechanical properties of unmineralized CFs

from a number of sources directly using AFM [20, 40]. AFM is often preferred to

other techniques as both high-resolution imaging and mechanical deformation

can be provided using a single AFM probe. First attempts on direct mechanical

testing of in vitro-assembled human type I collagen fibrils were carried out by

Graham and co-workers [39] using an AFM probe to pull an individual collagen

fibril from a connecting substrate. However, the relatively low collagen fibril

Young’s modulus (32 MPa) suggested sliding of the collagen fibril from the AFM

probe during testing, which was overcome using glue at the fibril-AFM probe

junction to give modulus values approaching 0.8 GPa [20].

Further research was carried out on the mechanical properties of unmineralized

individual CFs from sea cucumber using a microelectromechanical (MEMS)

device [40]. The selection from this source was beneficial as the fibril diameters

are typically relatively large, allowing manipulation of fibrils under optical

microscopy for nanomechanical tests.

Page 97: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

96

(A)

(B)

Figure. 2.20 Schematic of direct mechanical testing on mineralized collagen

fibrils. (A) Force spectroscopy experiment on collagen fibrils from Graham’s

work. The fibrils are adsorbed between a clean glass surface and an AFM

cantilever. The movement of the cantilever caused the resultant strain of

collagen fibril and recorded by laser optical sensor [39]. (B) A similar test

improved by van der Rijt et al using a drop of glue on the glass surface [20]. Both

of the tests recorded relatively low modulus and did not measure the fracture

properties of collagen fibrils.

Photodiodes

Laser

AFM probe

Collagen

fibril

AFM probe

Collagen

fibril Glue

Page 98: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

97

Both AFM and MEMS devices are limited as individual fibrils are required to be

first removed from the parent material sample, typically using chemical

treatments that will dissolve hydroxyapatite in mineralized fibrillar assemblies.

Other methods have overcome this limitation by using combinations of

synchrotron x-ray diffraction together with tensile testing [152-154] and

nanoindentation [31] to deform bulk bone materials and derive collagen fibril

and mineral mechanics. While powerful, these techniques average the

mechanical properties of the constituents over the volume of interest, typically

many tens of cubic microns in the case of x-ray diffraction. Therefore, the

mechanical properties of mineralized collagen fibril constituents as well as their

influence on bone deformation and fracture are still to be ascertained.

2.4.5 Sub-structural Components of Bone

a, Calcium Apatite Mineral

The natural calcium appetite crystals from bone have yet to be isolated and

tested due to the small size especially the extremely small thickness [68, 69].

Studies on artificial synthetic hydroxyapatite have shown that the Young’s

modulus of synthetic powdered carbonated apatite was 109 GPa, whereas a

larger value of 114 GPa can be observed for a large single crystal of

hydroxyapatite [3]. The reason why dahllite in bone has a unique plate shape,

Page 99: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

98

which differs from symmetrical hexagonal crystal structure of bulk dahllite, is

still unknown.

b, Mechanical Properties of Collagen

The mechanical testing on bulk collagen materials such as tendon and cartilage

that predominantly contain type I collagen has been previously carried out. The

recorded materials properties on collagen itself (usually assembled in vitro or

extracted from tendon and cartilage) generally show relatively large variations.

The Young’s modulus values recorded on hydrated type I collagen from tendon

ranged between 400 MPa and 1000 MPa, while failure strength having values

from 50 to 100 MPa and the ultimate strain ranged from 4 to 20% [42, 46,

155-158]. The water content and arrangement of fibrils were believed to be

crucial in determining the overall mechanical behaviour of the collagen material.

At the molecular level, plastic deformation of collagen involves the stretching of

molecules and unwinding of tropocollagen molecules due to the entropic first

and then the breaking of H bonds between molecules. The intermolecular sliding

provides large plastic strain (up to 50% in physiological conditions) within

collagen without catastrophic failure [105, 159-161].

Page 100: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

99

c, Interaction Between Mineral Crystals and Collagen Molecules

Within MCFs

Because the mineral phase has an elastic modulus that is more than an order of

magnitude higher than that of collagen, the presence of the hydroxyapatite phase

is critical to the stiffness of bone. Previous researches show a continuous

increase in Young’s modulus with mineralization of collagen fibrils, ranging up to

a factor of three for high mineral content [12]. Computational molecular

modelling and experimental X-ray analysis suggest that under tension, slippage

between mineral crystals and tropocollagen molecules is initiated and following

by a continuous sliding between tropocollagen molecules as well as between

mineral particles and tropocollagen molecules. This sliding at the molecular level

enables a large regime of dissipative deformation as soon as yield point is

reached which effectively increases the toughness [138] (Figure 2.21). Stress in

fibrils can be maintained after slip started because of additional resistance to slip

at the interface between collagen and mineral particles, which results in a

dramatic increase in energy dissipation.

Page 101: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

100

Figure 2.21 Computational simulation results of the stress-strain response of a

mineralized fibril versus an unmineralized collagen fibril, demonstrating the

significant effect of the presence of mineral crystals in collagen fibril on its

mechanical behaviour [138].

In addition to providing a plastic deformation mechanism, this molecular

behaviour of collagen and mineral phases under deformation within MCFs also

represents a fibrillar toughening mechanism, which increases the energy

dissipation compared to unmineralized collagen fibrils [105, 138, 151]. The

existence of the nanoscale mineral crystals in MCFs increases the fibril Young’s

modulus and fracture toughness. From the same computational simulation, the

Young’s modulus, yield strain, and fracture strength in tensile tests of MCFs were

found to be 6.2 GPa, 6.7%, and 0.6 GPa, compared to corresponding mechanical

properties of 4.6 GPa, 5%, and 0.3 GPa from testing collagen fibrils without a

Page 102: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

101

mineral phase [105, 138, 151].

d, Non-Collagenous Proteins (NCPs)

The final constituent of bone, which is often ignored in bone structure, are the

two hundred or more so-called non-collagenous proteins (NCPs) [71] generally

comprising less than 10% of the total protein content. Currently, there is little

data on the mechanical behaviour or composition of NCPs but MCFs are known

to be “glued” together by a thin layer (1–2 nm thick) of NCPs [162, 163]. When

bone is externally loaded in tension, the applied load is resolved into tensile

deformation of the MCFs and shear deformation in NCP regions [125]. Single

molecule spectroscopy of fractured bone surfaces has confirmed that the NCP

layer between fibrils is relatively weak but ductile and deforms by the successive

breaking of a series of sacrificial bonds [164, 165]. The separation of individual

MCFs during plastic deformation of bone is resisted by this NCPs layer via

sacrificial bonds. The force required to break these sacrificial bonds has been

estimated to be about 10%-50% of the force required to break the backbone of

the protein chains of NCPs [162]. The NCPs may be also partially mineralized

[166], which will increase its shear stiffness and reduce its deformability.

Previous work has indicated that the interface between the NCPs and MCFs is

formed of sacrificial bonds that break during plastic deformation of bone but

Page 103: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 2. Literature Review

102

reform, which leads to a toughening mechanism whereby the matrix (NPCs) /

fibre (MCFs) interface is disrupted beyond the yield point, and the matrix moves

past the fibres forming to reform the matrix/fibre bonds continuously [162, 167,

168].

2.5 Conclusions

To conclude, bone is a hierarchical composite material composed primarily of

assemblies of collagenous protein molecules, water, and mineral nanoparticles

made of carbonated hydroxyapatite. The resultant bone structure is extremely

tough, lightweight and adaptive to the physiological loading conditions

experienced by the bone material. Methods of mechanically testing bone material

at each hierarchical structural level and the correlated mechanical behaviour as

well as bone toughening mechanisms were reviewed but a general lack of

information at fibrillar length scales and their resultant contribution to overall

bone mechanics was highlighted. Mineralized collagen fibrils are particularly

crucial in defining overall mechanical behaviour of bone as they are the basic

building blocks of bone. Specifically, the fibrillar network in bone plays an

important role in the formation of microcracks and other toughening

mechanisms in plastic deformation of bone.

Page 104: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

103

Chapter 3

Methodology

3.1 Materials Preparation

Cortical bone samples were extracted from red deer and mouse for mechanical

testing of their fibrillar system at the nanoscale. Antler bone was chosen due to

its extraordinary toughness and absence of extrafibrillar mineralization. The

mouse limb bone was chosen because of its rapid growth of a skeletal system and

associated changes in the mineralization. Due to the considerable size variations

found in whole antler bone and mouse limb bone, a range of sample preparation

strategies were required. In this section, the preparation methods for antler and

mouse bone samples used in this study are reported in detail.

Page 105: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

104

3.1.1 Antler Bone Sample Preparation

A diversity of collagen fibril assemblies exist in bone materials with the antlers of

deer notable as one of the toughest natural materials [49, 103, 129, 169].

Importantly, many of these apatite minerals are intrafibrillar in antler and

provide a model composite fibre system of collagen reinforced with mineral, as

opposed to many other bones where the mineral is also found in extrafibrillar

spaces [103, 129, 154, 169, 170]. The prevalence of mineral within the collagen

fibrillar framework in antler makes this material an ideal source of model

mineralized collagen fibrils. Specifically, most of the mineral in antler is found

within the fibril, resulting in a simple composite fibre of mineral reinforcing

collagen molecules. Antler is notable as a bone material that exists outside the

body and the cortical bone tissue in antler is known to have low water content

varying from 13.2 1.2 % to 22.2 4.8 % depending on different times of the

year [171]. This relatively dry state of the antler is importantly potentially

compatible with the environment of an electron microscope chamber.

Samples were extracted from the main beam from an antler (after removal of

velvet) of a mature red deer (Cervus elaphus,). Because recent studies have

shown that the selection of tissue region within the antler beam affects the

composition and mechanical properties [154, 171], the samples were selected

from the same compact cortical shell near the antler-pedicle junction. Small

Page 106: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

105

beams of antler with dimensions of 3200.2 mm were cut from the bulk

material by using a water-cooling rotating diamond saw (Struers Accutom-5,

Figure 3.1).

Figure 3.1 Struers Accutom-5 water-cooled rotating diamond saw.

The long direction of antler samples was oriented parallel to the antler main

beam direction, which is also the principal osteonal axis as shown in Figure 3.2.

The beams were transferred into Hank’s buffered solution and left overnight.

This procedure allows full sample rehydration and mitigates mineral loss that

may occur in distilled water or physiological saline. Samples were fractured

perpendicular to the long axis to expose a number of individual fibrils at the

fracture surface. Water on the surface of sample was removed by filter paper to

avoid interference with SEM imaging.

Page 107: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

106

Figure 3.2 (A) Optical image of an antler cross section, highlighting the compact

cortical shell and the trabecular core. (B) Schematic representation of the sample:

S shows the sample location and the orientation within the cortical bone C of an

antler section; the trabecular region is indicated by T; (C) Optical microscopy

image of a polished and HCl etched cortical bone cross section surface, showing

(A) (B)

(C)

(D) (E)

Page 108: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

107

the typical tissue composition used for the experiments. (D) and (E) scanning

electron micrographs on fractured antler samples at different magnifications

showing exposed MCFs.

3.1.2 Mouse Bone Sample Preparation

Due to its relatively small size, the mouse limb bones were used directly as

specimens without further cutting as for the antler. Fresh femur and radius bones

from wild type 4 week and 10 week old mice were extracted as shown in Figure

3.3 (A). Limb bones were kept wet in Hank’s buffered solution during

preparation and stored at -200 C in wrapped gauze soaked in Hank’s buffered

solution for up to 1 week until mechanical testing. Mouse bone samples were

allowed to defrost at room temperature for 24 hours prior to mechanical testing.

Soft tissue was removed from the bone surface using a blade following the

defrosting stage and the fresh bone sample was kept in 70 % ethanol for 4 hours

for fixation. Limb bone samples were then transferred to Hank’s buffered

solution and left overnight to allow full sample rehydration and mitigate mineral

loss that may occur in distilled water or physiological saline. All the processes

above have been approved to have minimum influence on mechanical properties

of bone [1, 7, 12, 16, 48, 49, 56]. Femur samples were used for direct tensile

testing as it is easier to perform due to its larger size while radius samples were

fractured perpendicular to the osteon long axis to expose the individual MCFs at

Page 109: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

108

the fracture surface. Water on the surface of sample was removed by filter paper

to avoid interference with SEM imaging. The fractured mouse radius bone was

then transferred to the SEM chamber for MCF tensile testing as shown in the SEM

image in Figure 3.3.

Figure 3.3 (A) Schematic skeletal system of a mouse indicating the location of

femur and radius. (B) Fresh mouse radius bone from a wild type 10 week old

male mouse with muscle and connective tissue covering the bone surface. (C)

(A)

(B)

1 cm

(C)

MCFs

Bone

Page 110: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

109

Scanning electron micrograph showing individual mineralized collagen fibrils at

the bone fracture surface.

3.2 in situ Nanomechanical Testing

As proposed in Chapter 1, a mechanical testing method, having ability in

measuring nanofibres in their native state, is required. The nanomechanical

testing setup used for this thesis is developed by integrating an atomic force

microscope (AFM) for force measurements with a scanning electron microscope

(SEM) to provide imaging capabilities. In this chapter, relatively simple samples

of electrospun polyvinyl alcohol (PVA) and nylon-6 nanofibres were manipulated

and tensile tested using the AFM-SEM setup in order to validate the method to

evaluate the feasibility of this method. The complete stress-strain behaviour and

failure of individual electrospun nanofibres was recorded and a diversity of

mechanical properties observed, highlighting how this technique is able to

elucidate mechanical behaviour due to structural composition at nanometre

length scales.

3.2.1 Introduction

Advances in nanotechnology and the growth in synthetic nanomaterial

manufacture have required the development of specialized material mechanical

Page 111: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

110

characterization methods at low dimensions. Nanofibres are one particular form

of nanomaterial that show promise as a structural material in applications

predominantly where the nanofibre acts as reinforcement in a composite

[172-174]. Synthetic nanofibres typically exhibit superior mechanical properties

along the fibre’s principal axis, especially in polymeric materials incorporating

uniaxial orientation of molecular chains [173, 175], and low defect density [176,

177] comparing to bulk material equivalents. Investigating the mechanical

properties of nanofibres is therefore of critical importance in understanding the

inherent nanofibre behaviour but is often challenging due to the relatively small

length scales considered. Bone can also be considered as a nanofibrous

composite consisting of MCFs with additional complex structural hierarchy that

provides optimized mechanical properties. The effectiveness of biological

composite materials in a number of mechanical applications, most notably for

toughness, has motivated researchers to both understand structure-property

relationships in biology and develop bio-mimetic materials incorporating such

biological design features [178-181]. Current challenges exist in relating the

nanoscale, and typically fibrous, components found in structural biological

materials to their overall mechanical properties. Interestingly, nanofibrous

components are common to most biological materials, from MCFs in this thesis

that provide the organic framework in bone [17] to cellulose fibrils in wood

[182]. Measuring the mechanical properties in both synthetic and biological

Page 112: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

111

nanofibres is therefore relevant in determining the overall mechanical properties

of structural materials incorporating nanofibre components. Critically, difficulties

exist in testing constituents that have dimensions approaching the nanoscale,

especially when testing nanofibres beyond elastic limits to failure.

Mechanical testing of an individual nanofibrous sample is expected to be the

most direct method for examining mechanical behaviour but three practical

challenges exist in this approach: i) observation of the nanofibre, ii)

nanomanipulation of the nanofibre so that a mechanical test can be achieved and

iii) accurate recording of the applied force and deformation during mechanical

testing [30]. Conventional universal materials testing machines are typically able

to test a number of materials that are observable optically and can be

manipulated by hand, with the range of forces recorded during mechanical

testing reflecting the sample dimensions in the millimetre range and larger.

However, the relatively small size of nanofibres cannot be observed optically and

requires greater manipulation precision and force resolution than provided by

conventional methodologies. Simple approaches based on the earlier

instrumented indentation [183] or more recent nanoindentation techniques [184]

have been previously used to predict the tensile strength of materials, including

one-dimensional nanomaterials, based on applied load and contact radius of the

indenter tip with the sample [185]. However, the testing configuration is typically

Page 113: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

112

difficult to control, with considerable assumptions in the material behaviour

used to calculate mechanical quantities such as elastic modulus and yield

strength [186-188].

During the last decade, atomic force microscopy has been frequently employed as

a high-resolution force measurement tool coupled with its ability to image

surfaces with nanometre resolution. AFM force spectroscopy has sufficient force

resolution for investigating mechanical properties of surfaces [189-193],

molecular and nanofibrous samples [164, 177, 194], bimolecular interactions [22,

195, 196] as well as single polymer and protein mechanics [24, 197-199].

However, in small scale fibre mechanical testing studies, manipulation of the

fibrous samples through to mechanical testing requires direct observation of the

sample to ensure correct alignment of the fibre along the loading direction and

verification of sample failure.

The current state-of-the-art in direct tensile testing of nanofibres typically uses

dedicated nanomanipulators to move an individual nanofibre to the end of an

AFM probe while observing this manipulation using high resolution scanning

electron microscopy, and was pioneered from investigations into carbon

nanotube mechanics [164, 172, 200]. Gripping of the nanofibre to the AFM probe

is achieved either by electron beam-induced deposition (EBID) [200] at the

Page 114: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

113

nanofibre-AFM probe junction or manipulating the AFM probe into epoxy glue

[164] prior to the nanofibre attachment to the probe apex. The free end of the

nanofibre attached to the AFM probe is typically adhered to an immovable

surface, thus providing gripping at both of the nanofibre ends. Direct tensile

testing of the nanofibre is achieved by further manipulation of the AFM probe.

The force applied to the sample causes the AFM cantilever beam to deflect during

the test, with this deflection observed using high resolution microscopy. Thus,

cantilever deflection can be defined as force if the spring constant of the

cantilever is known. Similar tensile test procedures have been carried out on

electrospun nanofibres including polyethylene oxide (PEO) [201] and nylon-6,6

[202] using optical microscopy to observe the manipulation and testing due to

the fibre diameters being above the optical limits of light [30]. Finally,

microelectromechanical systems (MEMS) have been used to measure the

mechanical properties of fibrous materials but require significant manipulation

accuracy to position the sample between gripping elements using optical

microscopy or SEM to observe such manipulation [175, 203-205]. Typically, the

force sensing system of piezoelectric ceramic devices used in MEMS is less

accurate than the optical system used to measure force in the AFM.

Recent work has tensile tested electrospun polymer nanofibres with diameters

less than 100nm using AFM probes and SEM imaging [206]. Firstly, the individual

Page 115: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

114

electrospun nanofibre bridging across trenches in a substrate were hooked to an

AFM probe. Tensile testing was performed by translating the AFM probe away

from the fibre length to form a fibre ‘v’ shape during the deformation process. As

with other tensile testing techniques, the force applied to the nanofibre was

calculated from observing the AFM cantilever bending under the SEM while fibre

deformation was also recorded by the SEM. However, tensile testing of these

individual electrospun nanofibres requires both separation of nanofibres from

one another and bridging of the nanofibre across substrate trenches. This

preparation is often difficult to achieve in biological samples where nanofibre

separation only occurs in aggressive chemical environments that damage

samples [157, 207]. Tensile tests of such samples have to be carried out in their

as-received or as-fabricated state. Critically, the inherent measurement of force

applied to nanofibres from AFM cantilever observations is also less accurate than

the AFM force spectroscopy where dedicated optical sensors are used to

continually record cantilever deflections.

This section describes a novel in situ mechanical test method applied to a range

of different nanofibrous specimens by incorporating ‘true’ AFM force

spectroscopy with the imaging capabilities of a high resolution SEM. This ‘true’

AFM force spectroscopy uses an optical sensor to directly measure cantilever

deflection as is used in stand-alone AFM and does not rely on imaging to measure

Page 116: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

115

cantilever deflections with a correspondingly lower accuracy, as has been used

previously [1, 25, 34-40]. Subtle variations in nanofibre mechanical performance

are therefore expected to be accurately determined. Additional manipulation is

provided by a focused ion beam (FIB) housed within the chamber of a dual beam

configuration that allows sample sectioning following attachment to the AFM

probe. Nanofibre gripping and manipulation are achieved by using accurate AFM

piezopositioners while tensile testing experiments continually record cantilever

deflection at high force resolution using AFM force spectroscopy. This

configuration allows accurate stress-strain curves to be recorded during

nanofibre tensile testing to failure. For this study, the application of the AFM-SEM

system is explored by testing two synthetic materials of electrospun polyvinyl

alcohol (PVA) and polyamide (nylon-6) nanofibres.

3.2.2 AFM-SEM Setup and in situ Tensile Test

Electrospun polymer fibre samples of PVA and nylon-6 were provided by

Nanoforce Ltd. UK. Mechanical testing of individual nanofibres was carried out in

the chamber of the SEM incorporating an AFM system. A custom built AFM

system (attocube systems AG, Ger) with sample stage positioned 90 degrees to a

conventional AFM system was used to allow access of the electron beam of the

SEM (Quanta 3D ESEM, FEI Company, EU/USA) to the sample as shown in Figure

3.4. The AFM head consists of a cantilever on the cantilever plate attached to a

Page 117: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

116

piezoscanner, with the principal cantilever axis in the same axis as the electron

beam. Manipulation using the AFM setup is achieved using movement of the

sample stage or the AFM probe. Sample stage movement is enabled using xyz

piezopositioner situated underneath the sample stage plate as shown in Figure

3.4. The piezopositioners allow coarse movement with a maximum translation

distance of 5 mm. Fine manipulation is provided by movement of the AFM probe

attached to the cantilever using the piezoscanner connected to the AFM head.

This piezoscanner provides a maximum travel distance of 40 μm with

sub-nanometre resolution. The force detection scheme is based on an all fibre

low-coherence optical interferometer system situated behind the AFM cantilever.

The AFM in this study is compact enough to be installed within a SEM.

Samples were attached to the AFM sample stage as shown in Figure 3.4 by

carbon cement and positioned such that the direction of the nanofibres within

the sample was oriented in the horizontal plane and towards the AFM probe.

Vacuum compatible glue (Poxipol, Arg) was placed at the side of the sample on

the sample stage and the chamber taken to vacuum. Once under vacuum, the SEM

was used to image the manipulation of the sample and AFM probe. Manipulation

was carried out by moving the sample towards the AFM probe using the

piezopositioners behind the sample stage. The glue at the side of the sample

stage was first moved into contact with the AFM probe and then separated.

Page 118: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

117

Removal of the AFM probe from the glue deposited a small amount of glue at the

apex of the AFM probe. The piezopositioners were then used to move the fibrous

sample towards the AFM probe so that a nanofibre protruding from the sample

contacted the glue at the AFM probe apex, as demonstrated in Figure 3.5 (A).

Figure 3.4 (A) Schematic diagram showing the in situ configuration for tensile

testing of a nanofibrous sample using combined AFM-SEM. The insert shows

SEM FIB

Positioner

Piezoscanner AFM head

Fibre optics

Sample stage/

Fibrous sample

AFM cantilever

AFM probe Epoxy

glue Fibre optics

(A)

(B) (C)

SE

FIB

Page 119: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

118

fixing of the nanofibre with glue between the substrate and AFM cantilever. (B)

Optical photograph showing AFM fitted on SEM sample stage when door of SEM

chamber is open. Dashed rectangle indicates location of the AFM. Secondary

Electron axis (SE) and Focused Ion Beam axis are also labelled in the photo. (C)

Higher magnification optical side view image of the AFM on the SEM sample

stage.

The contact of the free length of the nanofibre to the glue at the apex of the AFM

probe was achieved consistently within a 3-5 minute timeframe from fixing the

glue within the SEM chamber. The maximum timeframe for attachment of the

nanofibre to the glue on the AFM probe was approximately 10 minutes, after

which the glue viscosity was too high to allow penetration of the nanofibre.

Manipulation and attachment of the nanofibre to the AFM probe was observed

using the electron beam of the SEM at a long sample working distance (15 mm),

low accelerating voltage (2 kV) and low electron beam current (40 pA) to ensure

that no electron beam damage occurs, as has been shown for soft polymer

systems [208]. In addition, previous work has indicated that the effect of vacuum

on hydrated samples using the SEM imaging conditions above is minimal if

specimens are exposed to vacuum for relatively short times of the order of a few

tens of minutes [17, 208, 209]. After contacting with the nanofibre, the glue was

allowed to cure for 10 minutes to ensure that one end of the nanofibre is securely

Page 120: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

119

attached to the AFM probe.

Figure 3.5 Scanning electron micrographs showing a typical tensile test of

electrospun nanofibres (PVA) using in situ AFM-SEM. (A) A single PVA nanofibre

was attached to an AFM probe using epoxy glue at the end of probe. (B)

Sectioning of the PVA nanofibre from the fibrous mat provides isolation at the

end of AFM probe. (C) Insertion of the free PVA nanofibre end into glue provides

the standard nanotensile testing configuration. (D) PVA nanofibre tensile testing

(B) (A)

(D) (C)

AFM

cantilever

probe

epoxy

glue

epoxy

glue

nanofibre

Page 121: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

120

carried out after solidification of the glue was achieved by translation of the AFM

probe away from the glue surface until the nanofibre failed at its mid-length. The

deformation of epoxy glue at the anchor point with fibre and was carefully

examined by pixel analysis and excluded from original tensile test result. The

maximum deformation of epoxy glue is relatively small (0.33 m) compared to

the length of the tested fibre recorded before and after testing (10.2 m and 12.1

m respectively)

A relatively high energy focused ion beam (FIB) integrated within the SEM

chamber, as well as electron beam with high accelerating voltage and beam

current, were used to section the polymer nanofibres to achieve a desired

isolated nanofibre length of 10-20 μm fixed to the AFM probe. FIB was used to

section the nylon-6 samples while PVA nanofibres were sectioned by focusing the

electron beam of the SEM at high magnification. The FIB accelerating voltage and

current used for sectioning was 30 kV and 1.5 nA respectively and the electron

beam was used at around 105 magnification with an accelerating voltage of 30

kV and beam current of 0.2 μA. Fresh glue (Poxipol, Arg.) was subsequently

introduced into the chamber and the free end of the nanofibre pushed into the

semi-liquid glue, as shown in Figure 3.5 (C). The glue was allowed to cure after

10 minutes to ensure that the second nanofibre end was securely fixed.

