impact of mechanical stimulation on …1 impact of mechanical stimulation on tendon tissue in a...

252
1 IMPACT OF MECHANICAL STIMULATION ON TENDON TISSUE IN A BIOREACTOR SYSTEM TAO WANG This thesis is submitted to The University of Western Australia in fulfillment of the requirement for degree of DOCTOR OF PHILOSOPHY IN BIOMEDICAL ENGINEERING School of Surgery The University of Western Australia 2014 The work presented in this thesis was performed in The University of Western Australia, School of Surgery, Centre for Orthopaedic Translational Research

Upload: others

Post on 01-Apr-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

1

IMPACT OF MECHANICAL STIMULATION ON TENDON TISSUE IN A

BIOREACTOR SYSTEM

TAO WANG This thesis is submitted to The University of Western Australia in fulfillment

of the requirement for degree of

DOCTOR OF PHILOSOPHY

IN

BIOMEDICAL ENGINEERING School of Surgery

The University of Western Australia

2014

The work presented in this thesis was performed in The University of Western

Australia, School of Surgery, Centre for Orthopaedic Translational Research

2

Declaration

This is to certify that all work contained herein was performed by myself, except where

indicated otherwise.

Tao Wang (PhD candidate)

W/Prof. Minghao Zheng (Supervisor)

A/Prof. Bruce Gardiner (Co-supervisor)

Prof. Brett Kirk (External supervisor)

3

TABLE OF CONTENTS

ABSTRACT 9 

PUBLICATIONS 11 

LIST OF AWARDS & PRESENTATIONS 12 

ABBREVIATIONS 14 

CHAPTER 1: INTRODUCTION 18 

1.1  Tendon Biology 18 1.1.1  Anatomy of human Achilles tendon 19 1.1.2  Histology of healthy tendon 20 1.1.3  Aetiology of tendon injuries 24 

1.1.3.1  Acute traumatic injury on Achilles tendon 24 1.1.3.2  Chronic Achilles tendinopathy 25 

1.1.4  Pathology of tendon injuries 30 1.1.4.1  Pathology of tendinopathy 30 1.1.4.2  Healing process 31 1.1.4.3  Molecules involved in tendon remodeling 34 

1.1.5  Treatments 43 1.1.5.1  Surgical treatment 43 1.1.5.2  Conservative treatment 44 

1.1.6  Summary 56 

1.2  Introduction to Bioreactor Design for Tendon/Ligament Engineering 56 1.2.1  Common key elements for tendon/ligament tissue engineering 57 1.2.2  Bioreactor design specific to tendon/ligament engineering 63 

1.2.2.1  Actuator and culture chamber design 66 1.2.2.2  Environmental control and medium circulation systems 71 1.2.2.3  Monitoring and feedback systems 75 

1.2.3  Commercial bioreactor systems for tendon/ligament engineering 79 1.2.4  Ideal bioreactors for tendon/ligament engineering 81 1.2.5  Previous work in this field 83 1.2.6  Conclusion 89 

1.3  Hypothesis and aims 90 1.3.1  Hypothesis 90 

1.3.1.1  Overall hypothesis 90 1.3.1.2  Specific hypothesis 91 

1.3.2  Aims 91 

CHAPTER 2 MATERIALS AND METHODS 94 

2.1  Material 94 2.1 94 2.1.1  Tissue Culture reagents 94 2.1.2  Chemical reagents 95 2.1.3  Molecular products 96 2.1.4  Oligonucleotide Primers 96 2.1.5  Antibodies 98 

4

2.1.6  Commercial kits 98 2.1.7  Other materials and equipment 99 2.1.8  Software 101 2.1.9  General solution 102 

2.2  Tissue culture procedures 104 2.1 104 2.2.1  Isolation of rabbit Achilles tendon 104 2.2.2  Culture of rabbit Achilles tendon 104 