Page 122: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

121

Fixed electrospun polymer samples were tensile tested using the force

spectroscopy of the AFM. Mechanical testing of all nanofibres along their

principal fibre axis in the SEM plane of view was achieved by first observing the

nanofibre by the SEM. The AFM probe was then moved by a small amount of the

order of tens of nanometres above and below the SEM plane of view and the

forced acting on the nanofibre as recorded by the AFM examined. A small

movement above or below this plane will cause the AFM cantilever to bend

towards the glue, resulting in a recorded force, when the nanofibre in the SEM

plane of view. Misaligned nanofibres not lying in the SEM plane of view cause

asymmetric AFM cantilever bending during both movement above and below the

SEM plane of view. Hence, only manipulated nanofibres with their length in the

SEM plane of view are tensile tested.

Force spectroscopy was achieved by translation of the AFM probe away from the

sample stage. This translation caused a bending of the AFM cantilever

corresponding to the applied force acting on the attached individual nanofibre

until rupture failure occurred. All nanofibres tensile tested failed in the middle of

their free length, away from the holding glue and the substrate. The force applied

to each nanofibre sample was calculated from the deflection of cantilever

recorded using the optical fibre setup behind the cantilever. Cantilever deflection

was converted to force using the thermal noise method [23].

Page 123: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

122

The calculation of the stress in the nanofibre specimens require the accurate

determination of their diameter. The measurement of nanofibre diameters was

achieved by recording a series of nanofibre images along the nanofibre length

using the SEM prior to mechanical testing. The SEM parameters used for

manipulation were selected for nanofibre diameter observation using fast

electron beam scanning speeds ( 60 Hz) and rapid switching between relatively

low to brief high magnification capture of the nanofibre image and back to low

magnification. This methodology was applied to ensure no melting or

deformation of the nanofibre occurred due to minimal exposure of the nanofibre

to the electron beam.

After nanofibre failure, images around the nanofibre failure point were selected

from the original series of SEM images taken prior to mechanical testing and a

total of approximately 6 images per nanofibre sample underwent pixel analysis

using ImageJ (NIH, USA) to determine an average nanofibre diameter. We note

that previous studies on mechanical properties of nanofibres from various

groups have also relied on measuring nanofibre diameter using the SEM [5, 6, 11,

13, 25, 34, 40].

The force applied on the specimen is detected by an all fibre low-coherence

optical interferometer system situated behind the AFM cantilever as shown in

Page 124: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

123

Figure 3.6. A non-linear sine curve of laser signal intensity vs. retract distance as

shown in Figure 3.7 (A) is produced with cantilever bending during mechanical

testing. The original data was recorded in the experiment and converted to a

stress-strain plot as shown in Figure 3.7 (B). The testing technique is different to

many previous AFM setups [25, 26] due to the non-linear data analysis applied to

the interferometer employed in this system.

Figure 3.6 Schematic diagram of the laser interferometer used in the AFM

illustrating the interference signal measured. Approximately 4 % of the laser

light is reflected at the glass-air interface with an intensity of I1 while ~96 % of

the light is transmitted and partially reflected at the AFM cantilever with an

intensity of I2. The intensity I2 depends on the reflectivity of the AFM cantilever.

Fibre optics

Fibre sample

Page 125: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

124

3.2.3 Data Analysis Principle

The fibrous samples were tensile tested by translation of the AFM probe away

from the sample stage as described above. The tensile stress applied to the

nanofibre can be calculated using Hooke’s Law:

=

2=

𝑘𝑑

2 Equation 3.1

where F is the tensile force acting on the nanofibre, k is the spring constant of the

cantilever, d is the cantilever deflection and r is the nanofibre radius.

When the cantilever is translated away from the sample stage, the retraction

distance of the piezoscanner X behind the cantilever is recorded by AFM. This

translation causes cantilever bending and deformation of the fibrous sample. The

retraction distance X is therefore equal to the sum of the cantilever bending d

and the sample deformation ΔL. Hence, the nanofibre strain ε can be calculated

using:

=Δ𝐿

𝐿= − 𝑑

𝐿 Equation 3.2

where ΔL is the elongation of the nanofibre which excludes the glue deformation

as shown in Figure 3.5 (D), L is the original fibre length taken from SEM images

of fibrous samples at the start of tensile testing and X is the retraction distance of

the cantilever.

Page 126: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

125

Equations 3.1 and 3.2 indicate that the determination of cantilever bending is

required to measure both the stress and strain behaviour of the nanofibre during

tensile testing. A laser interferometer optical fibre situated behind the AFM

cantilever was used to accurately determine the degree of bending during the

tensile test as shown in Figure 3.6. The interference of a laser beam reflected

from the back of the cantilever to the end surface of a fibre optic interact

constructively or destructively, defined as reflected laser intensity, as the optical

fibre-cantilever distance changes. A typical sinusoidal reflected laser

intensity-distance plot for nanofibre tensile testing is shown in Figure 3.7 (A).

Figure 3.7 Force spectroscopy plot from tensile testing of an individual nanofibre.

(A) Original sinusoidal laser intensity-distance curve recorded in force

spectroscopy. The Y-axis is reflected laser intensity while the X-axis indicates the

piezoscanner retraction distance. Failure of the nanofibre is observed at a

(A) (B)

Page 127: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

126

retraction distance of 15 μm. (B) Resultant stress-strain curve converted from

the sinusoidal curve based on Equation 3.4

Figure 3.7 (A) clearly indicates a sinusoidal variation in the collected laser

intensity as the cantilever bends due to piezoscanner translation during the

tensile test. The collected laser intensity I0 is defined by the intensity of incident

lights I1 and I2 as well as the optical path length difference between the two,

defined as 2(D+d) as shown in Figure 3.6, thus:

I0 = I1 + I2 + 2√I1I2 cos [2

𝜆× 2(D + 𝑑)] Equation 3.3

where I0 is the reflected laser intensity. I1 and I2 are the reflected laser intensities,

D is the initial cantilever-fibre optic distance and D+d is the cantilever-fibre optic

distance as the tensile test proceeds. is the wavelength of the laser (1330 nm)

used in the laser interferometer setup. The cantilever deflection d is therefore

calculated from:

𝑑 = arc cosI0 − I1 − I2

2√I1I2

𝜆

4 − D Equation 3.4

Equation 3.4 can therefore be applied to the collected laser light intensity data to

determine the cantilever bending. The force sensitivity of every test was

evaluated by defining the resolution of data collection in the sinusoidal

Page 128: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

127

intensity-displacement curve. The laser intensity changing between two adjacent

data points is converted to the minimum displacement recorded in the test by

Equation 3.4. The force resolution can be then calculated from the minimum

displacement of the cantilever recorded. For tensile testing of electrospun PVA

fibres, 618 data points were collected in total with the smallest force recorded as

(5.4 0.8) 10-4 μN. Tensile testing of electrospun nylon-6 fibres recorded 410

data points and the force resolution was (1.65 0.3) 10-3 μN. 768 data points

were collected for the tensile test and the minimum force recorded was (1.74

0.5) 10-3 μN. The accurate determination of laser intensity I1 and I2 is crucial in

force sensing but is usually difficult in practice due to the following challenges

below.

a. Intensity Drop

In Figure 3.7 (A), a noticeable drop of laser intensity occurred as the cantilever

deflected in the original data plot. This intensity reduction is mainly caused by

the tilt of the cantilever along the long axis, which is schematically demonstrated

in Figure 3.8. Bending of the cantilever as force is applied to the AFM probe will

cause a deviation x of the reflected laser onto the fibre optic and a corresponding

deviation angle θ as shown in the Figure 3.8. Reflected laser light is therefore

deflected away from the core of fibre optics. The displacement of the reflected

Page 129: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

128

spot can be calculated using (in a linear approximation):

𝑦

D + 𝑑≈ sin 𝜃 Equation 3.5

is given as:

|𝜃| = 𝐿2

2𝐸𝐼m Equation 3.6

where Im is the geometrical moment of inertia; E is Young’s modulus of material

of the cantilever. The deflection of the cantilever d can be calculated using this

equation:

𝑑 = 𝐿3

3𝐸𝐼m Equation 3.7

Figure 3.8 Schematic diagrams showing the deviation of laser light reflecting off a

deflecting AFM cantilever.

Under external force applied to the AFM probe, the deflection of the AFM

Cantilever

Laser

Reflected

laser

y Fibre optics

d

Page 130: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

129

cantilever leads to a shift of reflected laser beam away from core of fibre optics.

The laser intensity recorded by optical sensor is therefore reduced with the

deflection of cantilever, which causes the drop in the reflected laser light intensity.

The deviation of laser spot on fibre optics is y shown in Figure 3.8 where:

𝑦 ≈ (D + 𝑑) sin 𝜃 Equation 3.8

Hence, if we assume a 450 µm long cantilever, with a spring constant of 0.2 N.m-1

that displaces by 30 µm, which is the typical value in experiment, and a cantilever

positioned 40 µm away from the fibre optics initially, a minimum laser spot

displacement y of 3 µm will be produced. The intensity drop in the dataset can be

fitted with a Lorentzian function such as:

I0 =(1 + cos 2ax) × b

2[1 + (x/𝑐)2]+ background Equation 3.9

where a gives the frequency of the cosine function, b is a scaling constant, c

determines the decay rate of the laser intensity and the background offsets laser

intensity from zero. An example is shown in the following Figure 3.9

Page 131: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

130

Figure 3.9 A typical Lorentzian fitting on original data obtained in a tensile

testing experiment on fibre sample. The black line is the original data plot. Red

line is the Lorentzian fitting

The schematic in Figure 3.10 shows how the AFM cantilever bends in different

ways and the effect on the resultant data recorded under different experiment

modes.

Page 132: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

131

Figure 3.10 Schematic diagram of cantilever in different working modes. (A)

Z-Spectroscopy is the usual force spectroscopy tool used for indentation and

compression test. The cantilever bends up by applying compressive load. (B)

Dither-Spectroscopy is the maintenance mode in which the cantilever is driven

by dither piezo mounted behind the cantilever, which only changes the distance

between cantilever and fibre optics. The cantilever is not loaded and therefore

does not bend. (C) In nanofibre experiments, the cantilever is under tensile load

and bends down.

(A)

(B)

(C)

Page 133: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

132

b. Laser Beam Position

The laser intensity recorded by the fibre optic during AFM cantilever bending is

determined not only by the magnitude of the cantilever bending but also by the

laser spot location on cantilever, which is shown in Figure 3.11. If the laser spot is

located at a distance L1 away from the cantilever root, the bending d1 recorded by

the laser intensity will be much smaller than the real cantilever deflection d as

shown in the Figure 3.11 below.

Figure 3.11 Schematic diagrams showing different relative positions of the fibre

optics along the AFM cantilever (positions d1 and d0). Orange rectangles

represent different locations of fibre optics after installing AFM cantilever. When

considering a constant cantilever bending, the distance between cantilever and

fibre optics at each of these positions change due to different laser positions. The

d0

d1

Page 134: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

133

insert shows the bottom view from the AFM probe with the circular in orange

representing the fibre optics.

From Equation 3.7, d1 is given as:

𝑑1𝑑0

=𝐿13

𝐿3 → 𝑑1 = (

𝐿1𝐿)3

× 𝑑0 Equation 3.10

The laser intensity recorded at the d1 position in the experiment is:

I0 = I1 + I2 + 2√I1I2 cos {4

𝜆[D + (

𝐿1𝐿)3

× 𝑑0]} Equation 3.11

Therefore, the location of the laser spot has a third power effect on the laser

intensity recorded in data plotting. The period of the cosine function above is

defined by the relative location of the laser spot on the AFM cantilever. When a

new AFM cantilever is installed, the relative laser position will typically change.

The accurate determination of the laser position on the AFM cantilever is

relatively difficult in practice as an invisible laser is used in the optical system.

3.2.4 Calibration and Data Analysis

To solve the problems of the non-linear change in the reflected laser light

Page 135: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

134

collected by the optical fibre during AFM cantilever deflection, a calibration

approach is applied. A reference dataset is recorded after nanofibre tensile

testing by first firmly attaching an AFM probe to a solid substrate of silicon fixed

to the AFM sample stage using epoxy glue within the SEM chamber, in a similar

method to the nanofibre attachment. The AFM head was translated away from

the substrate, which caused the cantilever to bend away from the solid substrate.

Cantilever deflection in this case is exactly equal to the travelling distance of the

AFM head which is defined as cantilever displacement.

To eliminate the drop of laser intensity, the data obtained from calibration and

experiment are both normalized into I0(-1,+1) and I0’(-1,+1) which are shown

in Figure 3.12 (B) and (D). We noticed that the wavelength in the experimental

data plot is ‘expanded’ compared to the calibration data plot. The expanded

wavelength can be described as the displacement value in experimental data that

is equal to cantilever deflection plus sample elongation. Meanwhile, the

cantilever displacement recorded in calibration is equal to the cantilever

deflection.

Page 136: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

135

Figure 3.12 An example of a laser intensity versus cantilever displacement plot

obtained from tensile testing on an individual electrospun PVA fibre and

subsequent calibration. (A) Part of the original laser intensity with cantilever

displacement plot recorded during tensile test. (B) The same dataset after

normalization. (C) Part of the original laser intensity with cantilever

displacement plot of calibration. (D) Calibration data plot after normalization.

The relation between laser intensity and cantilever displacement after

normalization in calibration can be defined as:

I0′ = 𝑓(𝑥′) Equation 3.12

where I0’ is the normalized laser intensity and x’ is the cantilever displacement.

Similarly, the laser intensity and cantilever displacement relation in experiment

(A)

(B)

(C)

(D) (x, I0) (x’, I0’)

0 2 4 6 8 10

-1

0

1

Inte

ns

ity

(v

)

Displacement (m)

0 2 4 6 8 103

4

5

6

7

Inte

ns

ity

(v

)

Displacement (m)

0 2 4 6 8 103

4

5

6

7

Inte

ns

ity

(v

)

Displacement (m)

0 2 4 6 8 10

-1

0

1

Inte

ns

ity

(v

)

Displacement (m)

Page 137: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

136

data after normalization is:

I0 = 𝑓(𝑥) Equation 3.13

where I0 is the normalized laser intensity and x is the cantilever displacement. To

calculate sample stress and strain, the cantilever deflection needs to be

accurately determined which can be derived from the laser intensity recorded.

Equation 3.11 showed that the laser intensity is determined by the relative

position of the laser spot on the AFM cantilever, which is difficult to measure as

an invisible laser used. However, once the AFM probe is installed, the laser spot

remains at the same position on the cantilever and the relation between laser

intensity and cantilever deflection is constant. Therefore, the cantilever

deflection at the red spot (x, I0) in the experiment is the same as the cantilever

bending at the green spot (x’, I0’) in calibration shown in Figure 3.12 (B) and (D)

as the laser intensity at these two points have the same value and phase angle.

Meanwhile, at the green spot in the calibration, the cantilever deflection d is

exactly same to the cantilever displacement x’. Therefore, the sample strain at the

red spot (x, I) in the experiment can be calculated using:

=𝑥 − 𝑑

𝐿=𝑥 − 𝑥′

𝐿 (when I0 = I0

′ ) Equation 3.14

Page 138: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

137

where L is the original length of sample. The tensile load applied on sample is

calculated as the cantilever spring constant k times cantilever deflection. So the

stress of sample at red point is calculated:

=𝑘𝑑

2=𝑘𝑥′

2 (when I0 = I0

′ ) Equation 3.15

For example, at the red data point in Figure 3.12 (A) plot: x = 2.37 μm, I0=6.38 V.

After normalization: x=2.37 μm, I0=1 V. The corresponding data point in

normalized calibration plot is the green point Figure 3.12 (D) where the x= 1.25

μm, I0’=1 V. If the fibre length L=10 μm, radius r= 40 nm, spring constant of

cantilever k=0.2 N.m-1, the stress and strain of tested fibre at that point can be

calculated as:

=𝑥 − 𝑥′

𝐿= 11.2 % Equation 3.16

=𝑘𝑥′

2= 49.8 MPa Equation 3.17

3.2.5 Accuracy and Errors

As discussed in section 3.2.3, different laser positions on the AFM cantilever

affects the outputting cantilever deflection signal and influences the accuracy of

the force measurement. Therefore, in this section, the accuracy and errors of the

Page 139: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

138

force measurements in the AFM system are discussed by considering the laser

located at the back of AFM cantilever as shown at position d0 in Figure 3.11.

a. Accuracy

As defined in Equation 3.15, the force applied to the sample is calculated from the

cantilever deflection in the test, which is equal to the cantilever displacement

recorded in calibration as discussed in Equation 3.12 and Equation 3.13.

Therefore, the force resolution of the AFM is defined as the product of the

smallest recorded displacement of the cantilever and the spring constant of the

cantilever. The Attocube AFM system used in this study has a smallest

displacement of 0.36 nm at 300 K and 0.23 nm at 4 K for z axis piezo movement

[210]. Mechanical tests were carried out at room temperature, which

approximates to 300 K and therefore provides a cantilever deflection resolution

of 0.36 nm. The AFM cantilever spring constants (K) used in nanofibre

mechanical testing in this thesis are typically ranged between 0.01 Nm-1 to 0.2

Nm-1. Therefore, the smallest force that is measured by the AFM is equal to the

product of the cantilever deflection resolution (0.36 nm) and K = 0.01 Nm-1, to

give a minimum force of 3.6 pN. The nanomechanical tests produce the recorded

forces by movement of the piezo positioner in order to provide displacement of

the AFM cantilever. In this thesis, 2048 data points were sampled during each

Page 140: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

139

tensile test during a 20 μm z axis piezo displacement. The accuracy of the

smallest detectable piezo displacement is therefore 9.8 nm, which corresponds to

a force resolution of around 100 pN when using a cantilever with a spring

constant of 0.01 Nm-1. The forces and displacements used in the AFM setup are

considerably smaller than micro electro mechanical systems (MEMS). In

particular, MEMS devices measure forces in the range of tens of nano-Newtons up

to hundreds of micro-Newtons [211-214].

Finally, the SEM used in this study has a spatial resolution of 1.0 nm in secondary

electron imaging mode at 30 kV, 2.0nm at 2 to 3 kV and 3.0 nm at 1 kV under high

vacuum operation mode [215].

b. Errors

The mechanical oscillation of the AFM cantilever is the major factor in the force

measurement error and is dominated by the vibration of the SEM sample stage.

In addition, the interference of electronic components in the AFM controller unit

can also affect the stability of the signal output. All these effects will lead to a

resultant “oscillating” signal as shown by the recorded variation in the laser

intensity, in volts, over time as recorded in Figure 3.13. The average signal is

shown in Figure 3.13 as being, on average, 5.455 V with a range from 5.45 V to

Page 141: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

140

5.46 V. As shown in Figure 3.12 (A), the largest slope in the first period of the sine

curve is recorded to be 0.52 μm/V at the point with Y value of round 5.4 V. The

intensity-displacement relation is approximately linear near the largest slope

point where the cantilever displacement can be calculated from the slope directly.

Therefore, the 0.01 V noise in laser intensity signal corresponds to a

displacement of approximately 0.01 V 0.52 μm/V = 5.2 nm. Hence, the general

error of all the interferences on the force measurement was estimated to be

about 0.1 nN from the noise level of the original signal in a 1 kHz bandwidth

using a 0.02 Nm-1 cantilever as shown in Figure 3.13. AFM cantilever with lower

spring constants of 0.01 Nm-1 have smaller measured force noise ranged

between 0.05 to 0.08 nN. As the mechanical forces recorded using tensile testing

of nanofibres range from 314 to 1413 nN, a force error of 0.1 nN represents a

maximum 0.03 % error in the recorded forces.

0 1 2 3 4

5.44

5.45

5.46

5.47

Laser

inte

nsity (

V)

Time (s)

Figure 3.13 Plot of laser intensity reflected from the AFM cantilever against time

Page 142: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

141

for the AFM integrated within SEM (1 kHz data sampling bandwidth, cantilever

spring constant is 0.02Nm-1).

The accurate determination of the stress-strain behaviour when mechanically

testing samples requires the reliable measurement of the sample’s dimension,

especially the cross section area. As shown in Equation 3.12, the radius of the

nano fibrous samples has a square dependence on their stress level. High

magnification secondary electron images of sample were taken from different

view angles in prior of testing to verify a circular cross-section. The radius of the

nanofibre was carefully measured from SEM images by using pixel analysis

software (ImageJ, NIH, USA). However, the samples tested in the experiment are

made up of non-conductive polymeric materials. Hence, charging effects due to

electrons from the SEM beam residing on the nanofibre surface influences the

SEM image quality and causes distortion of the resultant image. The evaluation of

the charging effect was carried out by measuring the radius of nylon-6 samples

with and without gold sputter coating. The nanofibre radius was calculated by

averaging 6 measured radius values at different random positions along the fibre

length. The measured nanofibre radius for 20 individual non-coated nylon-6

nanfibres showed a larger standard deviation of 14 % compared to 8 % for

nylon-6 nanofibres with a gold-coating, indicating a 6% error caused by the

charging effect alone (non-coated sample are measured in images taken under 3

Page 143: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

142

kV and gold-coated sample are imaged under 30 kV). The error in values of stress

arising from inaccuracies in measuring the nanofibre radius from SEM images

are calculated to be around 12% which is significantly higher than the measured

force noise level of the AFM cantilever. Therefore, errors in values of stress are

dominated by the fibrous sample radius measurement using SEM.

3.2.6 Results of Tensile Test on Polymeric Nanofibres

and Discussion

The reliability and accuracy of the presented novel in situ mechanical testing

setup was evaluated by performing measurements on synthetic nanomaterials.

Electrospun PVA, nylon-6 nanofibres were tensile tested to failure with the

resultant stress-strain curves for individual nanofibres shown in Figure 3.14. The

stress-strain behaviour of all of the nanofibres are different, indicating that the

AFM-SEM technique used in this work is able to elucidate mechanical behaviour

due to varying structural compositions. A summary of the nanofibre mechanical

properties is listed in Table 3.1. The deformation of the electrospun nanofibres is

initially linear but with increasing non-linearity indicating viscoelastic behaviour

for both PVA and nylon-6 at increasing tensile strain. PVA and nylon-6

stress-strain curves indicate considerable linear behaviour compared with

typical polymer materials, which exhibit considerable plastic deformation. The

origin of this stress-strain behaviour in electrospun fibres is unknown but may

Page 144: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

143

be due to the difficulty in plastic deformation at reduced length scales which has

been previously suggested [175, 185, 204, 206, 208].

Table 3.1 Mechanical properties of PVA and nylon-6 nanofibres measured by in

situ AFM-SEM and their bulk equivalents. E is the Young’s modulus measured for

each sample in their linear region up to 4% strain. UTS is the ultimate tensile

strength and UTS is the ultimate tensile strain of the sample. Comparable film and

bulk properties were measured from PVA films or taken from literature for nylon

films (*see [206]).

Materials Status Diameter (nm) E (GPa) UST (Mpa) ƐUST (%)

PVA fibre 130.657.2 0.47 0.27 62.218.5 21.34.7

film 0.20 ±0.06 31.8 80.0

nylon-6 fibre 113.016.3 1.32 ±1.52 78.16.0 31.522.2

film* N/A* 47±0.5 168 ±14.8

Page 145: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

144

Figure 3.14 Tensile stress-strain curves for different fibrous samples. (A) 5 PVA

nanofibres exhibit a linear stress-strain relationship before 10 % strain followed

by non-linear behaviour indicating ductile yield in one of the samples. (B) One of

the five nylon-6 nanofibres (N3) shows double yield point behaviour in the

stress-strain plot, suggesting crystal structure changes during deformation, while

other nanofibres exhibit clear yield behaviour.

The Young’s modulus for PVA nanofibres is 0.47 0.27 GPa. The PVA films,

prepared by solution casting from the same polymer solution used for

electrospinning and measured using conventional tensile testing (Instron, UK) in

air, gave a Young’s modulus of 0.20 ± 0.06 GPa. The PVA nanofibres therefore

have mechanical properties that are slightly higher than the bulk isotropic PVA

films, as compared in Table 3.1. The high Young’s modulus of PVA nanofibres

indicates potential structural anisotropy along the tested direction [30, 175, 201,

205, 206, 216]. We note that the PVA nanofibres can swell under the influence of

0.00 0.05 0.10 0.15 0.20 0.25 0.30

0

20

40

60

80

Str

ess

(M

Pa)

Strain

P1

P2

P3

P4

P5

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

0

20

40

60

80

100

Str

ess

(M

Pa)

Strain

N1

N2

N3

N4

N5

(A) (B)

Page 146: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

145

water, with dry fibres giving Young’s modulus values of around 3 GPa [216]. As

the nanofibres in this work are far below such a value, we conclude that water is

still bound within the PVA nanofibres despite conducting the tests in the vacuum

chamber of an SEM.

The ultimate tensile strength of nylon-6 nanofibres obtained in our experiment is

comparable to the results recorded in a previous indirect tensile test on

electrospun nylon-6 nanofibres with similar diameters using a hooking method

[206]. Nylon-6 cast films showed a tensile strength of 47 ± 0.5 MPa with a strain

to failure of 168 14.8 % [206]. The nanofibres tested in this work exhibit lower

ductility but higher strength compared with mechanical properties for film or

bulk materials reported in other studies [201, 204, 206], possibly due to the

electrospinning processing method. Interestingly, the electrospun nylon-6

nanofibre exhibits a two stage stress-strain plot in test N 3 as shown in Figure

3.14, which can be related to the double yield point behaviour of nylon-6 detailed

in previous works and observed in other materials [217-221].

The mechanical deformation of polymer fibres at reduced length scales has been

previously studied extensively. For example, an increase in the elastic modulus of

electrospun fibres with smaller fibre diameter has been reported and confirmed

to be an effect associated with an increase in fibre crystallinity [222, 223] and

Page 147: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 3. Methodology

146

supramolecular confinement [224, 225]. The molecular organization within

electrospun fibres has been shown to be size dependent with a result in reduced

plasticity and corresponding ductility [223]. Indeed, plasticity effects at reduced

length scales have been observed in thin films, with a number of reviews on the

topic [226].

3.3 Conclusions

A novel in situ tensile testing method was developed to be used in this thesis for

measuring the mechanical stress-strain behaviour of nanofibrous materials in

tension to failure. This method has a capability of testing nanofibres in their

as-received or as-fabricated state and is therefore applicable to measure the

mechanical properties of biological nanofibres like the mineralized collagen

fibrils in bone. To demonstrate the feasibility of the in situ AFM-SEM method, the

stress-strain behaviour of nanofibrous samples of electrospun PVA and nylon-6

were measured. Electrospun PVA and nylon-6 nanofibres exhibited an increased

tensile strength and elastic modulus when compared to bulk equivalents. The

mechanical properties of PVA nanofibres were shown to be similar to hydrated

samples, indicated that the vacuum of the SEM does not dehydrate such samples

within the testing timeframes used in this work and therefore indicate that this

AFM-SEM technique can fulfil the requirements of this study.

Page 148: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

147

Chapter 4

Nanomechanical Properties of Individual

Mineralized Collagen Fibrils from Bone

Tissue

4.1 Introduction

Section 2.3.2 and 2.4.4 in Chapter 2 defined how mineralized collagen fibrils

(MCFs) are distinct building blocks for bone material and perform an important

mechanical function. A diversity of MCF assemblies exist in bone materials with

the antlers of deer [49, 103, 169] notable as one of the toughest natural materials.

Importantly, many mineral crystals are found to exist within the intrafibrillar

region in antler as opposed to many other bones where the mineral is also found

in extrafibrillar spaces [103, 154, 169, 170]. The mechanical properties of these

MCFs and their influence on the overall bone mechanics, such as toughness in

antler, has yet to be determined.

Page 149: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

148

In this chapter, the in situ nanomechanical testing method of combined AFM-SEM

defined in Chapter 3 is used to manipulate and measure the mechanical

properties of individual MCFs from antler. The recorded stress-strain response of

individual MCFs under tension shows an initial linear deformation region for all

fibrils followed by inhomogeneous deformation above a critical strain. This

inhomogeneous deformation is indicative of fibrils exhibiting either yield or

strain hardening and suggests possible mineral compositional changes within

each fibril. A phenomenological model is used to describe the fibril

nanomechanical behaviour.

4.2 Materials and Methods

4.2.1 in situ Nanomechanical Testing

Antler samples for in situ nanomechanical testing were prepared as reported in

Section 3.1.1. Cortical bone extracted from antler main beam was cut into small

beams with long axis parallel to osteons and rehydrated in prior of experiment.

Nanomechanical testing of individual MCFs was performed using a combined

AFM-SEM as described in Chapter 3. A schematic of the combined AFM-SEM

setup is shown in Figure 4.1 (B) and highlights how the AFM probe is

perpendicular to the fracture plane of the antler and along the principal axis of

the exposed collagen fibrils.

Page 150: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

149

Figure 4.1 (A) Scanning electron micrograph showing a typical testing

configuration for tensile testing of MCFs. The image shows a large number of

exposed collagen fibrils observed at the fracture surface of antler bone. An

individual collagen fibril protruding from the fracture surface is attached to the

glue at the end of the AFM probe. Translation of the AFM probe away from the

fibril causes tensile deformation of the fibril until failure occurs which is shown

in the inserted image. (B) Schematic diagram showing the combined SEM-AFM

setup.

Attachment of an individual collagen fibril was carried out according to the

methodology defined in Section 3.2 in the previous chapter. However, whereas

validation of the technique was performed on synthetic nanofibres that are

manufactured in an isolated form, MCFs are bound together in bone and require

adaption of the fibre mechanical testing technique. Clamping of an individual

collagen fibril to the end of the AFM probe was achieved by first translating the

end of the AFM probe into a droplet of glue (Poxipol, Arg) contained within the

(A) (B)

Page 151: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

150

SEM chamber as described on synthetic nanofibre manipulation in Chapter 3.

Removal of the AFM probe from the glue deposited a small amount of glue at the

apex of the AFM probe. The AFM probe was subsequently moved towards the

free end of the exposed collagen fibril until contact between the fibril and the

glue at the AFM probe apex was achieved as shown in Figure 4.1 (A). This contact

was achieved consistently within a 3-5 minute timeframe from the fixing of glue

within the SEM chamber. Manipulation and attachment of the collagen fibril to

the AFM probe was observed using the electron beam of the SEM at long working

distance (15 mm) and low accelerating voltage (2 kV) to ensure that no electron

beam damage occurs, as has been shown for soft polymer systems [227].

Solidification of the glue occurred approximately 10 minutes after contacting

with the fibril free end. Thus, the ends of the individual MCF was fixed both to the

apex of the AFM probe and the bone surface. This gripping of the individual MCF

is somewhat different to the validation testing on synthetic nanofibres in Chapter

3, where fixing of the nanofibre using glue was used for both of the fibre ends. We

note that the manipulation of collagen fibrils requires SEM imaging as opposed to

AFM imaging. AFM imaging is suitable for examining bone specimens where the

collagen fibrils are in the plane of the fracture surface [19, 228] but is unable to

image surfaces where the collagen fibrils are perpendicular to the fracture

surface due to the instability of this surface to imaging using the AFM probe.

Page 152: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

151

Fibrils attached to the AFM probe were used for subsequent tensile testing by

translation of the AFM probe away from the fracture surface. This translation

caused a corresponding bending of the AFM cantilever until failure of the

individual collagen fibril as shown in Figure 4.1(A) inset. The force applied to the

collagen fibril was calculated from the spring constant of the AFM cantilever,

found using the thermal noise method [23] and measured the cantilever

deflection using an optical interferometer setup (Attocube Systems, Ger) situated

behind the cantilever. The calculation of the stress in the collagen fibril requires

the accurate determination of the fibril diameter. The diameter of MCFs was

measured by pixel analysis in the SEM images captured before mechanical test

using ImageJ (NIH, USA). The diameters of collagen fibrils mechanically tested in

this work have a mean value of 92.4 12 nm. All mechanical testing ensured that

the collagen fibril axis is in the same plane as the SEM image, otherwise force

applied to the fibril will cause fibril orientation as opposed to deformation. This

configuration was achieved by moving the AFM probe attached to the collagen

fibril above and below the SEM plane of view. A small force will be recorded

during this out of plane movement if the fibril axis is aligned with the plane. Thus,

only fibrils with their principal axis in the SEM view plane are tested. All fibrils

tested failed in the middle of their free length, away from the holding glue and

the bone surface.

Page 153: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

152

4.2.2 Compositional Study Using EDS

X-ray energy dispersive spectroscopy (EDS) microanalysis within an SEM

(Inspect SEM, FEI Company, EU/USA) was used to investigate the composition of

antler samples used in experiments and verify if the regions containing the

mechanically tested collagen fibrils are mineralized. Chemical composition (Ca/P)

ratios have been previously used as a marker in EDS for the calculation of the

mineral distribution in bone tissue [28, 229, 230] and is thus employed in our

study. Twenty EDS spectra within an area of 100100 m2 were collected at the

tested fracture surface of antler in order to determine the calcium content.

4.2.3 Thermal Gravimetric Analysis

The effect of the SEM vacuum on the hydration of bone samples was examined

using Thermo Gravimetric Analysis (TGA). TGA is typically used to record the

weight of a sample as the sample temperature increases. For bone materials, TGA

has been used to heat a sample and record the weight loss due to removal of

water from the bone structure [102]. Therefore, TGA is a suitable technique to

record the hydration state of bone before and after exposure to the SEM vacuum.

Antler bone samples with dimensions 3100.2 mm were cut from the main

beam of same antler bone used for in situ mechanical testing. Bone samples were

first hydrated by leaving in Hank’s buffered solution for 20 hours. The samples

were removed from solution and excess water removed from the sample surface

Page 154: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

153

using filter paper. Samples were split into four groups, with each group

containing 5-6 bone samples. The first group was transferred into the TGA

furnace immediately and heated from 40 C to 250 C with a heating rate of 30

C .min-1 with 60 mL.min-1 nitrogen flow. The other three groups were moved

into the SEM chamber operating at a pressure of 3.4 10-3 Pa for 15, 30 and 45

minutes respectively, followed by analysis using the TGA.

4.3 Results

4.3.1 Mechanical Behaviour of Individual MCFs

The stress-strain behaviour of 6 individual collagen fibrils was measured using

the AFM-SEM with the results shown in Figure 4.2. All fibrils show a linear

stress-strain response during initial tensile loading up to strains of between 2 %

to 3.7 %. This strain value at the limit of the linear response corroborates

previous deformation of hydrated antler using x-ray studies [154] indicating that

the collagen fibrils are also hydrated despite the vacuum environment of the SEM.

The linear modulus in Region I of the collagen fibril stress-strain curve is highly

reproducible with a value of 2.4 ± 0.4 GPa and is considerably larger than

previous collagen moduli [20, 40], indicating that the fibrils have some degree of

mineralization that improves their stiffness.

Page 155: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

154

Figure 4.2 (A) The stress-strain plot for tensile testing of individual MCFs. The

MCFs show a linear stress-strain regime with a linear elastic modulus of 2.4 ± 0.4

GPa. At higher strains above approximately 2-3 % strain, the MCFs exhibit either

yield or an apparent increase in the tangential modulus. The insert shadow area

shows the inhomogeneity of the strain response with increasing stress. (B)

Failure strength verse ultimate strain recorded by stretching collagen fibrils until

failure. The arrow shows increasing strength along with decreasing fracture

strain, indicating a ‘brittle like’ change.

Further tensile deformation of a MCF caused an observed transition to a

mechanically inhomogeneous Region II. Fibrils in Region II exhibit either yield

behaviour or strain hardening including higher modulus and ultimate strength,

but a decrease in the fibril ultimate strain to failure. No fibril failure was

observed in the SEM image during the tensile testing until catastrophic failure.

The last data point in the stress-strain behaviour of the MCFs from Figure 4.2

(A) (B)

Page 156: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

155

was recorded just before this catastrophic failure. The detailed mechanical

properties of all six collagen fibrils are recorded in Table 4.1.

Table 4.1 Mechanical properties of six individual MCFs tensile tested to failure

using combined AFM-SEM

Test

No.

Diameter

(nm)

Modulus I

(GPa)

Modulus II

(GPa) ƐUTS (%) UTS (MPa)

1 72.9 2.30 1.95 5.14 125.68

2 102.5 2.29 0.96 6.46 89.05

3 90.7 1.91 0.98 6.72 82.22

4 96.0 3.03 3.85 5.81 185.00

5 86.0 2.30 2.78 5.37 129.50

6 106.0 2.46 1.18 6.23 100.66

Mean

SD 92.412.0 2.380.37 1.951.17 5.960.62 118.6837.67

4.3.2 Compositional Analysis

The antler bone mineral content at the fracture surface was evaluated using EDS

following previous works [28, 229, 230]. Preliminary SEM back scattered images

shown in Figure 4.3 indicated regions of high and low mineralization at the

fracture surface. The corresponding EDS analysis of elements present at the

fracture surface, including O, Na, Mg, S, P and Ca, indicated that the calcium

content varied considerably from 31.64 % to 61.34 %. This calcium content

shows that the bone is not only mineralized but there is a large variation in the

mineral content.

Page 157: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

156

Figure 4.3 Scanning backscattered electron micrograph of the antler bone

fracture surface. The light regions on the image indicate higher mineralization.

The corresponding Ca/P ratio in the image varied from 1.48 to 3.12, indicating

hydroxyapatite mineral is present [28, 230] and the content varies throughout

the antler bone.

4.3.3 Water Content Change in Bone Samples Exposed to

SEM Vacuum Studied by TGA

The amount of water in the bone samples calculated from TGA can be plotted

against exposure to SEM vacuum time as shown in Figure 4.4 below.

Page 158: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

157

0 10 20 30 40 50

8

10

12

14

16

18

20

Wa

ter

(%)

Time (mins)

Figure 4.4 Plot of water content in bone with time of SEM vacuum exposure. The

water content in bone was measured from the weight loss recorded during TGA

tests by heating bone samples up to 200 C.

The bone samples directly taken from Hank’s buffered solution contain 15.64

2.60 % of water while bone samples exposed to the SEM vacuum contained

progressively less water, with bone samples exposed to vacuum for 45 minutes

containing 8.79 0.61 % of water. The water content in bone remained at 13.35

1.52 % when exposed to vacuum for 15 minutes, which is the same (within

error) as the water content in bone samples prior to SEM vacuum exposure and

similar to the water content in fresh antler bone recorded in a previous study

[171].

Page 159: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

158

The manipulation and mechanical testing of the MCFs were accomplished within

a time frame of 10-15 minutes after introducing to the SEM vacuum, indicating

that the samples were still hydrated. This low water loss due to the short time of

exposure in vacuum has also been observed in synthetic nanofibres swollen with

water in Section 3.2. In addition, during the mineralization process of CFs, the

water in the gap regions of fibril is replaced by mineral phase which makes MCFs

less hydrated comparing to CFs [45]. We therefore conclude that our

nanomechanical techniques tested bone in its hydrated state. Table 4.2. shows

the average weight of bone samples with exposure to the SEM vacuum and the

weight percentage of water contained in the sample calculated from the removal

of water during TGA.

Table 4.2 TGA test on antler bone samples showing water content changes with

different exposure time to the vacuum in SEM chamber.

Exposure

time (Mins) 0 15 30 45

Weight(μg) 4.151.01 3.240.45 3.720.70 4.130.91

Water (%) 15.642.60 13.351.52 9.840.94 8.790.61

Page 160: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

159

4.4 Discussion

The deformation behaviour of individual collagen fibrils in tension described by

Region I and II are indicative of structural or compositional variation within the

fibrils themselves. A mechanistic description of the deformation of individual

collagen fibrils can be made by comparing our experimental results in this work

with molecular dynamics simulations [138]. These simulations show some

similarity with our individual MCF tensile tests, with an initial linear response

followed by heterogeneous deformation in the MCF stress-strain behaviour. The

initial linear behaviour was mostly from elastic behaviour of the MCF

constituents and the tensile modulus calculated from this Region I was

dependent on the amount of mineral contained within the fibril. The presence of

mineral phase increases the local yield regions in MCFs when tensile load is

applied, which suggests the MCFs fail locally to ensure that the majority of the

fibril length remains undamaged after exposure of the fibrils during sample

preparation [138]. The linear stress-strain fibril response of the MCFs in Figure

4.2 after primary mechanical tests provided an experimental validation of this

mechanism as testing of fibrils after yield will give a non-linear response.

The heterogeneous deformation zone, defined by Region II in this work, was

shown in the molecular simulation to be also due to the amount of mineral in the

Page 161: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

160

collagen fibrils [43, 138]. The mineralized collagen fibrils in the simulation work

displayed a characteristic stress-strain curve that is similar to Figure 4.5 (B) here,

indicating that deformation in Region II is due to intermolecular slippage and

failure at the mineral-tropocollagen macromolecule interface. We note that other

simulations show how intermolecular slippage in non-mineralized collagen

fibrils are suppressed when crosslink density is increased [105]. The

tropocollagen molecules in Buehler’s simulation work [138] do not vary the

crosslink density between these tropocollagen molecules. However, variations in

the mineral content within the MCFs may define the two observed stress-strain

behaviours in Figure 4.5, with the amount of mineral analogous to the

crosslinking behaviour in the simulations. A relatively large mineral content as

shown in Figure 4.5 (A) may enhance the binding between tropocollagen

molecules, causing molecular extensions at relatively high strains and an

increase in the tangential modulus. A relatively low amount of mineral as shown

in Figure 4.5 (B) conversely weakly binds the tropocollagen, with sliding

between the tropocollagen and resultant fibril yield observed.

Page 162: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

161

Figure 4.5 Tensile stress-strain curves for mineralized collagen fibrils showing

two distinct mechanical behaviours. Tropocollagen uncoiling occurs initially in

both types of fibrils within Region I. In (A) the fibril shows an enhanced elastic

modulus in Region II due to the mineral increasing the stress transfer between

tropocollagen macromolecules. In (B) the low mineral density within the fibrils

allows sliding between tropocollagen molecules, resulting in a plastic

deformation.

The proposed variation in mineral content causing the two observed MCF

stress-strain behaviours can be evaluated from EDS investigations within the

bone samples. The compositional EDS analysis in Section 4.3.2 shows a variation

mineral content throughout bone that suggests a corresponding variation in the

collagen fibril mineral content. Table 4.1 supports this assumption that the

heterogeneous fibril strain comes from the compositional variation with collagen

fibrils with a higher tensile modulus in Region I, as defined in Figure 4.5 (A), also

exhibiting an increased tensile modulus in Region II. The increased tensile

(A) (B)

Page 163: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

162

modulus in Region II relative to Region I cannot be simply due to mineral content

and must be due to a stiffening of the fibril with strain. Classical work from

tendon, essentially an aligned unmineralized CF array, shows how crimping of

tropocollagen molecules is removed with increasing strain, resulting in an

increased tensile modulus [105, 155, 231]. The increased tensile modulus in

MCFs as shown in Figure 4.5 (A) may therefore be due to the removal of

crimping in the MCF. However, molecular dynamics simulations do not show this

behaviour as the model assumes an uncrimped, uniaxially aligned tropocollagen

network [138]. The stress-strain curves recorded in MCFs tensile testing are

similar to the simulated stress-strain behaviour for uncrimped MCFs. The

mineral phase is acting as ‘crosslink’ between tropocollagen molecules to

enhance the fibril tensile modulus.

While the nanomechanics of collagen fibrils are expected to be due to mineral

content, there are some other alternative mechanisms could lead to the different

mechanical response of MCFs in the region II of the stress-strain curves. In the

first region of the stress strain curves, all the fibrils exhibit a similar response to

the applied load, which could be explained by the distinct mechanics of the

tropocollagen molecules and the apatite crystals within the fibrils. The tangential

modulus of the MCFs in the first region is defined by the interaction of the

tropocollagen and mineral phase which is indicative a possible similar degree of

Page 164: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

163

mineralization. The various responses recorded in the region II could come from

the different interactions between mineral crystals due to the different

distribution of the minerals within collagen when tropocollagen molecules start

to slip between each other under strain. Mineral crystals in strain-hardened

fibrils are brought together and contact to each other due to the slippage of

tropocollagen molecules and start to bear load to provide a higher resistance to

strain while this interaction of minerals may be absent in yield fibrils. In addition,

the change in tropocollagen residue sequences [41, 151, 160, 232],

hydroxyapatite crystal shapes [233, 234] as well as their texture and orientation

[233] are considered to have some degree of influence on the mechanical

behaviour of individual MCFs but are not as significant as the mineral

component.

The implications of nanomechanical heterogeneity have been more widely

studied and previous work has illustrated how heterogeneous deformation in

mineralized tissue aids energy dissipation [31, 122, 162]. The principal

mechanical function of antler is energy absorption during impact and therefore

promotion of heterogeneous deformation is favourable. Thus, our work

highlights how heterogeneous deformation originates from the nanomechanical

behaviour of the MCFs themselves.

Page 165: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 4. Nanomechanical Properties of Individual MCFs from Bone Tissue

164

4.5 Conclusions

In conclusion, the stress-strain behaviour of individual MCFs from antler bone

was measured using combination AFM-SEM. An initial region of homogeneous

fibrillar deformation was succeeded by inhomogeneous mechanical behaviour

above applied strains of 2-3.7 %. A molecular mechanism is proposed to explain

these different fibril mechanical responses to external load. The nanomechanical

testing technique developed in this work may also be applicable to the

measurement of bone at different mineralization and diseased states.

Page 166: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

165

Chapter 5

Mechanical Properties of MCFs from

Differing Ages of Bone

5.1 Introduction

The capacity to measure the mechanical properties of individual MCFs using

novel experimental AFM-SEM techniques has been defined in the previous two

chapters. This Chapter aims to address the ability of the technique to elucidate

potential differences that may exist in MCFs. To further this aim, MCF fibrils from

bone of different ages are considered due to bone being a biological material with

capability of renewing and remodelling its structure over time defining its

material properties [48, 51, 56]. The effect of this structural renewing and

remodelling in bone on MCF mechanics can therefore be assessed. The

mechanical properties of bone are generally known to change with age, with

Page 167: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

166

bone mineral density (BMD) typically used to describe the susceptibility of

different ages of bone to [235-237]. Considerable debate remains as to whether

the fracture behaviour of bone can be reliably predicted by assessing bone

density alone [238-243]. Recent clinical observations indicate a correlation in

bone mineral density of healthy people and patients who suffer fractures due to

age related diseases [238]. Moreover, experimental and simulation studies on

changes in collagen materials with age also show the limit of considering bone

mineral density alone in the mechanical study of aging bone [79, 104, 137, 156,

244]. An important aspect of bone tissue quality is the relative amounts and the

material properties of its main constituents of collagen fibrils and apatite crystals.

The influence of apatite/collagen ratio on mechanics of bulk bone material in

aging has been widely studied [245-248], although the basic building block of

MCFs from different ages has yet to be mechanically assessed. The evaluation of

the mechanical properties of MCFs from different ages of bone may therefore be

beneficial in potentially defining overall bone mechanical behaviour or providing

further understanding on the influence of MCFs in the changing mechanics of

whole bone. In this chapter, MCFs taken from the limb bones of 4 and 10 week

old mice are mechanically tensile tested to failure using in situ AFM-SEM

nanomechanical testing methods as described in Chapter 4. The mechanical

behaviour of individual MCFs from different ages of mouse bone was compared

to results from bulk samples. Thus, the effects of aging on bone nanomechanical

Page 168: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

167

behaviour can be elucidated.

5.2 Materials and Methods

5.2.1 Animals

C57BL/6J wild type mice (4 and 10 weeks) were freshly obtained from the

Medical Research Council (MRC, UK). Mice were kept in accordance with UK

Home Office welfare guidelines and license restrictions.

5.2.2 Sample Preparation of Bone Tissue

The right and left femur from each animal were exercised and cleaned to remove

adhering soft tissue from the femur limbs. As recorded in Section 3.1.2, mouse

femur bones were kept wet in Hank’s buffered solution during preparation and

stored frozenly. Bulk mechanical testing on whole mouse bone femur was

performed by fixing the distal and proximal end lobes in dental cement (FiltekTM

Supreme XT, 3M ESPE, USA) to grip the samples in a custom built testing device

(M110.1DG, Physic Instrumente, UK).

Page 169: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

168

Figure 5.1 Sample preparation for tensile testing. (A) Schematic of mouse

skeleton showing the location of the femur (ellipse). (B) Optical image of an

isolated femur bone. (C) Custom made micro milling setup used to remove bone

leaving 0.1 mm thick of anterior quadrants. (D) Bony ends of Femur included in

the dental cement (FiltekTM Supreme XT, 3M ESPE, USA) and milled mid

diaphysis.

Embedded femurs were machined to form a necked region in the mid diaphysis

1 cm

High speed

dremel drill

Specimen mounting chamber

Dental

cement

Femur

Milled

sample

(A)

(B) (C)

(D)

Page 170: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

169

using a custom-made micromilling machine. The milling machine consisted of

two motorised linear stages (M110.1DG linear stages Physic Instrumente, UK),

which are connected to the computer via 2 motor controllers. A high speed

rotatory milling tool (Dremel 300, Dremel Inc, UK) with 0.8 mm diameter cutting

tool was fixed and specimen was held in a Hank’s buffered solution fluid chamber

connected to the linear stage as shown in Figure 5.1(C). Specimens were

machined from the posterior quadrants leaving a 100 µm thick anterior quadrant.

The width and the thickness of each specimen were measured after milling using

a Basler A101f monochrome CCD camera (Basler Vision Technologies, Ger). High

resolution (1024 768) optical images of the milled specimens were captured,

transferred to computer and dimensions measured using ImageJ software (NIH,

USA). The typical dimensions of the gauge regions were approximately 0.10 mm

(thickness), 1.0 mm (width) and 4.0 mm (length). SEM was used to examine if

bone samples were damaged during the machining protocol and damaged

samples were discarded.

5.2.3 Tensile Testing of Bulk Bone Tissue

A customized tensile testing machine as shown in Figure 5.2 was used to deform

the femur samples by applying bi-directional motion of two DC encoder stages

(M110.1DG, Physic Instrumente, UK), which has a minimum incremental motion

of 0.1 µm.

Page 171: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

170

Figure 5.2 Schematic view of the micro tensile testing machine. The sample is

immersed in Hank’s buffered solution. Tensile load is applied along the

schematically prepared femur long axis as shown in the inset.

The main features of this tensile tester are (i) bi directional tensile testing

capability and (ii) sample testing in a fluid chamber (this provides physiological

environment to the sample during the experiment). Load was measured with a

22 N model 31 tension/compression load cell (SLC31/00005, RDP electronics,

UK) and sample holders were moved with linear stages. Due to an initial

slack-range in the grips, the stress-strain curve shows an initial toe-in region,

which was not considered during the data evaluation. A consistent strain rate of

Inside view of the

fluid chamber

Sample

holders

Load cell

Femur

Fluid

chamber

Page 172: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

171

0.02 %.s-1 was applied on the bone sample by moving bi-directional stages in

opposite direction at a velocity of 0.001 mm.s-1. 6 samples taken from 4 week old

mouse femur bone and 4 samples from 10 week old were tensile tested up to

applied strains of 3 %. Resultant stress-strain plots were recorded.

5.2.4 in situ Nanomechanical Testing of MCFs

The fresh radial bones from 4 and 10 week old mouse bone were extracted and

prepared as described in Chapter 3, Section 3.1.2. MCFs exposed at the fracture

surface were tensile tested to failures by in situ AFM-SEM setup reported in

Chapter 3, Section 3.2. To determine the stress-strain relation of samples, the

diameters and lengths of the MCFs were accurately measured by pixel analysis

using high magnification SEM secondary electron imaging taken before and after

nanomechanical experiments using ImageJ (NIH, USA). Resultant stress-strain

plots from tests on 6 MCFs from 4 week old mouse bone and 5 MCFs from 10

week old mouse bone were recorded.

5.3 Results

The mechanical stress-strain responses for tensile tested bulk bone samples are

shown in Figure 5.3 (A). All bone samples show a relatively linear increase in

stress with applied strain in these plots. The stress-strain behaviour for 10 week

Page 173: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

172

old mouse bone showed a larger linear (Young’s) modulus of approximately 4.5

GPa when compared to 1.2 GPa for 4 week old mouse bone as shown in Figure

5.3 (B).

Figure 5.3 Mechanical behaviour of femur samples from 4 and 10 week old

mouse bone recorded during tensile testing showing (A) stress-strain plots of

bone samples with linear regressions and (B) summary plot showing the Young’s

modulus calculated from stress-strain plot linear regressions for the 4 and 10

week old bone.

Stress-strain plots for individual MCFs from 4 and 10 week old mouse bone

tensile tested by the AFM-SEM are shown in Figure 5.4.

Strain %

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Stre

ss (

MP

a)

0

20

40

60

80

4 weeks 4 weeks - Linear regressions

10 weeks 10 weeks - Linear regressions

Age 4 weeks 10 weeks

Yo

un

g’s

Mo

du

lus

(GP

a)

0

1

2

3

4

5

6

7

(A) (B)

Page 174: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

173

0.00 0.02 0.04 0.06 0.08

0

50

100

150

200

Str

ess (

MP

a)

Strain

4 weeks

10 weeks

Figure 5.4 Stress-strain behaviour of individual mineralized collagen fibrils from

mouse radius bone at 4 and 10 weeks respectively.