2.3  RNA extraction 105 2.1 105 2.2 105 2.3.1  Extraction of total RNA 105 2.3.2  RNA concentration reading 106 

2.4  Quantitative polymerase chain reaction (QPCR) 107 2.4.1  Reverse transcription of mRNA 107 2.4.2  Quantitative polymerase chain reaction (QPCR) 108 

2.5  General histology 109 2.5.1  Tissue preparation 109 2.5.2  Haematoxylin and Eosin Staining 110 2.5.3  Histological scoring 112 

2.6  Type III collagen immunohistochemistry 113 

2.7  TUNEL assay 115 

2.8  Statistics analyses 116 

CHAPTER 3 (RESULTS): DESIGN AND CONSTRUCTION OF BIOREACTOR 118 

3.1  Design criteria 118 

3.2  Material selection 119 

3.3  Overall bioreactor design 120 3.3.1  The Actuating System 124 3.3.2  The Culture Chamber 125 3.3.3  Tissue Fixation 126 3.3.4  The Control System 127 

3.4  Bioreactor setup 128 

3.5  Discussion 130 

CHAPTER4 (RESULTS): PROGRAMMABLE MECHANICAL STIMULATION INFLUENCES TENDON HOMEOSTASIS IN A BIOREACTOR SYSTEM 133 

4.1  Abstract 133 

4.2  Introduction 134 

4.3  Material and Method 136 4.3.1  Tissue harvest and distribution 136 4.3.2  Histological preparation and assessment 138 

5

4.3.3  TUNEL assay 138 4.3.4  Immunostaining for type III collagen 138 4.3.5  Quantitative Real-time polymerase chain reaction (Q-PCR) 138 4.3.6  Statistical analyses 139 

4.4  Results 139 4.4.1  Histological examination of tendon undergoing different cyclic tensile strains 139 4.4.2  Impact of cyclic tensile strain on apoptosis of tenocytes 144 4.4.3  Type III collagen turnover in Achilles tendon under dynamic culture 146 4.4.4  Impact of cyclic tensile strain on ECM remodeling gene expression 147 

4.5  Discussion 148 

4.6  Conclusion 154 

CHAPTER 5 (RESULTS): CYCLIC MECHANICAL STIMULATION RESCUES RABBIT ACHILLES TENDON FROM DEGENERATION IN A BIOREACTOR SYSTEM 156 

5.1  Abstract 156 

5.2  Introduction 157 

5.3  Material and method 158 5.3.1  Tissue preparation and programmable mechanical stimulation 158 5.3.2  Histological preparation and assessment 161 5.3.3  TUNEL assay 161 5.3.4  Biomechanics testing 161 5.3.5  Real-time polymerase chain reaction 162 5.3.6  Immunohistochemistry 162 5.3.7  Statistical analyses 162 

5.4  Results 162 5.4.1  Cyclic mechanical stimulation improved the histological structure and cell viability of cultured tendon 162 5.4.2  Cyclic mechanical stimulation improved mechanical properties of cultured tendon 167 5.4.3  Rescue effect of cyclic mechanical stimulation on ECM remodeling gene expression 169 5.4.4  Cyclic mechanical stimulation decreased Type III collagen expression of cultured tendon 171 

5.5  Discussion 172 

5.6  Conclusion 178 

CHAPTER 6 GENERAL DISCUSSION AND FUTURE DIRECTIONS 180 

6.1  General discussion 180 6.1 181 6.1.1  The custom-made bioreactor (Chapter 3) 181 6.1.2  Mechanical stimulation in bioreactor system (Chapter 4 and 5) 183 6.1.3  Effect of various mechanical stimulation on tendon homeostasis (Chapter 4) 184 6.1.4  Rescue of degenerated tendon with mechanical stimulation. 187 