All fibrils show a linear stress-strain response during initial tensile loading up to

strains of between 2 % to 4 %, followed by a more variable stress-strain

behaviour beyond strains of 4 % as described for individual MCFs examined in

antler bone. Young’s moduli are calculated from this initial linear region and are

similar in both bone age groups, displaying values of 2.54 ± 0.60 GPa and 3.65 ±

0.81 GPa respectively. The calculated mechanical properties of the testing

individual MCFs is shown in Table 5.1 below and indicates an additional

similarity in the ultimate strength and strain in both age groups.

Page 175: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

174

Table 5.1 The mechanical behaviour of MCFs tensile tested to failure.

Age Length

(μm)

Diameter

(nm)

Strength

(MPa)

Ultimate

Strain

E (GPa)

4 weeks 1.25 92.09.3 138.43 0.052 2.86

2.55 118.89.7 89.05 0.065 2.34

1.67 103.810.2 82.22 0.067 1.48

1.71 110.412.5 185.25 0.054 2.83

1.27 93.110.5 129.50 0.054 3.41

1.92 109.58.7 100.10 0.062 2.35

MeanSD 1.730.44 104.69.6 120.7635.27 0.0590.006 2.540.60

10 weeks 1.02 97.29.1 113.73 0.070 4.12

2.08 118.76.3 131.89 0.078 2.36

2.45 105.214.6 154.10 0.068 3.51

1.57 89.38.6 188.94 0.055 4.81

1.48 108.17.5 204.66 0.041 3.44

MeanSD 1.720.50 103.710.0 158.6634.03 0.0620.013 3.650.81

5.4 Discussion

The mechanical behaviour of the bulk bone material and the MCFs can be

compared in order to ascertain the influence of the nanoscale on larger scale

bone mechanics. The linear stress-strain behaviour average over all of the

collected experimental data for the individual MCF and bulk bone at different

ages is plotted in Figure 5.5 below. The variation in Young’s modulus for bulk

bone samples with age is clearly observed in Figure 5.5 (A). This variation is

expected to due to the different composition of bone developing from 4 to 10

weeks. Critically, the increase in the Young’s modulus for bulk bone with

Page 176: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

175

increasing aging is not observed for the average stress-strain behaviour of MCFs

in Figure 5.5 (B). Within error, the stress-strain response of MCFs from 4 and 10

week old bone are comparable, indicating a weak or absent effect of increasing

age. Our results therefore indicate that the mechanical properties of the MCF unit

in bone do not define the mechanical properties of bone at larger length scales as

represented by the bulk bone data.

Figure 5.5 Linear regressions of average stress-strain plots for bulk bone and

MCFs samples from 4 week and 10 week old mouse bone. Error bars are

standard deviations (see Table 5.1).

Previous studies have shown how mechanical properties of bone at different ages

depend on the mineral content [51, 244, 249]. To obtain the information of

mineralization of bulk bone samples at different ages, compositional analysis

0.0 0.4 0.8 1.2 1.6

0

20

40

60

80

Str

ess (

MP

a)

Strain (%)

4 weeks

10 weeks

0.0 0.4 0.8 1.2 1.6

0

10

20

30

40

50

60

70

Str

ess (

MP

a)

Strain (%)

4 weeks

10 weeks

Bone MCFs

Page 177: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

176

using Quantitative Backscattered Scanning Electron microscopy (qBSE) was

performed. The qBSE images were collected from the mid shaft of the cortical

bone from 1 week to 16 weeks (left to right in Figure 5.6) for wild type mice. A

significant increase in mineralization shown as increased brightness in Figure 5.6

is observable, especially for the bone aging from 1 to 10 weeks.

Figure 5.6 BSE microscopy images of mid shaft cross sections of limb bones from

1 week to 16 weeks for wild type mice.

The increase in the Young’s modulus of bulk bone must therefore be related to

the increase in the degree of mineralization present on the bone as shown in

Figure 5.6 and observed in previous literature [1, 2, 51, 86, 132, 230, 244, 249,

250]. In particular, the stiffer mineral phase has been shown to influence the

Page 178: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

177

elastic properties and fracture behaviour of bone when varying mineral

percentage, orientation and morphology [1, 2, 15, 48, 56, 86, 132, 249] and has

been indicated as being the dominant phase in defining bone mechanics [1, 2, 56,

249, 251]. However, this increase in the mineralization of whole bone with age

does not show a corresponding increase in the Young’s modulus of individual

MCFs from the mouse bone, despite the expected mineral content influence on

fibril mechanics as discussed in Chapter 4. The slight increase in MCF Young’s

modulus from 4 to 10 weeks in Figure 5.5 is much smaller than in bone tissue. In

addition, a molecular dynamics simulation study on mechanistic description of

the deformation of individual MCFs showed that the Young’s modulus of MCFs is

dependent on the amount of mineral that exists within the fibrils [138] due to

the mineral phase increasing tropocollagen crosslinking, which promotes load

transfer within fibrils and leads to increased bone stiffness.

The discrepancy between increased mineralization increasing bulk bone Young’s

modulus but not the MCF Young’s modulus can be evaluated further by

considering the BSE microscopy images in Figure 5.6. Calcium concentration was

used to represent Bone Mineral Density (BMD) and was calculated from the grey

scale of each individual pixel in the BSE microscopy images of Figure 5.6,

baselined against standard reference materials of carbon and aluminium, as

carried out in previous literature [229]. Calcium concentration within each pixel

Page 179: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

178

in the BSE microscopy image is determined from its grey scale comparing to

reference materials. The ratio between the number of pixels with certain calcium

concentration and total number of pixels in one image is defined as the

frequency of appearance and used to evaluate the Bone Mineral Density

Distribution (BMDD). The Full Width at Half Maximum (FWHM) of the BMDD is

taken as the width at the half height of peaks in the diagram of frequency of

appearance verse calcium concentration for examining the distribution in local

variation of mineral content. A high FWHM represents a large variation in BMDD

within bone tissue.

Figure 5.7 highlights how a significant reduction in the distribution of calcium as

defined by the Ca FWHM values is observed from 1 to 4 weeks and stabilizes

from 4 weeks to 16 weeks. Table 5.2 summarizing the data presented in the

figure. In contrast, a dramatic increase in average calcium concentration was

observed from 1 to 4 weeks and gradually stabilizes from 4 weeks to 16 weeks.

These two observed phenomena are described by considering the mineralization

process in bone. In primary ossification, the mineralization occurs mainly in the

extrafibrillar regions as the narrow collagen fibrils formed have relatively limited

space for mineral to deposit. In the later secondary ossification, the collagen

fibrils have much larger diameters to allow mineral palettes to form within

internal spaces. However, the initial location of mineral deposition i.e. within the

fibrils or the intrafibrillar regions in secondary ossification is unclear [18, 83].

Page 180: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

179

Figure 5.7 Quantitative analytical diagrams of BSE microscopy images taken from

bone of different ages. (A) Frequency of Appearance against calcium

concentration with age. (B) Histogram of FWHM of BMDD plotted with

development which corresponds to the distribution in local variation of Calcium

content and (C) Mean calcium concentration in form of weighted fraction (Ca

Fre

qu

ency

of

Ap

pea

ran

ce

(N p

ixel

s/

N p

ixel

s)

Calcium concentration (wt%)

(A)

(B) (C)

Page 181: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

180

wt %) calculated from (A) was plotted as a function of development (1 week to

16 weeks).

Table 5.2 Quantitative analysis of qBSE images taken on bone at different

development stages (1 to 16 weeks).

Age (weeks) Ca % Mean Ca FWHM Ca % Peak

1 17.36 17.67 18.5

4 23.85 5.86 27.67

7 25.15 5.12 29.81

10 26.09 4.85 28.70

16 27.86 4.65 27.70

In ossification, the organic matrix acts as a template to induce inorganic crystal

growth. The organic matrix therefore promotes heterogeneous nucleation of

apatite and inhibits homogeneous nucleation [51, 83] by providing large

contacting area between extracellular fluid containing mineral ions and collagen

matrix. The formation of calcium apatite crystals occurs randomly in the

mineralization front area [12, 45, 51, 55] so a random distribution of nucleation

sites throughout bone tissue will occur, exhibiting a large local variation of

calcium concentrations in early stage development of bone as indicated in the

large Ca FWHM for 1 week old bone in Table 5.2. As the collagen fibril widths

produced in secondary ossification are significantly larger than the extrafibrillar

Page 182: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

181

spaces (~100 nm fibril diameters comparing to 1-2 nm spacing between fibrils),

mineralization in secondary ossification occurs initially in the relatively large

spaces found in the gap regions within collagen fibrils where water is replaced by

hydroxyapatite (HA) crystals [12, 18, 33, 45, 50]. Once the gap regions are filled

with mineral within a fibril, we propose that the HA mineral can only form

between the collagen fibrils in the spacing as shown in Figure 5.8:

Figure 5.8 Schematic sketches showing three stages of biomineralization on

collagen fibrils in mouse bone. Tropocollagen molecules are denoted as blues

cylinders and the hydroxyapatite mineral crystals are red particles with random

shapes. Negatively charged molecules (most of which are Non-Collagenous

Proteins as discussed in Chapter 2) existing in the extrafibrillar space are shown

in green and bonded via positive calcium ions. (A) Calcium and phosphate ions

migrate to the gap regions (initially filled with water) within collagen fibrils and

(A) (B) (C)

1 week 4 weeks 10 weeks

Page 183: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

182

initiate nucleation of HA crystals. (B) All the gap regions within collagen fibrils

are completely filled with mineral crystals. (C) Mineral continues to grow within

the extrafibrillar space and cover the fibril surfaces.

We can therefore assume that, in this study, the 1 week old mouse bone is at the

first stage of mineralization. The mineral crystals form predominantly and

randomly in the gaps regions found within the collagen fibrils. Relatively large

variation in local bone mineral density distribution (BMDD) but low average

calcium concentration is therefore observed. Beyond 1 week and up to 4 weeks

the mineral continues to deposit within the gap regions of the fibril (shown in

Figure 5.8 (B), resulting in an increase in the Ca concentration for 4 week old

bone as shown in Table 5.2. Beyond week 4, the gap regions within the collagen

fibrils are saturated with mineral and the process of mineralization slows as the

mineral ions become more difficult to nucleate in the intrafibrillar spaces. This

slowing of mineralization results in more stable BMDD and mineral

concentration values beyond week 4. While mouse skeleton is considered as

mature at 10 to 12 weeks [38, 68, 244], our results indicate that mouse bone

becomes mature beyond 4 weeks. Thus, this phenomenological model indicates

that the MCFs from 4 week old mouse bone are effectively fully mineralized

within their gap regions. MCFs from 10 week old mouse bone will show no

further increases in their mineralization, resulting in similar Young’s modulus

Page 184: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

183

values when compared to MCFs tested from 4 week old mouse bone. However,

the mineralization occurring between the MCFs is apparently responsible for

increases in bulk bone Young’s modulus. Structural difference at the nanoscale

between 4 and 10 week old bone occur due to the formation of extrafibrillar

mineral. Consideration of the space between collagen fibrils filled with

non-collagenous proteins (NCPs) may be the key to understanding the increased

Young’s modulus of bone at 10 weeks. In particular, the extrafibrillar space

contains mineral that provides more efficient linkage between fibrils and

therefore improved the load transfer between fibrils leading to a higher Young’s

modulus. Furthermore, the Young’s modulus recorded in 10 week MCFs from

mouse radial bone is comparable to the value of bone tissue at same age (3.65

1.8 GPa versus 4.7 1.3 GPa) indicating an effective stress transfer between

fibrils in bone. As the extrafibrillar space is relatively limited with only 1-2 nm in

thickness [152, 162], it is reasonable to assume the mineral crystals on the fibril

surfaces might contact mineral in adjacent fibrils to form a rigid network of

apatite mineral [252]. The increased connectivity between fibrils promotes the

stress transfer and may lead to a nanoscale stiffening mechanism.

On the other hand, the MCFs and bone tissue tested from antler bone showed two

different Young’s modulus values but in same order as recorded in Chapter 4 (2.4

0.4 GPa of MCFs versus 5 to 7 GPa for antler bone tissue [103, 170, 171]). The

Page 185: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

184

difference between measured Young’s modulus of MCFs and bone tissue could

arise from apatite mineral at the fibril surface connecting together leading to an

enhanced network stiffness as discussed for mouse bone above. Indeed, mineral

present at the surface of individual MCFs from antler may not contribute to the

mechanical behaviour measured using AFM-SEM as poor stress transfer between

the straining fibre and surface mineral may exist. This may indicate a potential

underestimation in mechanical properties of isolated MCFs compared to arrays

of MCFs found in bone where stress transfer can occur between all constituents.

The measurement of the Young’s modulus may also give rise to errors between

individual MCFs and bulk bone samples. Bone is a type of viscoelastic material

having a nonlinear stress-train response. The Young’s modulus is measured as

the tangent slope of the stress-strain curve at relatively small strain [12, 103, 109,

112]. However, in this thesis, the Young’s modulus of fibrils is calculated by linear

regression of the data points on the stress-strain curve due to the limited data

collected by AFM. These two measurement methods record different modulus

values as shown in Figure 5.9. In fact, some fibrils exhibit a Young’s modulus

higher than 8 GPa by choosing the first two points for linear regression. The

Young’s modulus of mouse bone tissue and corresponding MCFs were measured

using the same linear regression method and therefore have comparable values.

However, other bone samples may differ considerably due to the assignment of

this tangential slope in stress-strain curves.

Page 186: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

185

Figure 5.9 Young’s modulus measured by different methods. (A) Tangent slope of

stress-strain curve at small strain is recorded as Young’s modulus. (B) Young’s

modulus is measured as the linear regression of data points on a stress-strain

curve.

5.5 Conclusions

To conclude, 4 week and 10 week old mouse limb bones were representatively

chosen for evaluating mechanical properties of bone and MCFs at different ages.

Conventional mechanical testing methods were used to measure the mechanical

behaviour of bulk bone and stress-strain plots were recorded. Individual MCFs

from 4 and 10 week old mouse bone were tensile tested to failure using

AFM-SEM techniques demonstrated in Chapter 4. In addition, an increase in the

degree of mineralization at increasing bone age was examined by qBSE.

Stre

ss

Strain Strain

Tangent

modulus

Linear regression

modulus

Stre

ss

(A) (B)

Page 187: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 5. Mechanical Properties of MCFs from Different Ages of Bone

186

Homogeneity of mineralization and the bone mineral density were found to be

increasing with age. The Young’s modulus of both bulk bone and MCFs was found

to increase with age, with the increased amount of mineral in the older bone

responsible for this increase. However, the increase in the Young’s modulus for

MCFs was significantly less than for bulk bone tensile testing. A mechanism was

developed that considered the mineralization of collagen fibrils and suggested

that the mineral content in MCFs from 4 and 10 week old bone is compositionally

similar. The interphase region containing mineral between collagen fibrils is

described as varying with bone age and is more critical in defining the

mechanical properties of bulk bone. To study the extrafibrillar region further, the

NCPs layer between MCFs is mechanically evaluated in the next chapter.

Page 188: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

187

Chapter 6

Nanoscale Interfacial Behaviour in Bone

6.1 Introduction

The mechanical performance of bone has been evaluated throughout the thesis in

terms of the collagen fibril units. As with any composite material described in

Section 2.3.2 of Chapter 2, the mechanical properties of a composite are defined

not only by fibres but also by the matrix and fibre-matrix interface. Bone is

perhaps unique in being a composite material containing a high volume (> 80 %)

of fibrous material [12], resulting in a relatively small phase between MCFs.

Despite the relatively small volume fraction of the matrix material, the interface

formed between MCFs and NCPs should influence overall bone behaviour from

composite considerations. While interfaces in bone at constituent length scales is

the subject of this Chapter, interfacial behaviour in a number of layered biological

Page 189: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

188

structures, especially shell materials, have been previously evaluated [253-255].

The propensity for interfacial failure in biological layered composites and the

large fracture area involved has led to the pursuit of novel biomimetic layered

composite structures based on shell materials fabricated with high volume

fractions of ceramic mineral phase within a polymeric matrix [256-258]. These

biomimetic composites show unprecedented work of fracture values due to

extensive ductility and fracture at nanometre length scales. Bone is notable as a

composite structure that exhibits considerable toughness through using fibrous

constituents of MCFs as opposed to the layered platelets. The origin of bone

toughness is therefore contentious and potentially different to shell materials,

with hierarchical deformation over a range of length scales [7, 259],

microcracking mechanisms [128, 132], heterogeneous failure [31, 260] and

distinctive load transfer between constituents [261] proposed as defining bone

toughness. Many of the proposed failure mechanisms in bone depend, at least in

some part, on the mechanical behaviour of the nanomaterial constituent

properties found in bone. These bone constituents can be evaluated in composite

terms as a high volume fraction of collagen nanofibres reinforced by mineral

platelets of hydroxyapatite. The mineral is often found in the collagen fibrils

themselves but can be extrafibrillar. The critical material constituent not

considered in previous Chapters but present in what should be considered as the

interphase region, describing both the interface and the matrix phase between

Page 190: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

189

MCFs, is the non-collagenous protein (NCP) region. The NCP region found within

the relatively small spaces of 1-2nm between collagen fibrils is amorphous and

includes a number of proteins, most notably osteopontin [75, 262] and

proteoglycan tethers between collagen fibrils [231]. A number of studies have

examined the mechanical properties of both unmineralized collagen fibrils [20,

40] from various sources and mineralized collagen fibrils from bone tissue [17,

260]. However, the NCP region between the collagen fibrils is expected to be

critical in defining the toughness of whole bone but is often poorly understood.

Recent work has indicated deficiency of specific proteins in NCPs cause loss of

bone strength, while the transfer of stresses between collagen fibrils is expected

to be critically dependent on the mechanical behaviour of the NCP region [231,

263]. Indeed, classical mechanical analysis of composites of fibres bound

together by a polymer matrix highlights the importance of the fibre-matrix

interfacial adhesion on stress transfer. Composite theory has been extensively

exploited in nanocomposite interfacial mechanics, such as efficient stress

transfer at carbon nanotube-polymer interfaces [172, 264] or poor stress

transfer in graphene-polymer interfaces [265], but is rarely utilized in

understanding the effect of nanoscale interfaces between the collagen fibrils in

bone. Interfacial mechanical considerations in bone have been made [125, 261]

but lacks evaluation of collagen fibril-NCP interface mechanics.

Page 191: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

190

The mechanical characteristics of the NCP interface region between collagen

fibrils in bone therefore remains elusive despite its importance in bone and other

fibrous material toughness. Direct evaluation of nanoscale interfaces has been

achieved using advanced atomic force microscopy (AFM) techniques, utilized to

manipulate and remove nanofibres partially embedded within a polymeric

matrix material [172, 264]. These direct nanofibre pullout measurements give

quantitative information on the mechanical behaviour of the interface between

the nanofibre and matrix, allowing the evaluation of both strong and weak

nanocomposite interfacial mechanics. This Chapter therefore exploits the direct

mechanical testing ability of the AFM-SEM technique to evaluate the interfacial

properties at collagen fibril-NCP interfaces in bone using a pullout configuration

to provide understanding on the role of nanoscale interfaces in bone toughness.

6.2 Materials and Method

6.2.1 Material Preparation

Antler bone samples were prepared using methods described in Chapter 3 and 4.

Briefly, samples were extracted from the main beam of antler from a mature red

deer and the velvet removed. All samples were selected from the same compact

cortical shell near the antler-pedicle junction. Small beams of antler with

dimensions of 3200.2 mm with the long axis parallel to principal osteonal

Page 192: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

191

direction were cut from the bulk material using a water-cooling rotating diamond

saw and stored in 70% ethanol solution. Before mechanical evaluations, samples

were left in Hank’s buffered solution overnight to allow full sample rehydration

that mitigates mineral loss, which may occur in distilled water or physiological

saline. Water on the surface of the sample was removed by filter paper to avoid

interference with SEM imaging. Hydrated antler bone samples were

subsequently fractured perpendicular to their long axis to expose mineralized

collagen fibrils and immediately transferred to the chamber of an SEM containing

the AFM setup.

6.2.2 in situ AFM-SEM Fibril Pullout Experiment

Bundles of fibrils at the bone fracture edge were selected with an individual fibril

protruding from the centre of the bundle as shown in Figure 6.1. Mechanical

testing of the NCP was achieved following a pullout configuration, as has been

achieved in synthetic fibrous nanocomposites [164, 172]. The AFM probe within

the SEM chamber was first translated into a glue (Poxipol, Arg) droplet

containing within the chamber, to allow pickup of glue at the apex of the AFM

probe. The AFM probe was then translated towards the free end of the MCF

protruding from a bundle as shown in Figure 6.1.

Page 193: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

192

Figure 6.1 (A) Scanning electron micrographs showing an AFM probe containing

glue at its apex attached to an individual MCF partially embedded in a fibril

bundle at the fracture surface of antler bone. (B) The exposing fibril was pulled

out by the AFM with images at high magnification used to measuring the fibril

embedded length le.

The attachment of the free end of the exposed individual MCF to the AFM probe

containing glue was performed within a 10 minute time window to ensure that

the glue was still liquid during the MCF attachment. The AFM probe was

subsequently moved away from the bone surface, which caused an increase in

the tensile stress within the fibril and an equal but opposite shear stress within

the NCP surrounding the MCF. Fibrils were observed to detach from the bone

surface at a critically applied force, measured using the optical interferometer

setup to record the deflection of the AFM cantilever in the AFM system as

(B) (A)

10m

AFM cantilever

probe

epoxy glue

bone le

3m

Page 194: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

193

reviewed in Section 3.2. These MCF pullout experiments were performed in the

SEM vacuum chamber for less than 20 minutes to ensure that water loss from the

bone was minimized. Previous mechanical testing on bone constituents in SEM

described in Section 3.2 highlighted how bone at nanometre length scales retains

the same mechanical properties as fully hydrated bone samples [260] within the

testing time frame used here. We note that the selection of an individual MCF for

pullout was not controlled as the length of the MCF embedded within the bone

tissue was not known. Therefore, only around 2 in 10 individual MCFs selected

for pullout actually became pulled out of the bone tissue, with the other tested

MCFs fractured within the free length part.

6.3 Results and Discussion

Individual mineralized collagen fibrils were pulled out from rehydrated bone

sample using in situ AFM-SEM. The mechanical properties of the NCP interphase

region around the MCFs is calculated by recording the force applied to the MCF

by the AFM system. As shown in Figure 6.1, a force F is applied at the free end of

MCF with the effective force parallel to the protruding MCF long axis calculated

accurately by knowing the off-axis translation angle θ (<30). The force required

to pull the MCF out of the bone surface Fp = F cos θ. The force applied on the

Page 195: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

194

MCFs and the displacement of the AFM cantilever during the pullout was

recorded and shown in Figure 6.2 for tests on five different collagen fibrils within

the same bone region. The force applied to the MCF increased linearly with

progression time of the experiment until a maximum force, Fp, was reached,

which caused failure of the MCF-NCP interface and a rapid drop in the force F

exerted by the AFM until the MCF was separated from the bone sample. A linear

increase of applied force F with experimental progression time has been

observed for other nanofibre pullout experiments [264, 266-269], indicating that

the MCFs in this work pullout as opposed to fracture within the bulk bone

material and subsequent pullout.

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.00

0.05

0.10

0.15

0.20

0.25

0.30

Fo

rce (

N)

Time (s)

Figure 6.2 Plot showing the force applied to the partially exposed MCF during

pullout against progression time for the pullout experiment. The force increases

Page 196: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

195

linearly with progression time until a maximum force Fp is reached, which causes

failure of the interface and rapid separation of the MCF from the bulk bone

sample.

The strength of the interface between the MCF and surrounding NCP is

characterized by the interfacial shear strength (i) and is calculated from the

maximum force applied to the exposed MCF to cause pullout from the

surrounding NCP:

𝜏𝑖 = 𝑝

𝐷f 𝑙𝑒 Equation 6.1

where Df is the fibril diameter and le is the length of fibril embedded within the

bone. Equation 6.1 assumes the stress generated at the MCF-NCP interface

during pullout is uniformly distributed along the fibril embedded length. The

evaluation of the interfacial shear strength at the MCF-NCP interface thus

requires accurate determination of fibril diameter Df and embedded length le.

SEM was used to measure the diameter and length of the MCF before and after

pullout testing. The diameter of the MCF fibrils, measured at 5 equidistant points

along the MCF free length, were reasonably constant before and after pullout

testing with an average Df = 115.6 33 nm. The length of the MCF after pullout

consists of the MCF free length before pullout and the embedded length le. Thus,

subtraction of the MCF length before pullout from the MCF length after pullout

Page 197: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

196

provides le. SEM imaging of the MCF free lengths before and after pullout using

pixel analysis (ImageJ, NIH, USA) gave a range of le values from 510 to 880 nm.

The strain in the free length of the MCF during pullout testing is less than 1 %

when considering a maximum applied force of 0.3 N, the fibril diameter and the

MCF Young’s modulus defined in Chapter 4, indicating negligible fibril strain

contributions to the measured mechanical behaviour in this work.

The calculated MCF-NCP interfacial shear strength i is 0.65 0.15 MPa using

Equation 6.1 and the results in Figure 6.2. This shear strength for the nanoscale

interfaces found in bone is lower than values recorded from pullout of

engineering fibres from conventional fibre reinforced polymer composites [172,

264, 266, 269]. However, engineering composites are often optimized for

effective stress transfer between the reinforcing fibres and require relatively high

i values up to approximately 50 MPa [269, 270]. The MCF fibres in this work

therefore exhibit low interfacial shear strength, which is conducive for toughness

as cracks propagating through bone will be deflected at the weak interfaces

between the MCFs and surrounding NCP. The vast area available at these

nanoscale interfaces in bone may be a considerable energy absorbing process.

A stress-based analysis as described above can be indicative of the mechanical

behaviour of the MCF-NCP interface but provides little understanding of the

Page 198: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

197

fundamental nature of the interface. However, energy based criteria has been

used to understand molecular mechanisms present at interfaces in bone material

[152]. Specifically, the work done W to pullout the fibrils from the NCP is given

as:

𝑊 = ∫ 𝑙𝑒

𝐹𝑝

0

=1

2 𝑝𝑙𝑒 Equation 6.2

Previous studies suggest that macroscopic plasticity of bone is mainly due to the

plastic deformation occurring in extrafibrillar spaces as elastic deformation is

retained within the fibrils [5, 153, 234]. Thus, substantial work is required for the

fibril pullout process during breaking of bone, which contributes to the plastic

deformation of the extrafibrillar NCPs from MCF sliding and separation. At

molecular length scales, the NCP material consists of proteins including

osteopontin [262], proteoglycans [271, 272] and fetuin A [273] that are anchored

to the mineral sites on the MCFs and form an ionic network. The ionic bonds

existing in the thin layer of extrafibrillar NCPs bridge between negatively charged

protein molecules and divalent calcium ions, providing a molecular ‘glue’ to bind

the MCFs together. The work done to pullout the MCF from surrounding NCP

therefore causes failure of these sacrificial ion bonds in the NCP glue. However, a

number of works have shown how the interfacial region can reform [152, 167,

168, 274], indicating that the work of pullout may have to fail sacrificial ionic

bonds multiple times in the NCP before complete MCF separation from the bone

bulk.

Page 199: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

198

We therefore propose a model MCF-NCP system having partially contacting

protein molecules between MCFs as shown in Figure 6.3. Most of the mineral

content is found within the MCFs in antler bone, with little mineral found in the

NCP region. Mineral will therefore be present on the MCFs surface mainly at the

gap region of the MCF, as defined in the Hodge-Petruska scheme [82] and the

staggered arranged mineral arrangement [43], and will cover approximately 60%

of the MCF surface. The negatively charged protein molecules are therefore

bound to the mineral regions of the MCF as proposed by Salih et al. [275] as well

as Hartmann and Fratzl [168] as shown in Figure 6.3.

Figure 6.3 Schematic of the sacrificial ionic bonding system in NCPs region with

pink indicating MCFs, the blue region defining the extrafibrillar NCPs glue layer

and red denoting mineral crystals. Each short black dash line in NCPs layer

represents an activation volume. (A) NCPs are shown anchored to the mineral

MCFs

MCFs

NCPs

(A) (B)

(C

)

(D)

1 nm3

1 eV

Xn X0

Xn X0

Page 200: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

199

sites of the MCF. (B) The negatively charged NCP molecules are bonded together

by positively charged divalent calcium ions. (C) MCF pullout causes shear in the

NCP layer where ionic bonds are assumed to reversibly form process of MCFs

pullout. Thus, during plastic deformation of the interface glue region, ionic bonds

fail as negatively charged NCPs molecules are separated but reform when the

NCP molecules contact neighbouring NCP molecules during the pullout process.

The NCP bonding will reform and break continually with the fibril pullout

movement, resulting in a large number of bonds breaking relative to a

non-reforming pullout process as show in (D).

Under shear force applied to the MCF-NCP interface during pullout testing, the

negative charged protein molecules connected via calcium ions will slide to

separate and the ionic bonds will break and potentially reform the network in the

extrafibrillar region. To study this reforming behaviour of sacrificial bonds, two

extreme conditions can be assumed: i) failure and continued reforming of all

bonds from the NCP in contact with the separating MCF during the pullout as

indicated in Figure 6.3 (C) or ii) the complete failure of bonds at the MCF-NCP

interface during pullout. In i) the work done will be used to break a larger

number of bonds due to reforming events than condition ii). The activation

energy H is associated with the work done to break bonds over a unit volume and

has been previously recorded as approximately 1 eV for a 1 nm3 volume of bone

Page 201: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

200

material in which one or several sacrificial ionic bonds exist [152, 162, 167, 168].

Figure 6.3 (C) shows the condition where bonds existing at the MCF-NCP

interface reform during the interface failure. Each short dashed line represents a

unit activation volume defined as ‘unit’ in which a number of sacrificial bonds

exist. At position X0, bonds between the MCF and NCP will break but not reform

as the MCF is pulling out of the NCP. However, at position Xn, the unit will break

and reform by n times where n is the number of negatively charged molecules

existing along the long axis of the MCF. The whole extrafibrillar space can be

considered as a hollow cylinder with thickness of 1 nm as reported previously

[152, 168, 276] and the cylinder can be divided into n ‘rings’ along its long axis.

Each ring contains 𝐷√𝑉𝑎3⁄ units of activation volume. At the cross section

shown in Figure 6.3 (C), the units within the intrafibrillar space at one side of the

MCF will break and reform n(1+n)/2 times during fibril pullout. Therefore, over

the whole interface area, the total number of units breaking and reforming will

be:

𝑁100% = 𝐷

√𝑉𝑎3

×n(1 + n)

2≈

𝐷

√𝑉𝑎3

×n2

2 Equation 6.3

where the √𝑉𝑎3 represents the unit length of a side in a single cubic activation

volume which is equal to 1nm. 𝑉𝑎 = 1nm3, √𝑉𝑎23⁄ = 1nm2

Page 202: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

201

Conversely, the total number of units broken if no reforming occurs (N0%) as

shown in Figure 6.3 (D) can be stated as:

𝑁0% = 𝐷

√𝑉𝑎3

× 𝑛 Equation 6.4

In the Hodge-Petruska model of MCFs [82] with staggered arranged mineral

arrangement [43], approximately 60% of the MCF surface is covered with

mineral and attached with negatively charged protein chains. Therefore n can be

estimated using

𝑛 =0.6𝑙𝑒

√𝑉𝑎3

Equation 6.5

The activation energy H = W/N and provides boundary values of 0.011±0.0025

eV and 2.21±0.40 eV for bond reforming interfaces and complete interfacial

failure respectively. The calculated activation values lower than previous

literature value of 1 eV, and indicate complete bond reforming underestimates

the activation energy whereas no interfacial bond reforming overestimates this

activation energy. Thus, Equations 6.3 and 6.4 define boundary conditions,

indicating the interfacial bonding between the MCF and NCP will only reform

partially.

Page 203: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

202

To explore partial interface reforming, MCF pullout testing using the different

pullout velocities were compared to the work done W during pullout as shown in

Table 6.1. The calculated activation energies for the boundary conditions of

complete bond reforming (H100%) and complete bond failure (H0%) are also

shown in Table 6.1. The reforming of an individual sacrificial bond is a

probability event controlled by the molecular kinetics of NCPs protein chains.

However, over the whole interfacial space, the efficiency of the activation volume

units reforming has a statistical certainty in a single fibril pullout event due to a

large number of sacrificial bonds involved. Based on Equation 6.3, 6.4, and 6.5,

the total number of activation volume reforming can be calculated as:

𝑁 = 𝑁0% + 𝑁100% = 𝐷f(0.6𝛼𝑙𝑒)

2

2𝑉𝑎+ 𝐷f0.6𝑙𝑒(1 − 𝛼)

√𝑉𝑎23⁄

Equation 6.6

The reforming efficiency α is defined as the proportion of activation volume units

required to reform in order to provide a literature activation energy H=1 eV in

the fibril pullout and can be calculated as:

𝛼 =−𝑏 + √𝑏2 − 4𝑎𝑐

2𝑎 Equation 6.7

where: 𝑎 =0.36 𝐷f𝑙𝑒

2

2𝑉𝑎, 𝑏 = −

𝐷f0.6𝑙𝑒

√𝑉𝑎3

, 𝑐 = − 𝐷f0.6𝑙𝑒

√𝑉𝑎3

−𝑊

𝐻; 𝐻 = 1ev.

Page 204: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

203

By assuming all the sacrificial bonds within an activation volume unit share an

equal chance of breaking and reforming, the probability of a single bond

reforming P in the fibril pullout is the same as the activation volume reforming

efficiency α and listed in Table 6.1.

Table 6.1. Data showing the geometry, work done and calculated interfacial

behaviour in MCF pullout tests.

le

(nm)

Diameter

(nm)

Work

(×10-14J)

i

(MPa)

H100%

(eV)

H0%

(eV)

Reforming

efficiency

(%)

Pullout

velocity

(μm s-1)

700 168 7.89 0.61 0.011 2.22 7.88 2.55

510 102 2.96 0.71 0.012 1.89 7.94 3.21

740 110 7.00 0.74 0.013 2.85 9.36 2.26

880 121 6.03 0.41 0.0071 1.88 5.96 3.95

540 81 2.93 0.79 0.014 2.22 9.00 2.09

650

150

115.6

33

5.36

2.30

0.65

0.15

0.011

0.0026

2.21

0.40

8.02

1.32

2.55

0.76

Figure 6.4 shows the bond reforming efficiency plotted against the pullout

velocity. In particular, the bond reforming efficiency is expected to decrease as

the pullout velocity is increased due to the reduction in time available for ionic

bonds to reform at the MCF-NCP interface. Figure 6.4 supports this assumption

by showing a relatively rapid drop in bond reforming efficiency as the pullout

velocity increases fivefold from 0.5 to 2.5 m.s-1. The reforming efficiency of

Page 205: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

204

sacrificial ionic bonds in extrafibrillar space is highly dependent on the calcium

ions density within the NCP regions. The speed of calcium ions migration to new

bond reforming sites, which is decided by the permeability of NCP matrix, is

critical on the reforming process. More rapid pullout testing would therefore

suggest insufficient time for the calcium ions to transport sufficiently to bond

reforming sites, which leads to a lower reforming efficiency.

2 3 45

6

7

8

9

10

11

Bo

nd

refo

rmin

g p

rob

ab

ilit

y (

%)

Pull out speed (ms-1)

Figure 6.4 Plot of the sacrificial ionic bond reforming efficiency against individual

MCF pullout velocity speed for separation of MCFs from the surrounding NCP

region. The dashed line shows the trend of the reforming efficiency increasing

with decreasing pullout speed.

The calculation of activation energy for pullout of individual MCFs from a

Page 206: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

205

surrounding NCP binding ‘matrix’ based on AFM pullout experiments is

instructive in defining a partial capability for bonding reforming in bone at the

nanoscale. In addition, the molecular mechanism for bond reforming is expected

to be rate dependent and is supported by experimental pullout data. We also note

that the MCF pullout was achieved in less than 0.4 seconds. This pullout time

corresponds to an average testing strain rate of 4.27 ± 1.23 s-1, which is higher

than the strain rate (~1.59 s-1) used in testing to mimic the real loading condition

of antler bone in daily use and under sudden impact in combat [116]. Therefore,

under physiological loading conditions, antler bone will be expected to exhibit a

higher reforming efficiency than in this work.

6.4 Conclusions

In this chapter, individual MCFs were pulled out from surrounding NCP matrix

regions. Stress based analysis was used to show that the MCF-NCP interface is

weak, with energy based analysis suggesting an ability of the MCF-NCP interface

to reform during mechanical testing. The promotion of interfacial failure at the

nanoscale with ability to reform this interface is noted as being beyond current

synthetic nanocomposite design. Specifically, difficulties lie in producing

composites with dispersed, high volume fractions of nanofibres that provide

Page 207: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 6. Nanoscale Interfacial Behaviour in Bone

206

significant fracture surface after interfacial failure and the ability of interfacial

bonding to reform during the fracture process. The number of ionic bonds failing

at the fibril interface is therefore being maximized by reforming in order to

increase the overall work done to break the bone material.

Page 208: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

207

Chapter 7

Mineralized Collagen Fibril Contributions to

the Fracture Behaviour of Bone

7.1 Introduction

The nanoscale mechanical properties in bone, specifically the mechanical

behaviour of MCFs and their associated interfaces, have been elucidated in

previous Chapters but the relationship between nanoscale behaviour and whole

bone behaviour is unclear. This Chapter attempts to make predictions on the

contribution of the nanoscale to overall bone mechanics using simple fracture

considerations and composite models. The use of analytical composite models is

established in engineering materials and is therefore appropriate for the

MCF-type fibre reinforced structures of bone. The fracture toughness of bone

material is of particular importance and is related to the experimental fibril

Page 209: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

208

mechanics obtained in this work. In particular, the mechanical properties of

MCFs from antler bone have been evaluated in terms of their deformation and

failure behaviour in Chapter 4 and the mechanical properties of their interfaces

in Chapter 6. Therefore, the evaluation on the contribution of nanoscale fibril

fracture and pullout to the overall fracture resistance of antler bone is attempted

here.

7.1.1 Pullout Theory: Energy Balance Model

Observation of bone fracture surfaces in Chapters 4-6 indicates a propensity for

the MCFs to pullout from the bone matrix material. Indeed, the relatively weak

interfacial shear strength reported in the previous Chapter would support the

promotion of failure at the MCF-NCP interface. Descriptions of the contribution

of an individual interface on overall composite behaviour rely on consideration of

the fibre pulling out from the composite material. In fibre pullout testing, a

gradually increasing tensile force is applied to the fibre partially embedded

within the matrix until a critical force causes complete failure of the fibre-matrix

interface, resulting in the fibre being pulled out from the composite, with a

force-displacement curve typically recorded in this process. By applying a

suitable load transfer model, the interfacial shear strength, interfacial friction

coefficient and interface crack propagation behaviour can be measured. The

interfacial shear strength is given by using a simple force balance of:

Page 210: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

209

𝜏𝑖 = 𝑝

2 𝑙𝑒 Equation 7.1

where Fp is the maximum tensile force applied on fibre before pullout, r is the

fibre radius and le is embedded length of the fibre. This theory is straightforward

but ignores the effects of matrix elasticity, uneven distribution of tensile load in

the interfacial space, embedded length and Poisson’s ratio. More complete

descriptions of the fibre pullout process proposed by Chua and Piggott [277]

incorporate the following: i) lateral pressure applied on the fibre from matrix; ii)

interfacial friction coefficient; iii) interfacial fracture energy; iv) fibre embedded

length and free length.

The latest energy- based interface model proposed by Jiang and Penn [278] is

preferable in recent studies [264] as consideration of all factors listed above are

made based in the energy balance analysis. In this model, the experiment system

consists of three parts: i) fibre free length, ii) debonded region of fibre and iii)

bonded region between the fibre and matrix. The strain energy stored in these

regions as well as work of friction in fibre sliding out of the matrix are considered

in an energy balance analysis to derive the pullout load. Specifically, a crack

propagates at the interface between the fibre and matrix when the energy

released from the system provides sufficient energy to create new surfaces for

crack propagation in the bonded region of the composite and work to overcome

Page 211: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

210

friction in the debonded region is :

∂𝑈𝑓

∂𝑎+∂𝑈𝑑∂𝑎

+∂𝑈𝑏∂𝑎

= 2 𝐺f +∂𝑊𝑓

∂𝑎 Equation 7.2

where Uf, Ud and Ub, are the strain energies stored in the fibre free length region,

debonded region and bonded region respectively. Gf is the fibre interfacial

fracture energy and Wf is the work of friction at the debonded interface. The

crack length is describe by a so that no crack initiates when a=0.

By assuming negligible friction between the debonded fibre and matrix, the

interfacial fracture energy Gf can be calculated from the maximum pullout force

Fp :

𝑝 = (𝑛′

) 𝑙𝑒 √

𝐸f𝐴f2 𝐺f(2 + 𝛽)

Equation 7.3

where:

𝑛′ = √𝐸m

𝐸f(1 + 𝜈m) ln 𝑅 ⁄ Equation 7.4

and

𝛽 = √𝐸f𝐴f𝐸m𝐴m

Equation 7.5

where Ef and Em, are the Young’s modulus of the fibre and matrix respectively. Af

and Am are the cross section area of the fibre and matrix, νm is Poisson’s ratio of

Page 212: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

211

the matrix and R is the distance orthogonal from the fibre surface that is

perturbed when pulling the fibre from the matrix.

The analysis above can be applied to the pullout data using in situ AFM-SEM as

reported in Chapter 6. The nanoscale interfacial facture energy in fibrillar

systems of bone can be determined by using Equation 7.3, as has been carried

out in numerous fibrous composite systems [1]. One condition of applying

Equation 7.3 is that the interfacial friction between the fibre and matrix should

be negligible. When two contacting objectives are in relative movement, the

molecular interaction between the surfaces is generally described as a friction

force. The ‘friction’ in collagen fibril pullout from bone is essentially the break

and reforming of sacrificial ionic bonds existing within the extrafibrillar space. As

discussed in Chapter 6, the number of sacrificial bonds reforming in particular

pullout tests is less than 10 % of the total bonds under high pullout speed. The

friction raised from sacrificial bonds in the pullout is relatively low and is ignored

in this study. It is therefore expected to be reasonable to extend the concepts and

model above to MCFs pullout experiments.

When using Equation 7.3 for fibril pullout analysis, the matrix needs to be clearly

defined as n and α are highly dependent on the matrix properties. In this study,

the matrix may be defined as the antler bone tissue surrounding the tested fibril.

As shown in Figure 7.1 (A), when the fibril is pulled out from the bone matrix, the

Page 213: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

212

neighbouring MCFs are also affected by the shear stress transferred via the

extrafibrillar NCPs layer. Alternatively, the matrix could be defined as the NCPs

that exist between fibrils. Only the fibril with load applied directly with

surrounding NCPs layer is affected by shear stress as shown in Figure 7.1 (B).

However, this assumption indicates that the NCPs layer does not transfer load

between MCFs, which is contradictory to observations (fibril interfacial shear

strength recorded as 0.6 MPa) in Chapter 6.

Figure 7.1 Schematic diagrams showing stress distribution in the pullout fibril

and surrounding bone tissue. MCFs are represented by orange cylinders with

blue extrafibrillar NCP regions existing between the MCFs. (A) Adjacent MCFs are

deformed by the shear load transferred from NCPs whereas (B) Surrounding

bone tissue is not affected if assume NCP do not transfer load which is

contradictory to the phenomenon observed in Chapter 6.

Fp

R

Fp

R

(A) (B)

Page 214: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

213

7.2 Results

If we assume that a relatively large amount of matrix as bone tissue surrounding

is deformed during a pullout event, α in Equation 7.3 can be neglected as the

matrix area Am is much larger than the fibril area Af. Equation 7.3 can be

therefore rearranged as:

P𝑙𝑒= (

𝐾′𝑛′

)√𝐺f Equation 7.6

Where K’ is a constant defined by the geometry and Young’s modulus of MCFs:

𝐾′ = √𝐸f𝐴f Equation 7.7

Equation 7.6 can be solved by first plotting the variation in the maximum pullout

force FP against the embedded fibril length le using the data collected in Chapter 6.

The maximum pullout force FP plotted against the fibril embedded length is

shown in Figure 7.2 with a linear regression used to give the ratio 𝛿 𝛿𝑙𝑒⁄ .

Page 215: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

214

0 200 400 600 800 1000

0.00

0.05

0.10

0.15

0.20

0.25

0.30

Embedded length (nm)

Pu

llo

ut

forc

e (N

)

Figure 7.2 Maximum pullout force required for pulling out of individual MCFs

from bone matrix plotted against fibril embedded length. Red dash line shows

the linear regression.

The fibril interfacial fracture energy Gf can be calculated using Equation 7.6 from

the linear regression in Figure 7.2. In addition, average MCF radii of 57.8 16.5

nm were tested in pullout, the Poisson’s ratio of antler bone matrix νm is

recorded as 0.3-0.32 [103, 116, 170, 171, 279] and the Young’s modulus of wet

compact bone tissue from red deer antler is recorded to be ranged between 5.1

to 7.8 GPa [103, 116, 170, 171] (indeed, our samples of antler bone are the same

source as Currey’s 2009 work [171] with Young’s modulus of 7.2 GPa). Finally,

the constant n contains the stress transfer parameter R/r which cannot be

determined in this fibril pullout test. Therefore, the calculated fibril interfacial

Page 216: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

215

fracture energy Gf is plotted against R/r as shown in Figure 7.3. Typically, R/r

values vary from 2 - 3 for weak interfacial adhesion up to 9 for strong adhesion

[280, 281]. Hence, the calculated Gf from Equation 7.6 is plotted against stress

transfer parameter R/r as shown in Figure 7.3 and ranged from 2.17 to 6.87 J.m-2.

1 2 3 4 5 6 7 8 9 10

1

2

3

4

5

6

7

8

9

R/r

Gf (

Jm

-2)

Figure 7.3 Calculated interfacial fracture energy from pullout of individual MCFs

from bone matrix is plotted with different stress transfer parameters R/r. Error

bars indicate the standard deviation of the interfacial fracture energy values

calculated from five individual fibril pullout tests.

The calculation of interfacial fracture energy shown in Figure 7.3 requires

interpretation of R/r values. The results of fibril pullout experiments in Chapter 6

indicate relatively low interfacial shear strength values, at least relative to

Page 217: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

216

engineering composites, which suggest that small R/r values operate in bone.

However, R/r values form a complex relationship with the elastic properties of

composite constituents as shown in Equation 7.4. Additional complications in

determining Gf from R/r values arise from the elastic response assumed in the

analysis and the relatively small amount of matrix existing between the MCFs

that can be deformed.

Gf describes the fracture energy release in unit area of MCF-NCP interface

fractured. Figure 7.4 show a crack propagating perpendicular to the osteon

orientation (which is also the principle orientation of MCFs). Such crack

propagation will cause failure of individual MCFs, followed by pullout to leave

exposure of fibril ends and pullout voids at the fracture surface.

Figure 7.4 Schematic showing the propagation of a crack perpendicular to the

osteon orientation in bone leading to fibril fracture and pullout. MCFs are

Page 218: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

217

represented by orange cylinders with blue extrafibrillar NCP regions existing

between the MCFs.

The energy required for crack propagating at the fracture surface combines the

work done to break MCFs and the fibre interfacial fracture energy Gf. Considering

MCFs are closely packed in bone and share interfaces with neighbouring fibrils,

the total fracture energy required to fail the interfaces between MCFs can be

calculated as:

∆𝑈i =𝑓𝑙𝑝𝐺f

Equation 7.8

where f is the MCF volume fraction, taken as 82 % [12], lp is pullout length of the

fibrils exposed at the fracture surface, Gf has a value ranging from 2.17 to 6.87

J.m-2 as calculated from energy based analysis in Figure 7.3. The total fracture of

bone therefore consists of the total interfacial fracture energy given in Equation

7.14 but also a contribution from the failure of MCFs. The total energy required to

fracture MFCs during propagation of a crack in bone can be calculated form the

area under stress-strain curves recorded in Chapter 4, thus:

∆𝑈f = 𝑓σε𝑙𝑝 Equation 7.9

where is the maximum MCF failure stress and is the maximum MCF failure

Page 219: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

218

strain, taken from average values in Chapter 4. Hence, the total fracture energy Gc

released when considering MCF pullout and fibril failure is the summation of

Equations 7.8 and 7.9 to give:

𝐺c =𝑓𝑙𝑝𝐺f

+ 𝑓σε𝑙𝑝 Equation 7.10

The fibril pullout length lp can be measured by pixel analysis on scanning

electron microscopy images. However, fibrils might be oriented in an angle to the

SEM image plane and direct measurement on SEM image therefore might not

give real length of fibrils. To measure the true fibril length and eliminate the

error from various fibril orientations, SEM images on fibrils were taken from

different angles as shown in Figure 7.5. The real fibril length is given by the

trigonometry calculation as:

𝑙𝑝 = √𝑙302 − cos2 30𝑙0

2

1 − cos2 30 Equation 7.11

where l0 and l30 are the fibril pullout lengths measured when observing the

fracture surface at 0 and 30 degrees respectively in the SEM as shown in Figure

7.5.

Page 220: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

219

Figure 7.5 Scanning electron micrographs of fracture surfaces in antler bone from

different viewing angles. (A) Image taken with sample stage at 0 degree tilt. (B)

Image taken when sample stage was tilted by 30 degrees.

106 fibrils were randomly chosen with their lengths measured over the whole

fracture surface. The distribution of fibril pullout length was plotted in Figure 7.6.

An average fibril pullout length of 2.8 1.7 μm with an average fibril radius of

51.3 6.3 nm was recorded. The total fracture energy directly associated from

fibril pullout and failure mechanisms at a bone fracture surface calculated from

Equation 7.10 gives a range of values from 139.8 to 334.7 J.m-2, in which the fibril

fracture energy contributes less than 2 J.m-2.

(A) (B)

Page 221: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

220

0 2 4 6 8 100

5

10

15

20

Length Distribution

Average=2.8m

Median=2.3m

SD=1.7

Nu

mb

ers

Fibril pullout length (m)

Figure 7.6 Distribution of fibril pullout length at antler bone fracture surfaces.

7.3 Discussion

The fracture energy of nanoscale interfaces between MCFs in bone calculated

using energy based analysis is much lower than values for synthetic engineering

composites reinforced with fibres chemically modified to promote strong

interfacial adhesion, suggesting that the molecular interaction between MCFs is

weaker than covalent bonds in the chemically modified interface in artificial

composites. For example, a previous study [282] gave a range of interfacial

fracture energy values for glass fibres pulled from a variety of polymers such as

vinyl ester (13–34 J.m-2), polyamide 6 (24–93 J.m-2) and polyamide 6.6 (52–61

J.m-2) [264]. On the other hand, the fracture energy of the MCF interface

Page 222: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

221

calculated here is comparable to ionic bonded interfaces in ceramics [283-285],

which confirmed previous simulation studies indicating how ionic bonds are

dominant at fibrillar interfaces in bone [152, 167, 168, 274].

The relatively low interfacial fracture energy calculated in this Chapter

importantly needs to be placed in context with the fracture of whole bone. As

discussed in Section 2.4.2, the multiple length-scale toughening acting on cortical

bone leads to a unique rising resistance-curve (R-curve) behaviour for bone

fracture [6, 136, 137, 140]. The fracture toughness of bone material increases

with crack extension in bone. Vashishth et al. [286, 287] have reported rising

R-curve behaviour in antlers of red deer and demonstrated that the superior

toughness of antler bone is due to its enhanced ability to form microcracks

during deformation and fracture. The fracture work recorded previously in

R-curve measurements on antler bone using in situ synchrotron X-ray method

has shown low initiation fracture energy of 5.71 to 17.93 J.m-2 with crack length

of 0.02 mm that increases to almost 1 kJ.m-2 when crack length reaches 1 mm in

all loading directions [288]. Other studies have shown a much higher fracture

energy from 30 kJ.m-2 [170, 171] to 60 kJ.m-2 [15, 288] recorded when cracks

propagate in the transverse direction relative to the long axis of bone.

The difference the fracture energy of bone, which is of the order of 10 kJ m-2, and

Page 223: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

222

the initiation of cracking is approximately 3 orders of magnitude. This difference

is significant as initiation of cracking does not involve the large amount of plastic

deformation and microcracking occurring in bone beyond crack initiation. Indeed,

the energy based analyses used above ignore frictional behaviour associated with

pullout, which could contribute significantly to the overall fracture energy of

bone. Therefore, our calculated fracture energy results should be compared to

crack initiation behaviour only. Considering the fibril interfacial fracture energy

Gf calculated varies from 2.17 to 6.87 J.m-2, the nanoscale fibril interfaces

contribute roughly 50 % of the initiation fracture energy in antler bone. It is

therefore reasonable to assume that when cracking initiates in bone, the major

mechanism is the separation of fibrils around a crack. The interfacial fracture is

dominant in fracture initiation in antler bone at this stage.