6.2  Limitation of this work 188 6.2 188 6.3 188 6.2.1  Chapter 3 188 6.2.2  Chapter4 189 6.2.3  Chapter 5 190 

6

6.3  Future directions 190 6.3.1  Gene profile of tenocyte subjected to 2D or 3D mechanical stimulation 190 6.3.2  The effect of mechanical stimulation on tendon nutrient infiltration 192 6.3.3  The effect of mechanical stimulation on stem cell tenogenic differentiation 193 6.3.4  Therapeutic guideline 194 6.3.5  Robot- driven orthosis 195 6.3.6  Tendon engineering 196 

6.4  Conclusion 197 

REFERENCE 198 

APPENDIX 225 

7

Dedicated to

My Awesome parents, Jihong Wang and Junhua Xu

and beautiful lovely wife, Xiang Jiang

for their love and support

8

ACKNOWLEDGEMENTS

The work described in this thesis was performed at the Center for Orthopaedic

Translational Research, School of Surgery, The University of Western Australia. I

would like to express my appreciation to my supervisor, Winthrop Professor

Minghao Zheng, co-supervisor Associate Professor Bruce Gardiner and External

supervisor Professor Brett Kirk. Without their support and guidance, this work

would not have been finished.

Sincere thanks to Dr. Zhen Lin and Ms. Euphemie Landao-Bassonga for their valuable

contributions, technical training and support.

I would love to thanks one of my best friend, Weiming Zeng, for the technical support

of bioreactor construction.

At last, I would also like to extend my appreciation to my colleagues and the

administrative officer for their assistance, friendship and encouragement.

9

ABSTRACT

Tendons are force-transmitting tissues connecting muscle to bone. Because of this

physiological function, biomechanics plays an essential role in maintaining tendon

homeostasis. Indeed tenocytes have the ability to sense and respond to different

mechanical loads by remodeling the tendon tissue. Studies reported that load

deprivation can cause tendinopathy-like morphology such as disorientated collagen

fibers and rounded cell nuclei, while mechanical overloading can result in tears and

rupture of the tendon. In tendon engineering studies, mechanical stimulation has been

shown to improve the cell viability, proliferation, and neo-tendon structure and

mechanical properties. Obviously, the biomechanical environment has various effects on

tendon tissue; and yet the specific effects of different loading conditions on tendon have

not been well elucidated. Mechanical loading is both a source of degradation (e.g.

through damage and indirectly through protease activity) and a stimulus for tissue

synthesis/repair. The hypothesis of this PhD project is that there is some optimal

mechanical load to maintain or restore functional tendon tissue in a bioreactor.

Firstly, in order to study the mechanical stimulation on tendon tissue, a functional

bioreactor is necessary. A uniaxial bioreactor was designed and fabricated. This

bioreactor system is able to provide pre-programmable mechanical simulation and

sterilized environment for a maximum 6 individual tendons simultaneously.

Secondly, rabbit Achilles tendon were cultured under different mechanical environments

including no loading, 3% strain, 6% strain and 9% strain at 0.25Hz for 8h/day for 6 days

10

in the bioreactor system developed in this study. Histological and gene assessment

showed that, of the strains applied, only at 6% strain was the rabbit model able to

maintain the tendon homeostasis, other loading regimes resulted in pathological changes

to the tendon as observed through histology testing.

Thirdly, due to the fact that loading deprivation of tendon can lead to progressive tendon

degeneration; static cultured rabbit Achilles tendons was used as a degradation model.

Application of 6% mechanical stimulation for 6 days followed by 6 days loading

deprivation culture significantly improved the morphology of degenerative tendon and

successfully restored the mechanical properties.

In summary, this study identified the effect of various loading intensities on tendon

homeostasis in bioreactor system, and proposed a hypothetical model that there is only a

narrow range of mechanical stimulation can maintain tendon homeostasis. We then

further estimated the therapeutic effect of mechanical stimulation on a degenerative

model, and found out accurate mechanical stimulation was able to reverse the

early-stage degradation of Achilles tendon.

This leads to the conclusion that proper mechanical stimulation may have therapeutic

benefits and clinical application.