When a crack extends to up to 1 mm, the fracture energy is recorded to be

approximately 1 kJ.m-2 in all loading directions [288]. The total fracture energy

directly associated from fibril pullout and failure mechanisms at a bone fracture

surface calculated from Equation 7.10 gives a range of values from 139.8 to 334.7

J.m-2 in which fibril interfaces contribute more than 90 %. Our result indicates

the nanoscale contributions of MCFs, especially the fibril interfaces, to the

fracture process at this stage is significant with up to 33 % of the total energy

required to fail bone by assuming only simple flat surfaces are produced in

Page 224: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

223

fracture as shown in Figure 7.4 (B). The real contributions of MCFs could be

higher as i) crack propagates in deflected and twisted paths generate larger

fracture areas as demonstrated in Figure 7.7 (A), ii) there might be more cracks

existing away from fracture sites which leads to an underestimation on number

of interfacial failure events used for calculated fracture energies.

Crack propagation in bone material beyond crack lengths of 1 mm cause large

variability in the fracture energies of bone. For cracks propagating transverse to

the osteon direction, the fracture energy of antler bone can reach values of up to

31 kJ.m-2 [116, 171] to 60 kJ.m-2 [15, 288]. This fracture energy is considerably

higher than for longitudinal splitting of bone as considerable delocalized failure

of bone occurs. Thus the total number of MCF fibril events contributing to the

total fracture energy of bone in transversal direction cannot be simply made by

summation of the individual events at the fracture as described using Equation

7.10. Instead, the total number of events within the fracture volume, which

extends considerably beyond the apparent bone fracture surfaces observed, must

be considered but is difficult to know accurately. Similarly, splitting antler bone

along it longitudinal direction only requires a fracture energy of 6 to 10 kJ.m-2 [15,

116, 171, 288] due to failure being restricted more towards the apparent fracture

surfaces of bone as opposed to the larger failure zones occurring in transverse

cracking. However, the total number of failure events in both longitudinal

Page 225: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

224

splitting and transverse cracking are significantly larger that than assumed at the

propagating cracks as assumed in Equation 7.10. Figure 7.7 highlights the

differences in the damage zone for longitudinal and transverse crack growth. In

particular longitudinal crack propagation in Figure 7.7 (A) shows a smaller

damage zone than transverse cracking in (Figure 7.7 (B), which was confirmed

by X-ray tomography studies [288]. Transverse cracks therefore produce a larger

failure region than actual fracture surface, increases the fracture energy

considerably when compared to longitudinal cracking which greatly enhances

the fracture resistance of antler bone. Therefore, both transverse and

longitudinal failure in bone represents the failure mechanism away from the

failure described using Equation 7.10. Smaller cracks below 1 mm must therefore

reflect Equation 7.10 most closely.

In addition, weak interfaces between fibrils enable the occurring of nanocracks

away from fracture site due to the heterogeneous mechanical behaviour of fibrils

as recorded in Chapter 4. In the stress affected region ahead of a crack tip, the

strain mismatch between fibrils with different Young’s moduli leads to the failure

of interfaces and separation of fibrils. The nanocracks formed away from a crack

tip will accumulate, resulting in an increased stress away from the crack tip and

eventually formation of microcracks as shown in Figure 7.7 (B). The appearance

of microcracking absorbs additional energy and promotes other toughening

Page 226: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

225

mechanisms as discussed in Section 2.4.2.

Figure 7.7 (A) and (B) Scanning electron micrographs of crack growth path at a

fracture tip in antler bone [288]. Transverse crack propagates against osteon

orientation in (A) with red arrows show the crack ‘twisting’ during crack growth.

(B) Cross section image of osteons showing how ‘micro cracking’ occurs away

from a crack tip. (C) and (D) Synchrotron X-ray computed micro-tomography

images of cracks occurring in both the longitudinal and transverse orientations of

antler bone [288] with cracks shown in purple. Central canals in osteons (hollow

volume in osteons) and vascular channels are represented in brown. Both

longitudinal and transverse directions exhibit significant cracking. However,

(A) (B)

(C) (D)

Page 227: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

226

cracking in the transverse orientation (D) shows more failure volume due to

crack deflection and microcracking generated in the neighbouring space near the

principal crack tip [288].

7.4 Conclusions

In this chapter, energy based analyses using results from pullout testing of MCFs

was carried out. A fibril interfacial fracture energy ranging from 2.87 to 6.87 J.m-2

was calculated and was found to be comparable to ionic bonded interface values,

which is coincident with previous simulation result that fibrillar interface is

dominated by ion bonds. Around 50% of the work done to initiate the crack was

indicated to be due to the failure of the nanoscale fibrillar interfaces which makes

it a potent mechanism in crack initiation. The total energy required to both

fracture MCFs, using data from Chapter 4, and the calculated interfacial fracture

energies were used to develop total fracture energy for bone based on the MCF

contribution. Around 33 % of the energy required to crack bone with relatively

small crack lengths was needed to fail MCFs, although the fibril strength is a

minor contribution, and their associated interfaces. This total contribution from

MCFs assumed the failure was restricted to the surface, which is considerably

different to larger transverse and longitudinal splitting fracture behaviour where

Page 228: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 7. MCF Contributions to Fracture Behaviour of Bone

227

crack lengths are above 1 mm. Both longitudinal and transversal fracture of bone

have failure regions that are significantly larger than the apparent fracture

surface of bone although longitudinal cracking creates smaller fracture regions

when comparing to transverse fracture.

Page 229: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

228

Chapter 8

Conclusions and Future Work

8.1 Summary of the Thesis

In the first chapter, the need for research in the field of MCF mechanics and its

influence on overall mechanical behaviour of bone tissue was established.

Objectives of the project along with thesis outline were defined. Chapter 2

reviewed existing literature for bone material with emphasis on: i) the bone

formation process (ossification) and biomineralization, ii) compositional

information, iii) structural hierarchy, iv) bone mechanics at different length

scales including current studies on collagen fibrils and MCFs as well as multiple

length-scale toughening mechanisms. This review justified the importance of the

constituent mechanical behaviour on the overall mechanical properties of bone.

Page 230: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

229

Chapter 3 reported on the in situ nanomechanical testing methodology

combining AFM with SEM for investigating nanofibrous samples including MCFs

from bone in detail, including sample preparation processes and the

nanomechanical testing setup as well as data collection and analysis. Samples

were prepared without applying chemical treatments to retain the compositional

and mechanical characteristics of MCFs, validated by the similarity of MCF

mechanics to more incomplete fibril strain data in the literature and the apparent

initial linear stress-strain response of MCFs at the bone fracture surfaces. The

suitability of the AFM-SEM setup for mechanical testing of nanoscale fibrous

samples was evaluated by performing tensile tests on synthetic electrospun PVA

and nylon-6 polymer nanofibres. Stress-strain curves recorded on PVA and

nylon-6 indicated that the AFM-SEM testing method is suitable for testing

nanofibres. In addition, mechanical properties of PVA nanofibres measured after

exposing to vacuum in the SEM chamber for a relatively short time are

comparable to hydrated bulk PVA film, indicating that water is retained in the

sample within the testing time frame used.

The AFM-SEM technique in Chapter 3 was applied to individual MCFs in Chapter

4 to show a two-stage stress-strain behaviour of individual MCFs from antler

bone during tensile testing to failure. Fibrils exhibited a uniform linear

stress-strain response in a first stage followed by inhomogeneous deformation

Page 231: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

230

above MCF strains of 2-3.7 %. The inhomogeneous mechanical properties of

MCFs were explained by the compositional heterogeneity at the nanoscale

recorded using BSE microscopy. A phenomenological model was proposed to

explain the origin of high fracture resistance of antler bone at the fibrillar level

due to inhomogeneous fibril deformation.

As mineralization level in collagen fibrils was found to be crucial in defining

MCF’s mechanical behaviour in Chapter 4, MCFs isolated from different ages of

mouse bone with different degrees of mineralization were tensile tested in

Chapter 5 and compared to bulk mouse bone samples. Little variation in the

Young’s modulus of MCFs from different ages of bone was observed whereas bulk

bone samples showed a significantly stronger age dependent elastic response.

These results suggest that MCFs from the ages of bone tested have a similar

degree of mineralization despite the bulk bone samples having significantly

larger Young’s modulus variations. A fibrillar biomineralization process was

proposed where mineral ions first migrate to the free spaces within fibrils and

replace the water in the gap regions gradually. After all gap regions in fibrils are

filled with mineral, apatite crystals continue to grow into extrafibrillar spaces

beyond the fibril surface, leading to small increases in the overall bone

mineralization as the extrafibrillar space is limited for mineral to deposit within.

The appearance of extrafibrillar mineral does not change the Young’s modulus of

Page 232: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

231

the fibril itself but promotes stress transfer between fibrils and results in an

improved stiffness of bulk bone. The importance of nanoscale interfaces between

the fibrils in bone is therefore noted and studied further in next two chapters.

In Chapters 6 and 7, the interface between MCFs and surrounding NCPs in antler

compact bone was mechanically evaluated by pulling individual MCFs out of the

antler bone matrix. The shear strength of the MCF-NCP interface was calculated

and the sacrificial bonds reforming activities at the interface examined with the

reforming efficiency estimated in Chapter 6. The bond reforming activity was

found to be dependent on the strain rate of pullout testing, controlled by NCP

molecular kinetics. Thus, high pullout strain rates do not provide sufficient time

for sacrificial bonds to reform whereas slower pullout rates provide a higher NCP

reforming efficiency at the interface region. Critically, the interfacial shear

strength was found to be relatively small, at least compared to interfaces in

engineering composites, which indicated a propensity for bone to fail at the

MCF-NCP interface. The potential for interfacial failure at the nanoscale was

developed further in Chapter 7 in order to examine the contributions of failure at

nanoscale interfaces on whole bone mechanics. The MCF interfacial fracture

energy Gf was defined as a key parameter and evaluated using energy based

analysis approaches, showing values similar to the ionic bonded interfaces

coincident with previous simulation studies. The calculated MCF interfacial

Page 233: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

232

fracture energy was evaluated using analytical composite models and found to

contribute significantly more than MCF failure to toughness, with an overall 50 %

contribution of interfacial failure to the total fracture energy when cracks initiate

in antler bone. It is therefore reasonable to suggest that fibril mechanics at the

nanoscale have a critical role in defining overall mechanical properties, especially

the fracture behaviour of bulk bone material. The relative importance of MCF

mechanics is in contrast to a number of previous models that indicated the origin

of bone’s extraordinary toughness is built upon the multiple toughening

mechanisms due to microstructural arrangements in bone. These toughening

mechanics at micro and macroscopic scales must therefore be at least partially

based on MCF contributions.

8.2 Future Work

The work presented in the thesis is pioneering in direct mechanical study on

MCFs from bone tissues but present a number of future opportunities.

Specifically, mechanical testing usually requires large number of samples tested

but a larger statistical data set is difficult to achieve in this study due to the

difficulty in nanoscale manipulation and performing nanofibrous tensile testing

experiments. Although every experiment in this thesis was performed on at least

Page 234: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

233

5 to 6 samples, which meets minimum statistics requirement, larger data sets

would provide information on the material variation in bone.

The reforming efficiency of sacrificial bonds in extrafibrillar spaces during fibril

pullout was found to be dependent on strain rates. Further fibril pullout tests

using strain controlled testing devices to enable the pullout of fibrils at lower

physiological loading speeds would be obviously desirable and is not currently

available in the AFM setup used. In addition, studies on NCP molecular kinetics of

straining and relaxation is needed for a better explanation on reforming

efficiency. Molecular simulation studies linked to the experimental pullout data

could be instrumental in revealing molecular kinetics behaviour of NCPs.

Finally, the fibril interfacial fracture was found to contribute significantly to the

fracture of bone tissue when cracks initiate. However, the contribution of fibril

interfacial failure to larger cracking events is much more difficult to evaluate as

the number of interfacial failure events is difficult to evaluate as the failure zone

extends significantly beyond the crack tip. For example, transverse cracking of

bone samples show a significant diffuse cracking zone beyond the crack tip.

Future work investigating the total amount of surface area created within the

total failure zone in front of a propagating crack would be more instructive in

determining the contribution of nanoscale interfacial failure on overall bone

Page 235: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

234

toughness. One possible approach to examine this problem is to perform in situ

fracture testing on bone and observe cracks with high resolution X-ray

tomography to obtain a better approximation of the accumulation of nanoscale

crack surfaces during propagation of millimetre crack lengths in bone material.

8.3 Major Findings of the Thesis

All the objectives proposed in first chapter were successfully accomplished and

the main findings in this thesis are listed below:

The novel in situ AFM-SEM nanomechanical testing method established in

this thesis was proven to be suitable for nanofibrous samples.

Dehydration of specimens by exposing to vacuum in the SEM chamber was

found to be minimum within the time frame used in all the in situ

experiments (usually 20 mins) including testing on MCFs from bone and PVA

nanofibres indicating the applicability of this technique on hydrated samples.

MCFs from antler bone have a two-stage stress-strain behaviour. A linear

response with Young’s modulus of 2.4 0.4 GPa was recorded before 2 - 3.7 %

Page 236: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

235

strain of fibrils followed by an inhomogeneous deformation. An ultimate

strength of 118.7 38.7 MPa was recorded with failure strain of 6.0 0.6 %.

Fibrillar inhomogeneous mechanical behaviour is associated with its

compositional heterogeneity. Intrafibrillar mineral crystals perform as

crosslink points between tropocollagen molecules to enhance stiffness.

Different mechanical properties of fibrils are expected to lead to nanoscale

heterogeneous deformation in bone which promotes energy dissipation.

The appearance of extrafibrillar mineral during mineralization process does

not alter the mechanical properties of individual MCFs themselves but

provides stronger connectivity between fibrils, which enhances the stiffness

of bone tissue.

Fibrillar interfacial shear strength was calculated to be 0.6 0.15 MPa by

pulling out MCFs from antler bone, indicating relatively weak bonding

between fibrils. The reforming of sacrificial ionic bonds between fibrils was

found to be dependent on the pullout speed.

A fibrillar interfacial fracture energy Gf ranged from 2.17 to 6.87 J.m-2 was

calculated and found to be similar to values recorded on ionic bond

Page 237: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

Chapter 8. Conclusions and Future Work

236

dominated interfaces, which confirmed simulation results from the literature

describing MCFs linked together by ionic bonds.

The direct contribution from MCFs on antler bone fracture energy is about

50 % when crack initiates but potentially reduces when cracks extend further.

Page 238: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

237

References

1. Weiner, S. and H.D. Wagner, The material bone: Structure mechanical function

relations. Annual Review of Materials Science, 1998. 28: p. 271-298.

2. Weiner, S., W. Traub, and H.D. Wagner, Lamellar bone: Structure-function relations.

Journal of Structural Biology, 1999. 126(3): p. 241-255.

3. Wainwright, S.A., Mechanical design in organisms. 1976, New York: Wiley. xii, 423 p.

4. Fratzl, P. and R. Weinkamer, Nature's hierarchical materials. Progress in Materials

Science, 2007. 52(8): p. 1263-1334.

5. Fratzl, P., H.S. Gupta, E.P. Paschalis, and P. Roschger, Structure and mechanical quality

of the collagen-mineral nano-composite in bone. Journal of Materials Chemistry, 2004.

14(14): p. 2115-2123.

6. Vashishth, D., J.C. Behiri, and W. Bonfield, Crack growth resistance in cortical bone:

Concept of microcrack toughening. Journal of Biomechanics, 1997. 30(8): p. 763-769.

7. Rho, J.Y., L. Kuhn-Spearing, and P. Zioupos, Mechanical properties and the hierarchical

structure of bone. Medical Engineering & Physics, 1998. 20(2): p. 92-102.

8. Reilly, G.C., Stress-induced damage in bone: Microcracking and other toughening

mechanisms. Journal of Bone and Mineral Research, 1996. 11(10): p. P28-P28.

9. Nalla, R.K., J.J. Kruzic, and R.O. Ritchie, On the origin of the toughness of mineralized

tissue: microcracking or crack bridging? Bone, 2004. 34(5): p. 790-798.

10. Wright, T.M., B.A. Conway, J.S. Walton, and W.C. Hayes, Tensile properties of bone

effects of strain rate micro structure and density. IRCS (International Research

Communications System) Medical Science Library Compendium, 1976. 4(3): p. 109.

Page 239: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

238

11. Bonfield, W. and M.D. Grynpas, Anisotropy of youngs modulus of bone. Nature, 1977.

270(5636): p. 453-454.

12. Currey, J.D., Bones : structure and mechanics. 2002, Princeton, NJ: Princeton

University Press. xii, 436 p.

13. Ritchie, R.O., J.H. Kinney, J.J. Kruzic, and R.K. Nalla, A fracture mechanics and

mechanistic approach to the failure of cortical bone. Fatigue & Fracture of Engineering

Materials & Structures, 2005. 28(4): p. 345-371.

14. Zimmermann, E.A., M.E. Launey, H.D. Barth, and R.O. Ritchie, Mixed-mode fracture of

human cortical bone. Biomaterials, 2009. 30(29): p. 5877-5884.

15. Launey, M.E., M.J. Buehler, and R.O. Ritchie, On the Mechanistic Origins of Toughness in

Bone, in Annual Review of Materials Research, Vol 40, D.R. Clarke, M. Ruhle, and F. Zok,

Editors. 2010, Annual Reviews: Palo Alto. p. 25-53.

16. Melvin, J.W., Fracture-mechanics of bone. Journal of Biomechanical

Engineering-Transactions of the Asme, 1993. 115(4): p. 549-554.

17. Hang, F., D. Lu, and A.H. Barber, Combined AFM-SEM for mechanical testing of fibrous

biological materials, in Structure-Property Relationships in Biomineralized and

Biomimetic Composites, D. Kisailus, et al., Editors. 2009, Materials Research Society:

Warrendale. p. 135-140.

18. Beniash, E., Biominerals-hierarchical nanocomposites: the example of bone. Wiley

Interdisciplinary Reviews-Nanomedicine and Nanobiotechnology, 2011. 3(1): p.

47-69.

19. Hassenkam, T., G.E. Fantner, J.A. Cutroni, J.C. Weaver, D.E. Morse, and P.K. Hansma,

High-resolution AFM imaging of intact and fractured trabecular bone. Bone, 2004.

35(1): p. 4-10.

20. van der Rijt, J.A.J., K.O. van der Werf, M.L. Bennink, P.J. Dijkstra, and J. Feijen,

Micromechanical testing of individual collagen fibrils. Macromolecular Bioscience,

2006. 6(9): p. 697-702.

21. Hassenkam, T., R.B. Svensson, and M. Zalkovskij, Nano-science revelations in bone

Page 240: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

239

research. Current Nanoscience, 2007. 3(4): p. 345-351.

22. Lee, C.K., Y.M. Wang, L.S. Huang, and S.M. Lin, Atomic force microscopy: Determination

of unbinding force, off rate and energy barrier for protein-ligand interaction. Micron,

2007. 38(5): p. 446-461.

23. Sader, J.E., I. Larson, P. Mulvaney, and L.R. White, Mthod for the calibration of

atomic-force microscope cantilevers. Review of Scientific Instruments, 1995. 66(7): p.

3789-3798.

24. Ludwig, M., M. Rief, L. Schmidt, H. Li, F. Oesterhelt, M. Gautel, and H.E. Gaub, AFM, a

tool for single-molecule experiments. Applied Physics a-Materials Science &

Processing, 1999. 68(2): p. 173-176.

25. Turner, C.H., J. Rho, Y. Takano, T.Y. Tsui, and G.M. Pharr, The elastic properties of

trabecular and cortical bone tissues are similar: results from two microscopic

measurement techniques. Journal of Biomechanics, 1999. 32(4): p. 437-441.

26. Zysset, P.K., X.E. Guo, C.E. Hoffler, K.E. Moore, and S.A. Goldstein, Elastic modulus and

hardness of cortical and trabecular bone lamellae measured by nanoindentation in the

human femur. Journal of Biomechanics, 1999. 32(10): p. 1005-1012.

27. van der Linden, J.C., D.H. Birkenhager-Frenkel, J.A. Verhaar, and H. Weinans,

Trabecular bone's mechanical properties are affected by its non-uniform mineral

distribution. J Biomech, 2001. 34(12): p. 1573-80.

28. Ferguson, V.L., A.J. Bushby, and A. Boyde, Nanomechanical properties and mineral

concentration in articular calcified cartilage and subchondral bone. Journal of

Anatomy, 2003. 203(2): p. 191-202.

29. Ni, G.X., Y.S. Choy, W.W. Lu, A.H.W. Ngan, K.Y. Chiu, Z.Y. Li, B. Tang, and K.D.K. Luk,

Nano-mechanics of bone and bioactive bone cement interfaces in a load-bearing model.

Biomaterials, 2006. 27(9): p. 1963-1970.

30. Tan, E.P.S. and C.T. Lim, Mechanical characterization of nanofibers - A review.

Composites Science and Technology, 2006. 66(9): p. 1102-1111.

31. Tai, K., M. Dao, S. Suresh, A. Palazoglu, and C. Ortiz, Nanoscale heterogeneity promotes

Page 241: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

240

energy dissipation in bone. Nature Materials, 2007. 6(6): p. 454-462.

32. Wenger, M.P.E., L. Bozec, M.A. Horton, and P. Mesquida, Mechanical properties of

collagen fibrils. Biophysical Journal, 2007. 93(4): p. 1255-1263.

33. Roschger, P., E.P. Paschalis, P. Fratzl, and K. Klaushofer, Bone mineralization density

distribution in health and disease. Bone, 2008. 42(3): p. 456-466.

34. Tai, K., G. Pelled, D. Sheyn, A. Bershteyn, L. Han, I. Kallai, Y. Zilberman, C. Ortiz, and D.

Gazit, Nanobiomechanics of Repair Bone Regenerated by Genetically Modified

Mesenchymal Stem Cells. Tissue Engineering Part A, 2008. 14(10): p. 1709-1720.

35. Brandi, M.L., Micro architecture, the key to bone quality. Rheumatology, 2009. 48: p.

IV3-IV8.

36. Currey, J., Measurement of the Mechanical Properties of Bone: A Recent History. Clinical

Orthopaedics and Related Research, 2009. 467(8): p. 1948-1954.

37. Feng, L.A. and I. Jasiuk, Multi-scale characterization of swine femoral cortical bone.

Journal of Biomechanics, 2011. 44(2): p. 313-320.

38. Pathak, S., J.G. Swadener, S.R. Kalidindi, H.W. Courtland, K.J. Jepsen, and H.M. Goldman,

Measuring the dynamic mechanical response of hydrated mouse bone by

nanoindentation. Journal of the Mechanical Behavior of Biomedical Materials, 2011.

4(1): p. 34-43.

39. Graham, J.S., A.N. Vomund, C.L. Phillips, and M. Grandbois, Structural changes in

human type I collagen fibrils investigated by force spectroscopy. Experimental Cell

Research, 2004. 299(2): p. 335-342.

40. Shen, Z.L., M.R. Dodge, H. Kahn, R. Ballarini, and S.J. Eppell, Stress-strain experiments

on individual collagen fibrils. Biophysical Journal, 2008. 95(8): p. 3956-3963.

41. Parkinson, J., A. Brass, G. Canova, and Y. Brechet, The mechanical properties of

simulated collagen fibrils. Journal of Biomechanics, 1997. 30(6): p. 549-554.

42. Fratzl, P., K. Misof, I. Zizak, G. Rapp, H. Amenitsch, and S. Bernstorff, Fibrillar structure

and mechanical properties of collagen. Journal of Structural Biology, 1998. 122(1-2):

p. 119-122.

Page 242: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

241

43. Jager, I. and P. Fratzl, Mineralized collagen fibrils: A mechanical model with a staggered

arrangement of mineral particles. Biophysical Journal, 2000. 79(4): p. 1737-1746.

44. Eppell, S.J., W.D. Tong, J.L. Katz, L. Kuhn, and M.J. Glimcher, Shape and size of isolated

bone mineralites measured using atomic force microscopy. Journal of Orthopaedic

Research, 2001. 19(6): p. 1027-1034.

45. Lees, S., Mineralization of type I collagen. Biophysical Journal, 2003. 85(1): p.

204-207.

46. Orgel, J., T.C. Irving, A. Miller, and T.J. Wess, Microfibrillar structure of type I collagen in

situ. Proceedings of the National Academy of Sciences of the United States of America,

2006. 103(24): p. 9001-9005.

47. Gautieri, A., S. Vesentini, A. Redaelli, and M.J. Buehler, Intermolecular slip mechanism

in tropocollagen nanofibrils. International Journal of Materials Research, 2009.

100(7): p. 921-925.

48. Martin, R.B. and D.B. Burr, Structure, function, and adaptation of compact bone. 1989,

New York: Raven Press. xii, 275 p.

49. Currey, J.D., The design of mineralised hard tissues for their mechanical functions.

Journal of Experimental Biology, 1999. 202(23): p. 3285-3294.

50. John, D.C., Role of collagen and other organics in the mechanical properties of bone.

Osteoporos Int, 2003. 14 Suppl 5: p. S29-36.

51. Lowenstam, H.A. and S. Weiner, On biomineralization. 1989, New York: Oxford

University Press. ix, 324 p.

52. Bonucci, E., Calcification in biological systems. 1992, Boca Raton: CRC Press. 406 p.

53. Brighton, C.T. and R.M. Hunt, Histochemical-localization of calcium in the fracture

callus with potassium pyroantimonate - Possible role of chondrocyte mitochondrial

calcium in callus calcification. Journal of Bone and Joint Surgery-American Volume,

1986. 68A(5): p. 703-715.

54. Carter, J.G., Skeletal biomineralization : patterns, processes, and evolutionary trends.

1990, New York: Van Nostrand Reinhold.

Page 243: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

242

55. Thompson, D.A.W., On growth and form. 1992, New York: Dover. 1116 p.

56. Termine, J.D., Bone proteins and mineralization, in Kuehn, K. And T. Krieg. 1986. p.

184-196.

57. Pedrini-Mille, A., V. Pedrini, D.D. Hunt, and I.V. Ponseti, Chemical studies on the ground

substance of human epiphyseal-plate cartilage. The Journal of bone and joint surgery.

American volume, 1967. 49(8): p. 1628-35.

58. Cameron, D.A., The fine structure of bone and calcified cartilage. A critical review of

the contribution of electron microscopy to the understading of osteogenesis. Clin

Orthop Relat Res, 1963. 26: p. 199-228.

59. Morris, D.C., K. Masuhara, K. Takaoka, K. Ono, and H.C. Anderson, Immunolocalization

of alkaline-phosphatase in osteoblasts and matrix vesicles of human fetal bone. Bone

and Mineral, 1992. 19(3): p. 287-298.

60. Wuthier, R.E., A review of the primary mechanism of endochondral calcification with

special emphasis on the role of cells, mitochondria and matrix vesicles. Clinical

Orthopaedics and Related Research, 1982(169): p. 219-242.

61. Boskey, A.L. and A.S. Posner, Invitro nucleation of hydroxyapatite by a bone

calcium-phospholipid-phosphate complex. Calcified Tissue Research, 1977. 22: p.

197-201.

62. Boyan, B.D., Z. Schwartz, L.D. Swain, and A. Khare, Role of lipids in calcification of

cartilage. Anatomical Record, 1989. 224(2): p. 211-219.

63. Boyan, B.D., C.H. Lohmann, D.D. Dean, V.L. Sylvia, D.L. Cochran, and Z. Schwartz,

Mechanisms involved in osteoblast response to implant surface morphology. Annual

Review of Materials Research, 2001. 31: p. 357-371.

64. Brighton, C.T., Y. Sugioka, and R.M. Hunt, Cytoplasmic structures of epiphyseal plate

chondrocytes quantitive evaluation using electron micrographs of rat costo chondral

junctions with special reference to the fate of hypertrophic cells. Journal of Bone and

Joint Surgery American Volume, 1973. 55(4): p. 771-784.

65. Tortora, G.J. and M.T. Nielsen, Principles of human anatomy. 11th ed. 2009, Hoboken,

Page 244: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

243

NJ: J. Wiley.

66. Sarikaya, M. and I.A. Aksay, Biomimetics : design and processing of materials. AIP

series in polymers and complex materials. 1995, Woodbury, N.Y.: AIP Press. xi, 285 p.,

4 p. of plates.

67. Weiner, S. and P.A. Price, Disaggregation of bone into crystals. Calcified Tissue

International, 1986. 39(6): p. 365-375.

68. Fratzl, P., M. Groschner, G. Vogl, H. Plenk, J. Eschberger, N. Fratzlzelman, K. Koller, and

K. Klaushofer, Mineral crystals in calcified tissues - A coparative-study by SAXS. Journal

of Bone and Mineral Research, 1992. 7(3): p. 329-334.

69. Wachtel, E. and S. Weiner, Small-angle X-ray-scattering study of dispersed crystals

from bone and tendon. Journal of Bone and Mineral Research, 1994. 9(10): p.

1651-1655.

70. Siperko, L.M. and W.J. Landis, Atomic-force microscopy imaging of hydroxyapatite.

Journal of Materials Science Letters, 1993. 12(14): p. 1068-1069.

71. Delmas, P.D., R.P. Tracy, B.L. Riggs, and K.G. Mann, Identification of the noncollagenous

proteins of bovine bone by two-dimensional gel-electrophoresis. Calcified Tissue

International, 1984. 36(3): p. 308-316.

72. Giachelli, C.M. and S. Steitz, Osteopontin: a versatile regulator of inflammation and

biomineralization. Matrix Biology, 2000. 19(7): p. 615-622.

73. Gericke, A., C. Qin, L. Spevak, Y. Fujimoto, W.T. Butler, E.S. Sorensen, and A.L. Boskey,

Importance of phosphorylation for osteopontin regulation of biomineralization.

Calcified Tissue International, 2005. 77(1): p. 45-54.

74. Linde, A., Dentin and dentinogenesis. 1984, Boca Raton, Fla.: CRC Press. v. <1 >.

75. Nanci, A., Content and distribution of noncollagenous matrix proteins in bone and

cementum: Relationship to speed of formation and collagen packing density. Journal of

Structural Biology, 1999. 126(3): p. 256-269.

76. Bourne, G.H., The biochemistry and physiology of bone. 2d ed. 1971, New York,:

Academic Press. v.

Page 245: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

244

77. Hay, E.D., Cell biology of extracellular matrix. 2nd ed. 1991, New York: Plenum Press.

xvii, 468 p.

78. Glimcher, M.J., Recent studies of the mineral phase in bone and its possible linkage to

the organic matrix by protein-bound phosphate bonds. Philosophical Transactions of

the Royal Society of London Series B-Biological Sciences, 1984. 304(1121): p. 479-&.

79. Miller, A., Collagen - The organic matrix of bone. Philosophical Transactions of the

Royal Society of London Series B-Biological Sciences, 1984. 304(1121): p. 455-477.

80. Ramachandran, G.N. and University of Madras., Aspects of protein structure;

proceedings of a symposium held in Madras 14-18 January 1963 and organized by the

University of Madras. 1963, London, New York,: Academic Press. x, 380 p.

81. Yamauchi, M., E.P. Katz, K. Otsubo, K. Teraoka, and G.L. Mechanic, Cross-linking and

stereospecific structure of collagen in mineralized and nonmineralized skeletal tissues.

Connect Tissue Res, 1989. 21(1-4): p. 159-67; discussion 168-9.

82. Petruska, J.A. and A.J. Hodge, A subunit model for the tropocollagen macromolecule.

Proceedings of the National Academy of Sciences of the United States of America,

1964. 51: p. 871-6.

83. Maas, M., P. Guo, M. Keeney, F. Yang, T.M. Hsu, G.G. Fuller, C.R. Martin, and R.N. Zare,

Preparation of Mineralized Nanofibers: Collagen Fibrils Containing Calcium Phosphate.

Nano Letters, 2011. 11(3): p. 1383-1388.

84. Dorozhkin, S.V. and M. Epple, Biological and medical significance of calcium

phosphates. Angewandte Chemie-International Edition, 2002. 41(17): p. 3130-3146.

85. Suvorova, E.I., P.A. Stadelmann, and P.A. Buffat, HRTEM simulation in determination of

thickness and grain misorientation for hydroxyapatite crystals. Crystallography

Reports, 2004. 49(3): p. 343-352.

86. Suga, S., H. Nakahara, Nihon Gakujutsu Kaigi., and Kanagawa Academy of Science and

Technology., Mechanisms and phylogeny of mineralization in biological systems. 1991,

Tokyo New York: Springer-Verlag. xiv, 517 p.

87. Martin, R.B., D.B. Burr, and N.A. Sharkey, Skeletal tissue mechanics. 1998, New York:

Page 246: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

245

Springer. xiv, 392 p.

88. Reid, S.A., Micromorphological characterization of normal human-bone surfaces as a

function of age. Scanning Microscopy, 1987. 1(2): p. 579-597.

89. Weiner, S., T. Arad, and W. Traub, Crystal organizatioin in rat bone lamellae. Febs

Letters, 1991. 285(1): p. 49-54.

90. Marotti, G., Lees, and Reeve, A new theory of bone lamellation. Calcified Tissue

International, 1993. 53: p. S47-S56.

91. Weiner, S., T. Arad, I. Sabanay, and W. Traub, Rotated plywood structure of primary

lamellar bone in the rat: Orientations of the collagen fibril arrays. Bone, 1997. 20(6): p.

509-514.

92. Sabanay, I., T. Arad, S. Weiner, and B. Geiger, Study of vitrified, unstained frozen

tissue-sections by cryoimmunoelelctron microscopy. Journal of Cell Science, 1991. 100:

p. 227-236.

93. Ziv, V., I. Sabanay, T. Arad, W. Traub, and S. Weiner, Transitional structures in lamellar

bone. Microscopy Research and Technique, 1996. 33(2): p. 203-213.

94. Du, C., X.W. Su, F.Z. Cui, and X.D. Zhu, Morphological behaviour of osteoblasts on

diamond-like carbon coating and amorphous C-N film in organ culture. Biomaterials,

1998. 19(7-9): p. 651-658.

95. Akiva, U., H.D. Wagner, and S. Weiner, Modelling the three-dimensional elastic

constants of parallel-fibred and lamellar bone. Journal of Materials Science, 1998.

33(6): p. 1497-1509.

96. Havers, C., Osteologia Nova. 1691, London: Samuel Smith.

97. Bone Ossification & Growth. 1997; Available from:

https://courses.stu.qmul.ac.uk/smd/kb/microanatomy/bone/index.htm.

98. Uzwiak, A. Anatomy and Physiology course note. 2010; Available from:

http://www.rci.rutgers.edu/~uzwiak/AnatPhys/APFallLect8.html.

99. Currey, J.D., The mechanical adaptations of bones. 1984, Princeton, N.J.: Princeton

University Press. viii, 294 p.

Page 247: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

246

100. Giannuzzi, L.A. and F.A. Stevie, A review of focused ion beam milling techniques for

TEM specimen preparation. Micron, 1999. 30(3): p. 197-204.

101. Butler, M.F., A.M. Donald, and A.J. Ryan, Time resolved simultaneous small- and

wide-angle X-ray scattering during polyethylene deformation 3. Compression of

polyethylene. Polymer, 1998. 39(4): p. 781-792.

102. Fantner, G.E., H. Birkedal, J.H. Kindt, T. Hassenkam, J.C. Weaver, J.A. Cutroni, B.L.

Bosma, L. Bawazer, M.M. Finch, G.A.G. Cidade, D.E. Morse, G.D. Stucky, and P.K.

Hansma, Influence of the degradation of the organic matrix on the microscopic

fracture behavior of trabecular bone. Bone, 2004. 35(5): p. 1013-1022.

103. Rajaram, A. and N. Ramanathan, Tensile Properties of Antler Bone. Calcified Tissue

International, 1982. 34(3): p. 301-305.

104. Nalla, R.K., J.J. Kruzic, J.H. Kinney, M. Balooch, J.W. Ager, and R.O. Ritchie, Role of

microstructure in the aging-related deterioration of the toughness of human cortical

bone. Materials Science & Engineering C-Biomimetic and Supramolecular Systems,

2006. 26(8): p. 1251-1260.

105. Buehler, M.J., Nanomechanics of collagen fibrils under varying cross-link densities:

Atomistic and continuum studies. Journal of the Mechanical Behavior of Biomedical

Materials, 2008. 1(1): p. 59-67.

106. Siegmund, T., M.R. Allen, and D.B. Burr, Failure of mineralized collagen fibrils:

Modeling the role of collagen cross-linking. Journal of Biomechanics, 2008. 41(7): p.

1427-1435.

107. Berkson, E., G.H. Lee, A. Kumar, N. Verma, B.R. Bach, and N. Hallab, The effect of cyclic

loading on rotated bone-tendon-bone anterior cruciate ligament graft constructs.

American Journal of Sports Medicine, 2006. 34(9): p. 1442-1449.

108. Currey, J.D., Effects of strain rate, reconstruction and mineral content on some

mechanical-Properties of bovine bone. Journal of Biomechanics, 1975. 8(1): p. 81-&.

109. Reilly, D.T. and A.H. Burstein, The elastic and ultimate properties of compact bone

tissue. Journal of Biomechanics, 1975. 6(6): p. 393-405.

Page 248: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

247

110. Walsh, W.R., D.P. Labrador, H.D. Kim, and N. Guzelsu, Ultrasonic properties of cortical

bone following in-vitro fluoride-ion treatment. Hydroxyapatite and Related Materials,

ed. P.W. Brown and B. Constantz. 1994, Boca Raton: Crc Press Inc. 311-317.

111. Ashman, R.B., S.C. Cowin, W.C. Vanbuskirk, and J.C. Rice, A continuous wave technique

for the measurement of the elastic properties of cortical bone. Journal of Biomechanics,

1984. 17(5): p. 349-361.

112. Pope, M.H. and J.O. Outwater, Mechanical-Properties of bone as a function of position

and orientation. Journal of Biomechanics, 1974. 7(1): p. 61-66.

113. Liu, D., H.D. Wagner, and S. Weiner, Bending and fracture of compact circumferential

and osteonal lamellar bone of the baboon tibia. Journal of Materials Science-Materials

in Medicine, 2000. 11(1): p. 49-60.

114. Currey, J.D., K. Brear, and P. Zioupos, Dependence of Mechanical-Properties on Fiber

Angle in Narwhal Tusk, a Highly Oriented Biological Composite. Journal of

Biomechanics, 1994. 27(7): p. 885-&.

115. Currey, J.D., Strain rate and mineral-content in fracture models of bone. Journal of

Orthopaedic Research, 1988. 6(1): p. 32-38.

116. Currey, J.D., Strain rate dependence of the mechanical-properties of reindeer antler and

the cumulative damage model of bone-fracture. Journal of Biomechanics, 1989. 22(5):

p. 469-475.

117. Sammarco, G.J., A.H. Burstein, W.L. Davis, and V.H. Frankel, Biomechanics of torsional

fractures - Effect of Loading on ultimate properties Journal of Biomechanics, 1971.

4(2): p. 113-&.

118. Wright, T.M. and W.C. Hayes, Fracture mechanics of fatigue crack-Propagation in

compact bone. Journal of Biomedical Materials Research, 1976. 10(4): p. 637-648.

119. Wright, T.M. and W.C. Hayes, Tensile testing of bone over a wide-range of strain rates -

Effects of strain rate, microstructure and density. Medical & Biological Engineering,

1976. 14(6): p. 671-680.

120. Griffith, A.A., The phenomena of rupture and flow in solids. Philosophical Transactions

Page 249: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

248

of the Royal Society of London Series B-Biological Sciences, 1920. 221A: p. 163-197.

121. Warren, G.L., A.L. Moran, H.A. Hogan, A.S. Lin, R.E. Guldberg, and D.A. Lowe, Voluntary

run training but not estradiol deficiency alters the tibial bone-soleus muscle functional

relationship in mice. American Journal of Physiology-Regulatory Integrative and

Comparative Physiology, 2007. 293(5): p. R2015-R2026.

122. Phelps, J.B., G.B. Hubbard, X. Wang, and C.M. Agrawal, Microstructural heterogeneity

and the fracture toughness of bone. Journal of Biomedical Materials Research, 2000.

51(4): p. 735-741.

123. Kahler, B., M.V. Swain, and A. Moule, Fracture-toughening mechanisms responsible for

differences in work to fracture of hydrated and dehydrated dentine. Journal of

Biomechanics, 2003. 36(2): p. 229-237.

124. Ji, B.H. and H.J. Gao, A study of fracture mechanisms in biological nano-composites via

the virtual internal bond model. Materials Science and Engineering a-Structural

Materials Properties Microstructure and Processing, 2004. 366(1): p. 96-103.

125. Gupta, H.S., W. Wagermaier, G.A. Zickler, D.R.B. Aroush, S.S. Funari, P. Roschger, H.D.

Wagner, and P. Fratzl, Nanoscale deformation mechanisms in bone. Nano Letters, 2005.

5(10): p. 2108-2111.

126. Nalla, R.K., J.S. Stolken, J.H. Kinney, and R.O. Ritchie, Fracture in human cortical bone:

local fracture criteria and toughening mechanisms. Journal of Biomechanics, 2005.

38(7): p. 1517-1525.

127. Behiri, J.C. and W. Bonfield, Orientation dependence of the fracture-mechanics of

cortical bone. Journal of Biomechanics, 1989. 22(8-9): p. 863-&.

128. Koester, K.J., J.W. Ager, and R.O. Ritchie, The true toughness of human cortical bone

measured with realistically short cracks. Nature Materials, 2008. 7(8): p. 672-677.

129. Zioupos, P. and J.D. Currey, The Extent of Microcracking and the Morphology of

Microcracks in Damaged Bone. Journal of Materials Science, 1994. 29(4): p. 978-986.

130. Nalla, R.K., J.H. Kinney, and R.O. Ritchie, Mechanistic fracture criteria for the failure of

human cortical bone. Nature Materials, 2003. 2(3): p. 164-168.

Page 250: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

249

131. Yeni, Y.N. and T.L. Norman, Calculation of porosity and osteonal cement line effects on

the effective fracture toughness of cortical bone in longitudinal crack growth. Journal

of Biomedical Materials Research, 2000. 51(3): p. 504-509.

132. Peterlik, H., P. Roschger, K. Klaushofer, and P. Fratzl, From brittle to ductile fracture of

bone. Nature Materials, 2006. 5(1): p. 52-55.

133. Yeni, Y.N. and D.P. Fyhrie, A rate-dependent microcrack-bridging model that can

explain the strain rate dependency of cortical bone apparent yield strength. Journal of

Biomechanics, 2003. 36(9): p. 1343-1353.

134. Taylor, D., J.G. Hazenberg, and T.C. Lee, Living with cracks: Damage and repair in

human bone. Nature Materials, 2007. 6(4): p. 263-268.

135. Ritchie, R.O., R.M. Cannon, B.J. Dalgleish, R.H. Dauskardt, and J.M. McNaney,

Mechanics and mechanisms of crack-growth at or near ceramic-metal interfaces -

Interface enginerring strategies for promoting toughness. Materials Science and

Engineering a-Structural Materials Properties Microstructure and Processing, 1993.

166(1-2): p. 221-235.

136. Malik, C.L., S.M. Stover, R.B. Martin, and J.C. Gibeling, Equine cortical bone exhibits

rising R-curve fracture mechanics. Journal of Biomechanics, 2003. 36(2): p. 191-198.

137. Nalla, R.K., J.J. Kruzic, J.H. Kinney, and R.O. Ritchie, Effect of aging on the toughness of

human cortical bone: evaluation by R-curves. Bone, 2004. 35(6): p. 1240-1246.

138. Buehler, M.J., Molecular nanomechanics of nascent bone: fibrillar toughening by

mineralization. Nanotechnology, 2007. 18(29): p. 9.

139. Parsamian, G.P. and T.L. Norman, Diffuse damage accumulation in the fracture process

zone of human cortical bone specimens and its influence on fracture toughness. Journal

of Materials Science-Materials in Medicine, 2001. 12(9): p. 779-783.

140. Nalla, R.K., J.J. Kruzic, J.H. Kinney, and R.O. Ritchie, Mechanistic aspects of fracture and

R-curve behavior in human cortical bone. Biomaterials, 2005. 26(2): p. 217-231.

141. Runkle, J.C. and J. Pugh, The micro-mechanics of cancellous bone. II. Determination of

the elastic modulus of individual trabeculae by a buckling analysis. Bulletin of the

Page 251: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

250

Hospital for Joint Diseases, 1975. 36(1): p. 2-10.

142. Townsend, P.R., R.M. Rose, and E.L. Radin, Buckling studies of single human trabeculae.

Journal of Biomechanics, 1975. 8(3-4): p. 199-&.

143. Wolff, J., Das Gesetz der transformation der knochem. 1892, Berlin: Hirschwald.

144. Sargis, E.J. and M. Dagosto, Mammalian evolutionary morphology : a tribute to

Frederick S. Szalay. Vertebrate paleobiology and paleoanthropology series. 2008,

Dordrecht, The Netherlands: Springer. xxviii, 439 p.

145. Bendsøe, M.P., Optimization of structural topology, shape, and material. 1995, Berlin ;

New York: Springer. xi, 271 p.

146. Mullender, M.G., R. Huiskes, and H. Weinans, A physiological approach to the

simulation of bone remodeling as a self-organization control process. Journal of

Biomechanics, 1994. 27(11): p. 1389-1394.

147. Mullender, M.G. and R. Huiskes, Proposal for the regulatory mechanism of Wolff's law.

Journal of Orthopaedic Research, 1995. 13(4): p. 503-512.

148. Weiss, A.J., J. Iqbal, N. Zaidi, and J.I. Mechanick, The skeletal subsystem as an

integrative physiology paradigm. Current osteoporosis reports, 2010. 8(4): p. 168-77.

149. Van Rietbergen, B., R. Muller, D. Ulrich, P. Ruegsegger, and R. Huiskes, Tissue stresses

and strain in trabeculae of a canine proximal femur can be quantified from computer

reconstructions. Journal of Biomechanics, 1999. 32(2): p. 165-173.

150. Fyhrie, D.P. and M.B. Schaffler, Failure mechanisms in human vertebral cancellous bone.

Bone, 1994. 15(1): p. 105-109.

151. Buehler, M.J., S. Keten, and T. Ackbarow, Theoretical and computational hierarchical

nanomechanics of protein materials: Deformation and fracture. Progress in Materials

Science, 2008. 53(8): p. 1101-1241.

152. Gupta, H.S., P. Fratzl, M. Kerschnitzki, G. Benecke, W. Wagermaier, and H.O.K. Kirchner,

Evidence for an elementary process in bone plasticity with an activation enthalpy of 1

eV. Journal of the Royal Society Interface, 2007. 4(13): p. 277-282.

153. Gupta, H.S. and P. Zioupos, Fracture of bone tissue: The 'hows' and the 'whys'. Medical

Page 252: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

251

Engineering & Physics, 2008. 30(10): p. 1209-1226.

154. Krauss, S., P. Fratzl, J. Seto, J.D. Currey, J.A. Estevez, S.S. Funari, and H.S. Gupta,

Inhomogeneous fibril stretching in antler starts after macroscopic yielding: Indication

for a nanoscale toughening mechanism. Bone, 2009. 44(6): p. 1105-1110.

155. Roeder, B.A., K. Kokini, J.E. Sturgis, J.P. Robinson, and S.L. Voytik-Harbin, Tensile

mechanical properties of three-dimensional type I collagen extracellular matrices with

varied microstructure. Journal of Biomechanical Engineering-Transactions of the

Asme, 2002. 124(2): p. 214-222.

156. Viguet-Carrin, S., P. Garnero, and P.D. Delmas, The role of collagen in bone strength.

Osteoporosis International, 2006. 17(3): p. 319-336.

157. Christiansen, D.L., E.K. Huang, and F.H. Silver, Assembly of type I collagen: fusion of

fibril subunits and the influence of fibril diameter on mechanical properties. Matrix

Biology, 2000. 19(5): p. 409-420.

158. Sun, Y.L., Z.P. Luo, A. Fertala, and K.N. An, Direct quantification of the flexibility of type

I collagen monomer. Biochemical and Biophysical Research Communications, 2002.

295(2): p. 382-386.

159. Sun, Y.L., Z.P. Luo, A. Fertala, and K.N. An, Stretching type II collagen with optical

tweezers. Journal of Biomechanics, 2004. 37(11): p. 1665-1669.

160. Buehler, M.J., Atomistic and continuum modeling of mechanical properties of collagen:

Elasticity, fracture, and self-assembly. Journal of Materials Research, 2006. 21(8): p.

1947-1961.

161. Buehler, M.J. and S.Y. Wong, Entropic elasticity controls nanomechanics of single

tropocollagen molecules. Biophysical Journal, 2007. 93(1): p. 37-43.

162. Fantner, G.E., T. Hassenkam, J.H. Kindt, J.C. Weaver, H. Birkedal, L. Pechenik, J.A.

Cutroni, G.A.G. Cidade, G.D. Stucky, D.E. Morse, and P.K. Hansma, Sacrificial bonds and

hidden length dissipate energy as mineralized fibrils separate during bone fracture.

Nature Materials, 2005. 4(8): p. 612-616.

163. Craig, A.S., M.J. Birtles, J.F. Conway, and D.A.D. Parry, An estimate of the mean length of

Page 253: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

252

collagen fibrils in rat tail-tendon as a function of age. Connective Tissue Research,

1989. 19(1): p. 51-62.

164. Barber, A.H., R. Andrews, L.S. Schadler, and H.D. Wagner, On the tensile strength

distribution of multiwalled carbon nanotubes. Applied Physics Letters, 2005. 87(20):

p. 3.

165. Smith, B.L., T.E. Schaffer, M. Viani, J.B. Thompson, N.A. Frederick, J. Kindt, A. Belcher,

G.D. Stucky, D.E. Morse, and P.K. Hansma, Molecular mechanistic origin of the

toughness of natural adhesives, fibres and composites. Nature, 1999. 399(6738): p.

761-763.

166. Bonar, L.C., S. Lees, and H.A. Mook, Neutron-diffraction studies of collagen in fully

mineralized bone. Journal of Molecular Biology, 1985. 181(2): p. 265-270.

167. Hansma, P.K., G.E. Fantner, J.H. Kindt, P.J. Thurner, G. Schitter, P.J. Turner, S.F. Udwin,

and M.M. Finch, Sacrificial bonds in the interfibrillar matrix of bone. J Musculoskelet

Neuronal Interact, 2005. 5(4): p. 313-5.

168. Hartmann, M.A. and P. Fratzl, Sacrificial Ionic Bonds Need To Be Randomly Distributed

To Provide Shear Deformability. Nano Letters, 2009. 9(10): p. 3603-3607.

169. Zioupos, P., J.D. Currey, and A.J. Sedman, An Examination of the Micromechanics of

Failure of Bone and Antler by Acoustic-Emission Tests and

Laser-Scanning-Confocal-Microscopy. Medical Engineering & Physics, 1994. 16(3): p.

203-212.

170. Chen, P.Y., A.G. Stokes, and J. McKittrick, Comparison of the structure and mechanical

properties of bovine femur bone and antler of the North American elk (Cervus elaphus

canadensis). Acta Biomaterialia, 2009. 5(2): p. 693-706.

171. Currey, J.D., T. Landete-Castillejos, J. Estevez, F. Ceacero, A. Olguin, A. Garcia, and L.

Gallego, The mechanical properties of red deer antler bone when used in fighting.

Journal of Experimental Biology, 2009. 212(24): p. 3985-3993.

172. Barber, A.H., S.R. Cohen, A. Eitan, L.S. Schadler, and H.D. Wagner, Fracture transitions

at a carbon-nanotube/polymer interface. Advanced Materials, 2006. 18(1): p. 83-87.

Page 254: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

253

173. Ji, Y., B.Q. Li, S.R. Ge, J.C. Sokolov, and M.H. Rafailovich, Structure and nanomechanical

characterization of electrospun PS/clay nanocomposite fibers. Langmuir, 2006. 22(3):

p. 1321-1328.

174. Liu, L.Q., A.H. Barber, S. Nuriel, and H.D. Wagner, Mechanical properties of

functionalized single-walled carbon-nanotube/poly(vinyl alcohol) nanocomposites.

Advanced Functional Materials, 2005. 15(6): p. 975-980.

175. Inai, R., M. Kotaki, and S. Ramakrishna, Structure and properties of electrospun PLLA

single nanofibres. Nanotechnology, 2005. 16(2): p. 208-213.

176. Wong, E.W., P.E. Sheehan, and C.M. Lieber, Nanobeam mechanics: Elasticity, strength,

and toughness of nanorods and nanotubes. Science, 1997. 277(5334): p. 1971-1975.

177. Wang, W., A.J. Bushby, and A.H. Barber, Nanomechanical thermal analysis of

electrospun polymer fibers. Applied Physics Letters, 2008. 93(20).