11

Publications

Publications arising from this thesis

1. Wang, T., Gardiner, B. S., Lin, Z., Rubenson, J., Kirk, T. B., Wang, A., Xu, J., Smith,

D. W., Lloyd, D. G., Zheng, M. H., Bioreactor design for tendon/ligament

engineering. Tissue Eng Part B Rev 2013;19(2):133-46 (Appendix 1)

2. Wang, T., Lin, Z., Day, R. E., Gardiner B. S., Landao-Bassonga, E., Rubenson, J.,

Kirk, T. B., Smith, D. W., Lloyd, D. G.,Hardisty, G., Wang, A., Xu, J., Zheng, Q.

Zheng M. H. Programmable mechanical stimulation influences tendon homeostasis

in a bioreactor system. Biotechnol Bioeng 2013;110(5):1495-507 (Recommended

by Faculty 1000) (Appendix 2)

3. Wang, T., Lin, Z., Ni M., Day, R. E., Gardiner B. S., Landao-Bassonga, E.,

Rubenson, J., Kirk, T. B., Smith, D. W., Wang, A., Lloyd, D. G., Wang Y., Zheng Q.,

Zheng M. H., Cyclic mechanical stimulation rescues rabbit Achilles tendon from

degeneration in a bioreactor system. (manuscript submitted to Tissue engineering

Part A)

12

List of Awards & Presentations

Awards

1. “Peoples’ Choice Poster Award”, Combine Biology Science Meeting (CBSM), Perth,

Australia, 2012: Wang, T., Lin, Z., Landao-Bassonga, E., Zheng, M.H.,

Development and optimization of ex vivo bioreactor system for tendon tissue

engineering.

2. 2013 China Distinguished International student scholarship

Oral Presentations

1. Wang, T., Lin, Z., Landao-Bassonga, E., Zheng, M.H., Development and

optimization of ex vivo bioreactor system for tendon tissue engineering. In

Australian & New Zealand Orthopaedic Research Society 18th annual meeting.

Perth, Australia, 2012

2. Wang, T., Lin, Z., Landao-Bassonga, E., Zheng, M.H., Development and

optimization of ex vivo bioreactor system for tendon tissue engineering. In

Australian Society of Medical Research, 2012 annual meeting. Perth, Australia,

2012

3. Wang, T., Lin, Z., Day, R. E., Gardiner B. S., Rubenson, J., Kirk, T. B., Smith, D.

W., Lloyd, D. G., Wang, A., Zheng, M.H., The effect of programmable mechanical

stimulation on tendon repair in a bioreactor system. In Australian & New Zealand

Orthopaedic Research Society 19th annual meeting. Sydney, Australia, 2013

4. Wang, T., Lin, Z., Day, R. E., Gardiner B. S., Rubenson, J., Kirk, T. B., Smith, D.

W., Lloyd, D. G., Wang, A., Zheng, M.H., The effect of programmable mechanical

13

stimulation on tendon repair in a bioreactor system. In Combio 2013. Perth,

Australia, 2013

5. Wang, T., Lin, Z., Day, R. E., Gardiner B. S., Rubenson, J., Kirk, T. B., Smith, D.

W., Lloyd, D. G., Wang, A., Zheng, M.H., The effect of programmable mechanical

stimulation on tendon repair in a bioreactor system. In International Symposium on

Ligaments & Tendons. Arezzo, Italy, 2013

Poster Presentations

1. Wang, T., Lin, Z., Landao-Bassonga, E., Zheng, M.H.Minghao Zheng,

Development and optimization of ex vivo bioreactor system for tendon tissue

engineering. In Australian and New Zealand Bone and Mineral Society 22nd

annual meeting. Perth, Australia, 2012

2. Wang, T., Lin, Z., Day, R. E., Gardiner B. S.,Rubenson, J., Kirk, T. B., Smith, D. W.,

Lloyd, D. G., Wang, A.,Zheng, M.H., The effect of programmable mechanical

stimulation on tendon repair in a bioreactor system. In Australian and New Zealand

Bone and Mineral Society 23nd annual meeting. Perth, Australia, 2012

14

Abbreviations

ADAMTS A Disintegrin and Metalloproteinase with Thrombospondin Motifs

ADSCs Adipose-derived stem cells

AT Achilles tendon

ATT Autologous tenocyte therapy

bFGF Basic fibroblast growth factor

BMSCs Bone marrow-derived stem cells

BMP Bone morphogenetic proteins

COX-2 Prostaglandin-endoperoxide synthase 2

DMEM Dulbecco’s modified Eagle’s medium

ESCs Embryonic stem cells

ECM Extracellular matrix

GDF-5 Growth differentiation factor 5

GAG Glycosaminoglycan

H&E Haematoxylin-eosin

H2O2 Hydrogen peroxide

LE Lateral epicondylitis

IGF-I Insulin-like growth factor-I

15

IL-6 Interleukin-6

LVDT Linear variable differential transformers

MCP1 Monocyte chemotactic protein 1

MIF Macrophage inhibitory factor

MMP Matrix metalloproteinase

MSCs Mesenchymal stem cells

NSAID Non-steroid anti-inflammation drug

OCT Optical coherence tomography

PBS Phosphate buffered saline

PDGF Platelet-derived growth factor

PRP Platelet-rich plasma

PMMA Polymethylemethacrylate

PMS Programmable mechanical stimulation

PGE2 Prostaglandin E2

Scx Scleraxis

SDSCs Synovium-derived stem cells

SMBS Step motor-ball screw transmission system

SP Substance P

16

TDSCs Tendon-derived stem cells

TGF- Transforming growth factor

VEGF Vascular endothelial growth factor

17

Chapter 1

Introduction

18

CHAPTER 1: INTRODUCTION

1.1 Tendon Biology

The Achilles tendon (AT), also known as the calcaneal tendon, is the strongest and

thickest tendon in human body. It connects the gastrocnemius and soleus muscles to

calcaneus bone. As a force transferring tissue, AT serves the function of transferring the

power generated by the muscle to the calcaneus, allowing plantar flexion about the ankle

joint. It also helps store elastic energy while walking/running, increasing locomotion

efficiency (Bressel and McNair 2001).

AT injury is a serious and common injury. AT rupture was first reported in 1575 by

Ambrose Paré. The AT injuries rate is estimated at between 2 to 12 in 100,000 people

(So and Pollard 1997; Schepsis, Jones et al. 2002; Suchak, Bostick et al. 2005).

Typically, this injury affects males between 30 to 50 years old (Gebauer, Beil et al. 2007).

It is thought that due to the specific geometry of the AT, the mechanical weak point is 2

to 6 cm from the calcaneal insertion, that is the thinnest part of the AT and the common

rupture site(Wren, Lindsey et al. 2003). In addition, 11% of regular runners suffer from

Achilles tendinopathy(Rees, Wilson et al. 2006), which can cause significant pain and

restricts the activities in daily living, and potentially tendon rupture (Leppilahti and

Orava 1998; Smith 2000). Traditionally, overuse and/or overloading has been

considered to be the cause of tendinopathy, 30% to 50% of sports injuries were reported

to be an overuse injury, and tendon injuries comprise a large part of that(Schechtman

and Bader 1997; Khan and Cook 2003).

19

In the United States, injuries to tendons and ligaments represent about half of the 33

million musculoskeletal injuries(Huang, Qureshi et al. 2000). Each year more than

33,000 tendon reconstructions occur in U.S., costing $30 billion USD(Butler, Gooch et

al. 2010; Chen, Yin et al. 2010). In Australia, $250 million is spent annually just on

rotator cuff repair(Chen, Xu et al. 2009). However, despite the high prevalence of

tendon injury and associated tendinopathy worldwide, treatment options remain poorly

defined.