178. Mayer, G., Rigid biological systems as models for synthetic composites. Science, 2005.

310(5751): p. 1144-1147.

179. Heuer, A.H., D.J. Fink, V.J. Laraia, J.L. Arias, P.D. Calvert, K. Kendall, G.L. Messing, J.

Blackwell, P.C. Rieke, D.H. Thompson, A.P. Wheeler, A. Veis, and A.I. Caplan, Innovative

materials processing strategies a biomimetic approach. Science, 1992. 255(5048): p.

1098-1105.

180. Kikuchi, M., T. Ikoma, S. Itoh, H.N. Matsumoto, Y. Koyama, K. Takakuda, K. Shinomiya,

and J. Tanaka, Biomimetic synthesis of bone-like nanocomposites using the

self-organization mechanism of hydroxyapatite and collagen. Composites Science and

Technology, 2004. 64(6): p. 819-825.

181. Wang, X.J., Y.B. Li, J. Wei, and K. de Groot, Development of biomimetic

nano-hydroxyapatite/poly(hexamethylene adipamide) composites. Biomaterials, 2002.

23(24): p. 4787-4791.

182. Iwamoto, S., W.H. Kai, T. Isogai, T. Saito, A. Isogai, and T. Iwata, Comparison study of

TEMPO-analogous compounds on oxidation efficiency of wood cellulose for

preparation of cellulose nanofibrils. Polymer Degradation and Stability, 2010. 95(8): p.

Page 255: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

254

1394-1398.

183. Oliver, W.C. and G.M. Pharr, An improved technique for determining hardness and

elastic-modulus using load and displacement sensing indentation experiments. Journal

of Materials Research, 1992. 7(6): p. 1564-1583.

184. Kermouche, G., J.L. Loubet, and J.M. Bergheau, Extraction of stress-strain curves of

elastic-viscoplastic solids using conical/pyramidal indentation testing with application

to polymers. Mechanics of Materials, 2008. 40(4-5): p. 271-283.

185. Kim, J.Y., B.W. Lee, D.T. Read, and D. Kwon, Influence of tip bluntness on the

size-dependent nanoindentation hardness. Scripta Materialia, 2005. 52(5): p.

353-358.

186. Bull, S.J., Nano-indentation of coatings. Journal of Physics D-Applied Physics, 2005.

38(24): p. R393-R413.

187. Constantinides, G., K.S.R. Chandran, F.J. Ulm, and K.J. Van Vliet, Grid indentation

analysis of composite microstructure and mechanics: Principles and validation.

Materials Science and Engineering a-Structural Materials Properties Microstructure

and Processing, 2006. 430(1-2): p. 189-202.

188. Wang, M.J., K.M. Liechti, J.M. White, and R.M. Winter, Nanoindentation of polymeric

thin films with an interfacial force microscope. Journal of the Mechanics and Physics

of Solids, 2004. 52(10): p. 2329-2354.

189. Thormann, E., A.C. Simonsen, P.L. Hansen, and O.G. Mouritsen, Force trace hysteresis

and temperature dependence of bridging nanobubble induced forces between

hydrophobic surfaces. Acs Nano, 2008. 2(9): p. 1817-1824.

190. Thormann, E., A.C. Simonsen, P.L. Hansen, and O.G. Mouritsen, Interactions between a

polystyrene particle and hydrophilic and hydrophobic surfaces in aqueous solutions.

Langmuir, 2008. 24(14): p. 7278-7284.

191. Ducker, W.A., T.J. Senden, and R.M. Pashley, Direct Measurement of Colloidal Forces

Using an Atomic Force Microscope. Nature, 1991. 353(6341): p. 239-241.

192. Cappella, B. and G. Dietler, Force-distance curves by atomic force microscopy. Surface

Page 256: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

255

Science Reports, 1999. 34(1-3): p. 1-+.

193. Pettersson, T. and A. Dedinaite, Normal and friction forces between mucin and

mucin-chitosan layers in absence and presence of SDS. Journal of Colloid and Interface

Science, 2008. 324(1-2): p. 246-256.

194. Wang, W. and A.H. Barber, Diameter-dependent melting behaviour in electrospun

polymer fibres. Nanotechnology, 2010. 21(22).

195. Strunz, T., K. Oroszlan, R. Schafer, and H.J. Guntherodt, Dynamic force spectroscopy of

single DNA molecules. Proceedings of the National Academy of Sciences of the United

States of America, 1999. 96(20): p. 11277-11282.

196. Thormanna, E., J.K. Dreyer, A.C. Simonsen, P.L. Hansen, S. Hansen, U. Holmskov, and

O.G. Mouritsen, Dynamic strength of the interaction between lung surfactant protein D

(SP-D) and saccharide Ligands. Biochemistry, 2007. 46(43): p. 12231-12237.

197. Rief, M., M. Gautel, F. Oesterhelt, J.M. Fernandez, and H.E. Gaub, Reversible unfolding

of individual titin immunoglobulin domains by AFM. Science, 1997. 276(5315): p.

1109-1112.

198. Li, H., M. Rief, F. Oesterhelt, and H.E. Gaub, Force spectroscopy on single xanthan

molecules. Applied Physics a-Materials Science & Processing, 1999. 68(4): p.

407-410.

199. Hugel, T., M. Grosholz, H. Clausen-Schaumann, A. Pfau, H. Gaub, and M. Seitz,

Elasticity of single polyelectrolyte chains and their desorption from solid supports

studied by AFM based single molecule force spectroscopy. Macromolecules, 2001.

34(4): p. 1039-1047.

200. Yu, M.F., O. Lourie, M.J. Dyer, K. Moloni, T.F. Kelly, and R.S. Ruoff, Strength and breaking

mechanism of multiwalled carbon nanotubes under tensile load. Science, 2000.

287(5453): p. 637-640.

201. Tan, E.P.S., C.N. Goh, C.H. Sow, and C.T. Lim, Tensile test of a single nanofiber using an

atomic force microscope tip. Applied Physics Letters, 2005. 86(7): p. 3.

202. Zussman, E., M. Burman, A.L. Yarin, R. Khalfin, and Y. Cohen, Tensile deformation of

Page 257: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

256

electrospun nylon-6,6 nanofibers. Journal of Polymer Science Part B-Polymer Physics,

2006. 44(10): p. 1482-1489.

203. Altman, G.H., R.L. Horan, H.H. Lu, J. Moreau, I. Martin, J.C. Richmond, and D.L. Kaplan,

Silk matrix for tissue engineered anterior cruciate ligaments. Biomaterials, 2002.

23(20): p. 4131-4141.

204. Tan, E.P.S., S.Y. Ng, and C.T. Lim, Tensile testing of a single ultrafine polymeric fiber.

Biomaterials, 2005. 26(13): p. 1453-1456.

205. Naraghi, M., I. Chasiotis, H. Kahn, Y.K. Wen, and Y. Dzenis, Mechanical deformation and

failure of electrospun polyacrylonitrile nanofibers as a function of strain rate. Applied

Physics Letters, 2007. 91.

206. Hwang, K.Y., S.D. Kim, Y.W. Kim, and W.R. Yu, Mechanical characterization of

nanofibers using a nanomanipulator and atomic force microscope cantilever in a

scanning electron microscope. Polymer Testing, 2010. 29(3): p. 375-380.

207. Culler, R.D., F.C. Parrish, G.C. Smith, and H.R. Cross, Relationship of myofibril

fragmentation index to certain chemical, physical and sensory characteristics of bovine

longissiums muscle. Journal of Food Science, 1978. 43(4): p. 1177-1180.

208. Hang, F., D. Lu, S.W. Li, and A.H. Barber, Stress-strain behavior of individual electrospun

polymer fibers using combination AFM and SEM. Probing Mechanics at Nanoscale

Dimensions, 2009. 1185: p. 87-91.

209. Hang, F. and A.H. Barber, Nano-mechanical properties of individual mineralized

collagen fibrils from bone tissue. Journal of The Royal Society Interface, 2010.

210. Attocube. Brochure of attoAFM systems. 2011; Available from:

http://www.attocube.com/nanoSCOPY/download/Brochure_attoAFM.pdf.

211. Fukuda, T., M. Nakajima, M.R. Ahmad, Y.J. Shen, and M. Kojima, Micro- and

Nanomechatronics. Ieee Industrial Electronics Magazine, 2010. 4(4): p. 13-22.

212. Beyeler, F., S. Muntwyler, and B.J. Nelson, A Six-Axis MEMS Force-Torque Sensor With

Micro-Newton and Nano-Newtonmeter Resolution. Journal of Microelectromechanical

Systems, 2009. 18(2): p. 433-441.

Page 258: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

257

213. Abbott, J.J., Z. Nagy, F. Beyeler, and B.J. Nelson, Robotics in the small - Part I:

microrobotics. Ieee Robotics & Automation Magazine, 2007. 14(2): p. 92-103.

214. Arai, F., D. Ando, T. Fukuda, Y. Nonoda, and T. Oota, Micro manipulation based on

micro physics - Strategy based on attractive force reduction and stress measurement.

Iros '95 - 1995 Ieee/Rsj International Conference on Intelligent Robots and Systems:

Human Robot Interaction and Cooperative Robots, Proceedings, Vol 2, ed. K. Ikeuchi

and P. Khosla. 1995, Los Alamitos: I E E E, Computer Soc Press. 236-241.

215. FEI. Quanta™ Scanning Electron Microscope: Specifications. 2011; Available from:

http://www.fei.com/products/scanning-electron-microscopes/quanta.aspx.

216. Wang, W., P. Ciselli, E. Kuznetsov, T. Peijs, and A.H. Barber, Effective reinforcement in

carbon nanotube - polymer composites. Philosophical Transactions of the Royal

Society a-Mathematical Physical and Engineering Sciences, 2008. 366(1870): p.

1613-1626.

217. Seguela, R. and F. Rietsch, Double yield-point in polyethylene under tensile loading.

Journal of Materials Science Letters, 1990. 9(1): p. 46-47.

218. Muramatsu, S. and J.B. Lando, Double yield points in poly(tetramethylene

terephthalate) and its copolymers under tensile loading. Polymer Engineering and

Science, 1995. 35(13): p. 1077-1085.

219. Brooks, N.W.J., R.A. Duckett, and I.M. Ward, Modeling of double yield points in

polyethylene - temperature and strain-rate dependence. Journal of Rheology, 1995.

39(2): p. 425-436.

220. Plaza, A.R., E. Ramos, A. Manzur, R. Olayo, and A. Escobar, Double yield points in

triblends of LDPE, LLDPE and EPDM. Journal of Materials Science, 1997. 32(2): p.

549-554.

221. Takenaka, K. and S. Muramatsu, Effects of copolymerization and temperature on

double yield points in poly(butylene terephthalate). Kobunshi Ronbunshu, 1997.

54(4): p. 285-287.

222. Liu, Y., S. Chen, E. Zussman, C.S. Korach, W. Zhao, and M. Rafailovich,

Page 259: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

258

Diameter-Dependent Modulus and Melting Behavior in Electrospun Semicrystalline

Polymer Fibers. Macromolecules, 2011. 44(11): p. 4439-4444.

223. Lim, C.T., E.P.S. Tan, and S.Y. Ng, Effects of crystalline morphology on the tensile

properties of electrospun polymer nanofibers. Applied Physics Letters, 2008. 92(14).

224. Arinstein, A., M. Burman, O. Gendelman, and E. Zussman, Effect of supramolecular

structure on polymer nanofibre elasticity. Nature Nanotechnology, 2007. 2(1): p.

59-62.

225. Ji, Y., C. Li, G. Wang, J. Koo, S. Ge, B. Li, J. Jiang, B. Herzberg, T. Klein, S. Chen, J.C.

Sokolov, and M.H. Rafailovich, Confinement-induced super strong PS/MWNT

composite nanofibers. Epl, 2008. 84(5).

226. Arzt, E., Overview no. 130 - Size effects in materials due to microstructural and

dimensional constraints: A comparative review. Acta Materialia, 1998. 46(16): p.

5611-5626.

227. Hang, F., D. Lu, S.W. Li, and A.H. Barber. Stress-strain behavior of individual electrospun

polymer fibers using combination AFM and SEM. in 2009 MRS Spring Meeting. 2009.

San Francisco, USA: Institute of Physics Publishing.

228. Tai, K. and C. Ortiz, Nanomechanical heterogeneity of bone at the length scale of

individual collagen fibrils. Abstracts of Papers American Chemical Society, 2006. 231:

p. 179-PMSE.

229. Boyde, A., Bone and BSE. Veterinary Record, 1998. 142(11): p. 288-288.

230. Eidelman, N., A. Boyde, A.J. Bushby, P.G.T. Howell, J.R. Sun, D.E. Newbury, F.W. Miller,

P.G. Robey, and L.G. Rider, Microstructure and mineral composition of dystrophic

calcification associated with the idiopathic inflammatory myopathies. Arthritis

Research & Therapy, 2009. 11(5): p. 21.

231. Cribb, A.M. and J.E. Scott, Tenden response to tensile-strss- An untrastructural

investigation of collagen-proteoglycan interactions in stressed tendon. Journal of

Anatomy, 1995. 187: p. 423-428.

232. Gautieri, A., S. Vesentini, A. Redaelli, and M.J. Buehler, Single molecule effects of

Page 260: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

259

osteogenesis imperfecta mutations in tropocollagen protein domains. Protein Science,

2009. 18(1): p. 161-168.

233. Dubey, D.K. and V. Tomar, Tensile and Compressive Loading Effects on Texture

Dependent Nanoscale Mechanical Behavior of Model Tropocollagen-Hydroxyapatite

Biomaterials. Journal of Computational and Theoretical Nanoscience. 7(7): p.

1306-1316.

234. Dubey, D.K. and V. Tomar, Role of the nanoscale interfacial arrangement in mechanical

strength of tropocollagen-hydroxyapatite-based hard biomaterials. Acta Biomaterialia,

2009. 5(7): p. 2704-2716.

235. Block, J.E., R. Smith, C.C. Glueer, P. Steiger, B. Ettinger, and H.K. Genant, Models of

spinal trabecular bone loss as determined by quantitative computed-tomography.

Journal of Bone and Mineral Research, 1989. 4(2): p. 249-257.

236. Genant, H.K., J.E. Block, P. Steiger, C.C. Glueer, and R. Smith, Quantitative

computed-tomography in assessment of osteoporosis. Seminars in Nuclear Medicine,

1987. 17(4): p. 316-333.

237. Looker, A.C., H.W. Wahner, W.L. Dunn, M.S. Calvo, T.B. Harris, S.P. Heyse, C.C. Johnston,

and R. Lindsay, Updated data on proximal femur bone mineral levels of US adults.

Osteoporosis International, 1998. 8(5): p. 468-489.

238. Cummings, S.R., Are patients with hip-fractures more osteoporotic - Review of the

evidence. American Journal of Medicine, 1985. 78(3): p. 487-494.

239. Marshall, D., O. Johnell, and H. Wedel, Meta-analysis of how well measures of bone

mineral density predict occurrence of osteoporotic fractures. British Medical Journal,

1996. 312(7041): p. 1254-1259.

240. McCreadie, B.R. and S.A. Goldstein, Biomechanics of fracture: Is bone mineral density

sufficient to assess risk? Journal of Bone and Mineral Research, 2000. 15(12): p.

2305-2308.

241. Sandor, T., D. Felsenberg, and E. Brown, Comments on the hypotheses underlying

fracture risk assessment in osteoporosis as proposed by the World Health Organization.

Page 261: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

260

Calcified Tissue International, 1999. 64(3): p. 267-270.

242. Sandor, T., D. Felsenberg, and E. Brown, Discriminability of fracture and nonfracture

cases based on the spatial distribution of spinal bone mineral. Journal of Computer

Assisted Tomography, 1997. 21(3): p. 498-505.

243. Wilkin, T.J., Changing perceptions in osteoporosis. British Medical Journal, 1999.

318(7187): p. 862-864.

244. Akkus, O., F. Adar, and M.B. Schaffler, Age-related changes in physicochemical

properties of mineral crystals are related to impaired mechanical function of cortical

bone. Bone, 2004. 34(3): p. 443-453.

245. Currey, J.D., K. Brear, and P. Zioupos, The effects of ageing and changes in mineral

content in degrading the toughness of human femora. Journal of Biomechanics, 1996.

29(2): p. 257-260.

246. Ding, M., M. Dalstra, C.C. Danielsen, J. Kabel, I. Hvid, and F. Linde, Age variations in the

properties of human tibial trabecular bone. Journal of Bone and Joint Surgery-British

Volume, 1997. 79B(6): p. 995-1002.

247. McCalden, R.W., J.A. McGeough, M.B. Barker, and C.M. Courtbrown, Age

related-changes in the tensile properties of cortical bone - The relative importance of

changes in porosity, mineralization, and microstructure. Journal of Bone and Joint

Surgery-American Volume, 1993. 75A(8): p. 1193-1205.

248. McCalden, R.W., J.A. McGeough, and C.M. CourtBrown, Age-related changes in the

compressive strength of cancellous bone - The relative importance of changes in density

and trabecular architecture. Journal of Bone and Joint Surgery-American Volume,

1997. 79A(3): p. 421-427.

249. Currey, J.D., Effects of differences in mineralization on the mechanical-Properties of

bone. Philosophical Transactions of the Royal Society of London Series B-Biological

Sciences, 1984. 304(1121): p. 509-&.

250. Currey, J.D., P. Zioupos, P. Davies, and A. Casinos, Mechanical properties of nacre and

highly mineralized bone. Proceedings of the Royal Society of London Series

Page 262: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

261

B-Biological Sciences, 2001. 268(1462): p. 107-111.

251. Ji, B.H. and H.J. Gao, Mechanical Principles of Biological Nanocomposites, in Annual

Review of Materials Research, Vol 40. 2010, Annual Reviews: Palo Alto. p. 77-100.

252. Fratzl, P., F. Hang, and R.O. Ritchie, Personal communication during 2009 MRS spring

meeting. 2009: San Fronsisco.

253. Li, X.D., W.C. Chang, Y.J. Chao, R.Z. Wang, and M. Chang, Nanoscale structural and

mechanical characterization of a natural nanocomposite material: The shell of red

abalone. Nano Letters, 2004. 4(4): p. 613-617.

254. Yao, H.M., M. Dao, T. Imholt, J.M. Huang, K. Wheeler, A. Bonilla, S. Suresh, and C. Ortiz,

Protection mechanisms of the iron-plated armor of a deep-sea hydrothermal vent

gastropod. Proceedings of the National Academy of Sciences of the United States of

America, 2010. 107(3): p. 987-992.

255. Chai, H., J.J.W. Lee, P.J. Constantino, P.W. Lucas, and B.R. Lawn, Remarkable resilience

of teeth. Proceedings of the National Academy of Sciences of the United States of

America, 2009. 106(18): p. 7289-7293.

256. Bonderer, L.J., A.R. Studart, and L.J. Gauckler, Bioinspired design and assembly of

platelet reinforced polymer films. Science, 2008. 319(5866): p. 1069-1073.

257. Podsiadlo, P., A.K. Kaushik, E.M. Arruda, A.M. Waas, B.S. Shim, J.D. Xu, H. Nandivada,

B.G. Pumplin, J. Lahann, A. Ramamoorthy, and N.A. Kotov, Ultrastrong and stiff layered

polymer nanocomposites. Science, 2007. 318(5847): p. 80-83.

258. Munch, E., M.E. Launey, D.H. Alsem, E. Saiz, A.P. Tomsia, and R.O. Ritchie, Tough,

Bio-Inspired Hybrid Materials. Science, 2008. 322(5907): p. 1516-1520.

259. Currey, J.D., Materials science - Hierarchies in biomineral structures. Science, 2005.

309(5732): p. 253-254.

260. Hang, F. and A.H. Barber, Nano-mechanical properties of individual mineralized

collagen fibrils from bone tissue. Journal of the Royal Society Interface, 2011. 8(57): p.

500-505.

261. Gupta, H.S., J. Seto, W. Wagermaier, P. Zaslansky, P. Boesecke, and P. Fratzl, Cooperative

Page 263: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

262

deformation of mineral and collagen in bone at the nanoscale. Proceedings of the

National Academy of Sciences of the United States of America, 2006. 103(47): p.

17741-17746.

262. Sodek, J., B. Ganss, and M.D. McKee, Osteopontin. Critical Reviews in Oral Biology &

Medicine, 2000. 11(3): p. 279-303.

263. Arteaga-Solis, E., L. Sui-Arteaga, M. Kim, M.B. Schaffler, K.J. Jepsen, N. Pleshko, and F.

Ramirez, Material and mechanical properties of bones deficient for fibrillin-1 or

fibrillin-2 microfibrils. Matrix Biology, 2011. 30(3): p. 188-194.

264. Barber, A.H., S.R. Cohen, S. Kenig, and H.D. Wagner, Interfacial fracture energy

measurements for multi-walled carbon nanotubes pulled from a polymer matrix.

Composites Science and Technology, 2004. 64(15): p. 2283-2289.

265. Gong, L., I.A. Kinloch, R.J. Young, I. Riaz, R. Jalil, and K.S. Novoselov, Interfacial Stress

Transfer in a Graphene Monolayer Nanocomposite. Advanced Materials, 2010. 22(24):

p. 2694-+.

266. Hsueh, C.H., Interfacial debonding and fiber pull-out stresses of fiber-reinforced

composites .6. interpretation of fiber pull-out curves. Materials Science and

Engineering a-Structural Materials Properties Microstructure and Processing, 1991.

149(1): p. 11-18.

267. Kim, J.K., C. Baillie, and Y.W. Mai, Interfacial debonding and fiber pull-out stress .1.

Critical coparision of existing theories with experiments. Journal of Materials Science,

1992. 27(12): p. 3143-3154.

268. Yue, C.Y. and W.L. Cheung, Interfacial properties of fibre-reinforced composites. Journal

of Materials Science, 1992. 27(14): p. 3843-3855.

269. Nuriel, S., A. Katz, and H.D. Wagner, Measuring fiber-matrix interfacial adhesion by

means of a 'drag-out' micromechanical test. Composites Part a-Applied Science and

Manufacturing, 2005. 36(1): p. 33-37.

270. Delfolie, C., C. Depecker, and J.M. Lefebvre, Interfacial phenomena in glass fibre

reinforced polyester resin with low profile additives - Part I - Micromechanical

Page 264: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

263

evaluation by pull out testing. Journal of Materials Science, 1999. 34(3): p. 481-495.

271. Scott, J.E., Proteoglycan fibrillar collagen interactions. Biochemical Journal, 1988.

252(2): p. 313-323.

272. Scott, J.E. and M. Haigh, Proteoglycan-type-I collagen fibril interactions in bone and

non-calcifying connective tissues. Bioscience Reports, 1985. 5(1): p. 71-81.

273. Heiss, A., A. DuChesne, B. Denecke, J. Grotzinger, K. Yamamoto, T. Renne, and W.

Jahnen-Dechent, Structural basis of calcification inhibition by alpha(2)-HS

glycoprotein/fetuin-A - Formation of colloidal calciprotein particles. Journal of

Biological Chemistry, 2003. 278(15): p. 13333-13341.

274. Currey, J., Biomaterials - Sacrificial bonds heal bone. Nature, 2001. 414(6865): p.

699-699.

275. Salih, E., S. Ashkar, L.C. Gerstenfeld, and M.J. Glimcher, Identification of the

phosphorylated sites of metabolically P-32-labeled osteopontin from cultured chicken

osteoblasts. Journal of Biological Chemistry, 1997. 272(21): p. 13966-13973.

276. Fantner, G.E., J. Adams, P. Turner, P.J. Thurner, L.W. Fisher, and P.K. Hansma, Nanoscale

ion mediated networks in bone: Osteopontin can repeatedly dissipate large amounts of

energy. Nano Letters, 2007. 7(8): p. 2491-2498.

277. Chua, P.S. and M.R. Piggott, The glass-fiber polymer interface. 1. Theoretical

consideration for ingle fibre pull-out. Composites Science and Technology, 1985.

22(1): p. 33-42.

278. Jiang, K.R. and L.S. Penn, Improved analysis and experimental evaluation of the single

filament pull-out test. Composites Science and Technology, 1992. 45(2): p. 89-103.

279. Akhtar, R., M.R. Daymond, J.D. Almer, and P.M. Mummery, Elastic strains in antler

trabecular bone determined by synchrotron X-ray diffraction. Acta Biomaterialia, 2008.

4(6): p. 1677-1687.

280. Galiotis, C., Interfacial studies on model composites by laser Raman spectroscopy.

Composites Science and Technology, 1991. 42(1-3): p. 125-150.

281. Detassis, M., E. Frydman, D. Vrieling, X.F. Zhou, H.D. Wagner, and J.A. Nairn, Interface

Page 265: in situ Nanomechanical Investigations of Bone · 2016. 4. 17. · in situ nanomechanical investigations of bone a thesis submitted to the university of london for the degree of doctor

References

264

toughness in fibre composites by the fragmentation test. Composites Part A: Applied

Science and Manufacturing, 1996. 27(9): p. 769-773.

282. Zhandarov, S., E. Pisanova, E. Mader, and J.A. Nairn, Investigation of load transfer

between the fiber and the matrix in pull-out tests with fibers having different diameters.

Journal of Adhesion Science and Technology, 2001. 15(2): p. 205-222.

283. Phillipps, A.J., W.J. Clegg, and T.W. Clyne, Fracture behavior of ceramic laminates in

bending. 2. Comparision of model predictions with experimental data. Acta

Metallurgica Et Materialia, 1993. 41(3): p. 819-827.

284. Gostemeyer, G., M. Jendras, M.P. Dittmer, F.W. Bach, M. Stiesch, and P. Kohorst,

Influence of cooling rate on zirconia/veneer interfacial adhesion. Acta Biomaterialia,

2010. 6(12): p. 4532-4538.

285. Lin, S., F. Narita, and Y. Shindo, Comparison of energy release rate and energy density

criteria for a piezoelectric layered composite with a crack normal to interface.

Theoretical and Applied Fracture Mechanics, 2003. 39(3): p. 229-243.

286. Vashishth, D., K.E. Tanner, and W. Bonfield, Experimental validation of a

microcracking-based toughening mechanism for cortical bone. Journal of

Biomechanics, 2003. 36(1): p. 121-124.

287. Vashishth, D., Rising crack-growth-resistance behavior in cortical bone: implications

for toughness measurements. Journal of Biomechanics, 2004. 37(6): p. 943-946.

288. Launey, M.E., P.Y. Chen, J. McKittrick, and R.O. Ritchie, Mechanistic aspects of the

fracture toughness of elk antler bone. Acta Biomaterialia, 2010. 6(4): p. 1505-1514.