1.1.1 Anatomy of human Achilles tendon

The Achilles tendon is the extension of two muscles, the gastrocnemius and soleus

muscles (Suchak, Bostick et al. 2005). Depending on knee positioning, these separate

muscles have their own movement (Schepsis, Jones et al. 2002), but mainly control the

plantar flexion of the ankle, allowing for locomotion at the ankle (Suchak, Bostick et al.

2005). These muscles form together at the mid-calf region of the lower leg, where the

AT begins. The AT then merges into a single tendon at around 5 to 6 cm from the

calcaneal insertion (So and Pollard 1997). The AT is covered by a peritenon, a

single-cell layer tissue, which act as a lubricant during dynamic activity and provides

vascular supply to the tendon tissue (Schepsis, Jones et al. 2002). Beside the peritenon,

the main nutrient supply is from the muscle-tendon junction and bone tendon junction,

therefore, the nutrient delivery to the AT is very limited (Schepsis, Jones et al. 2002).

20

Figure 1.1 Anatomy of human Achilles tendon (Figure from: http://www.webmd.com/a-to-z-guides/achilles-tendon)

1.1.2 Histology of healthy tendon

Tendons are relatively hypovascular and hypocellular tissues. Tendon cells, primarily

tenocytes and fibroblasts, comprise less than 5% of the total volume. The morphology

of tenocytes and fibroblasts are “sharp” and usually elongated along collagen fibers in

normal tissues, but some rounded tenocytes are found occasionally(Manske 1988; Bray,

Rangayyan et al. 1996; Lo, Ou et al. 2002; Hildebrand, Frank et al. 2004). Tendons

consist of collagens, cells, proteoglycans, elastin, glycolipids and water. Although

roughly 65-70% of the total weight is water(Lin, Cardenas et al. 2004), tendon is a

highly organized hierarchical structure. Collagen type I is the main structural/functional

component, and comprises around 70-80% of the dry weight, and is generally aligned

along the long axis of tendon (Buckwalter and Hunziker 1996; Jarvinen, Jarvinen et al.

2004). Type III collagen comprises less than 5% of total collagen, and is mainly present

21

in the endotenon and epitenon, but is also present in the early phase of tendon repair

(Buckwalter and Hunziker 1996; Jarvinen, Jarvinen et al. 2004). The hierarchical

structure of tendon is shown in Figure 1.2, which is modified from the study of Kastelic

et al. (Kastelic, Galeski et al. 1978).

Type 1 collagen molecules synthesized by tenocytes bind together into a triple helix

tropocollagen (Kuhn 1969), which then self-assemble into microfibril and bind to

adjacent helixes by molecules such as decorin and biglycan, so forming collagen

fibrils(Tkocz and Kuhn 1969). This helixes structure is able to provide high resistance

to tensile strain while maintaining its flexibility (Screen, Lee et al. 2004). Due to this

structural characteristic, the tendon is able to efficiently transfer force from muscle to

bone (Magnusson, Hansen et al. 2003). The collagen fibrils align parallel to each other

assemble into collagen fiber with a waveform structure, enabling the tendon to absorb

the sudden initial force generate by muscle(Kastelic, Palley et al. 1980). As shown as

Figure 1.3, when unloaded the collagen fiber appears ‘wavy’. The wavy structure

disappears at around 4% strain, micro and partial rupture starts at ~8% strain, beyond

that, complete rupture is likely to occur (Maffulli 1998).

Tenocytes adhere to the collagen fiber; and sense the mechanical loading on the tendon

(Benjamin and Ralphs 1998; Murata, Nishizono et al. 2000). In the haematoxylin-eosin

(H&E) stained normal and healthy tendon section, tenocytes are evenly distributed

between the straight and parallel collagen fibrils with sharp and elongate nuclei (Chen,

22

Willers et al. 2007).

Although proteoglycans comprised only ~1% dry mass of tendon, they still play an

essential role in tendon biomechanics(Thompson 2013). Small proteoglycans such as

decorin control collagen fibril assembly and alignment(Parkinson, Samiric et al. 2011),

and the hydrophilic nature of the large proteoglycans has a strong effect on tissue

mechanical properties (Screen, Chhaya et al. 2006). Moreover, proteoglycan

concentration varies within and between tendon due to different local loading

characteristics(Birch 2007).

Elastin and microfibrils form elastic fibers, which comprise around 0.1% to 2% dry

mass of tendon(Kannus 2000). They plays a role in the tissue low strain and resilience

propertie by bridging between collagen fiber bundles with a network structure(Smith,

Vaughan-Thomas et al. 2011).

23

Figure 1.2 The structure of a healthy tendon. figure modified from Kastrlic et al. (Kastelic, Galeski et al. 1978)

Figure 1.3 Force versus strain curve of tendon (figure modified from the study of Wang et al.) (Wang 2006)

24

1.1.3 Aetiology of tendon injuries

Tendon injuries can be classified into acute and chronic. In acute traumatic injury

extrinsic factors predominate, whereas in chronic tendinopathy both intrinsic and

extrinsic factors interact (Williams 1993; Khan and Maffulli 1998).

1.1.3.1 Acute traumatic injury on Achilles tendon

Traumatic AT injuries are often associated with sport activities, although AT is the

strongest tendon in human body, sudden, excessive strain of AT from athletic or

recreational activities can cause tears or even rupture the AT (Kannus and Natri 1997).

Statistics show that 60-75% of AT ruptures happen during sport activities, especially in

soccer, badminton and basketball (Kannus and Natri 1997). Most of the AT ruptures are

associated with sudden acceleration and jumping, that is those activities which require

massive sudden force generation from muscle, and the rupture site tend to locate at 2 to

6 cm above the calcaneus-Achilles tendon junction. Interestingly the left tendon is more

commonly affected than the right tendon (Hattrup and Johnson 1985; Leppilahti and

Orava 1998).

In acute traumatic injuries, AT failure is mainly due to excessive strain caused by

sudden explosive power produced by the triceps surae (Langberg, Skovgaard et al.

1999). Hoffmeyer et al. found some pathological changes including increased lipid

droplet and ultrastructural muscle changes indicating partial ischemia in the triceps

25

surae muscle after AT rupture, and this might contribute to the abnormal stiffness in the

triceps muscle causing excessive muscle contraction. (Hoffmeyer, Freuler et al. 1990)

1.1.3.2 Chronic Achilles tendinopathy

Tendinopathy of the AT is a chronic non-inflammatory, degenerative condition, which

mainly affect male runners between 35 and 45 years old(Alfredson and Lorentzon 2000).

Tendinopathy is normally characterized by matrix disorganization, hypercellularity and

vascular hyperplasia (Teitz, Garrett et al. 1997). In some cases heterotopic

mineralization was considered as a feature of tendinopathy as well (Jarvinen, Jozsa et al.

1997). The clinical symptom of tendinopathy in the AT is pain, which generally happens

at the beginning and the end of training (Maffulli, Sharma et al. 2004). As the

pathological process progresses, it can affect daily activities. Approximately 3-10% of

chronic Achilles tendinopathy develop to AT ruptures (Khan, Cook et al. 1999).

Traditionally, overuse is considered as an extrinsic factor of Achilles tendinopathy.

Therefore, tendinopathy mainly affects runner, dancers and gymnasts, who constantly

and repetitively overload their tendon (Teitz, Garrett et al. 1997). During slow walking,

human AT is subjected to ~3kN force, and it can reach up to 9kN at a speed of 6m/s,

which corresponds to 12.5 times body weight (Komi, Fukashiro et al. 1992). Moreover,

the peak stress of AT is more than twice that of other tendons in human body(Wren,

Yerby et al. 2001). Although tendinopathy affects such a large population, the exact

aetiology remains uncertain beyond the catch-all ‘overuse’. There are several theories

26

proposed, including vascular supply, hypoxia, hyperthermia, apoptosis and abnormal

biomechanical environment.

1.1.3.2.1 Vascular supply

When tendon is subjected to mechanical loading, microruptures or damage of collagen

fibrils start to accumulate, and need to be repaired by tenocytes. However, the vascular

supply of the “high risk, high stress zone” (2-6 cm above the calcaneus insertion) is

relatively poor compared to the other parts of the tendon, presumably due to the hostile

mechanical environment (Hattrup and Johnson 1985; Carr and Norris 1989; Leppilahti

and Orava 1998). Under the situation of repetitive damage without enough blood supply,

the repair mechanism may be unable to maintain the tendon’s structural integrity, and

then the AT undergoes degeneration(Kannus and Jozsa 1991).

1.1.3.2.2 Hypoxia

The oxygen supply of tendon and ligament is only 13% that of skeletal muscles(Kannus

and Jozsa 1991). The degree and duration of the hypoxia plays a key role in cellular

viability(Nathan 2002). Several studies suggest that a hypoxic environment can induce

inflammatory cytokines, including interleukin-6 (IL-6), interleukin-8 (IL8), monocyte

chemotactic protein 1 (MCP1) and Platelet-derived growth factor (PDGF), and so might

significantly disrupt the balance between reparative and degenerative processes in the

tendon (Berse, Hunt et al. 1999; Zamara, Galastri et al. 2007; Millar, Reilly et al. 2012).

27

Moreover, hypoxia has been shown to increase the total collagen production but with a

‘shift’ in production to type III collagen instead of type I, thus disturbing the original

collagen composition ratio (Millar, Reilly et al. 2012). Lastly, when tendon repair

mechanisms are activated, tendon requires oxidative energy metabolism to maintain

cellular ATP levels. That is local tissue hypoxia may result in reduced repair capacity

and/or tenocyte death (Birch, Rutter et al. 1997).

1.1.3.2.3 Hyperthermia

During locomotion, 5% to 10% of energy stored in tendons is converted into heat (Ker

1981; Riemersma and Schamhardt 1985). Due to the hypovascular structure of tendon,

the blood supply to the structure is not sufficient to dissipate the heat generated. The

temperature of equine superficial digital flexor tendon has been recorded to reach 45℃

during galloping(Wilson and Goodship 1994). The in vitro study of Birch et al. indicates

that tenocytes appear to have high thermal tolerance; even under 45℃ culture condition

the viability of tenocyte is still unaffected. However, compromised cellular function

might be induced by repeated exposure to short periods of hyperthermia, as it has been

shown to be associated with reduced collagen synthesis and disturbed cell

metabolism(Arancia, Crateri Trovalusci et al. 1989; Birch, Wilson et al. 1997).

28

1.1.3.2.4 Apoptosis

Significant tenocyte apoptosis has been found in tendinopathy in various tendons,

including rotator cuff and AT(Yuan, Wang et al. 2003; Chen, Wang et al. 2010; Wang,

Lin et al. 2013). Deformation of the cytoskeleton of tenocytes can produce

stress-activated protein kinase that triggers apoptosis (Arnoczky, Tian et al. 2002;

Skutek, van Griensven et al. 2003). In ruptured tendon, apoptotic cell number is

elevated compared to normal tendon (Chen, Willers et al. 2007). For example,

quadriceps femoris tendons with tendinopathy exhibited a 1.6 times higher apoptosis

rate than normal tendon (Machner, Baier et al. 2003).

1.1.3.2.5 Biomechanical environment

As a force transferring tissue, it is unsurprising that the biomechanical environment is

essential to tendon homeostasis; overloading and underloading can both lead to tendon

abnormality. Repetitive and excessive loading may result in release of cytokines by

tenocytes leading to abnormal cellular activities(Leadbetter 1992). The release of

cytokines may induce the expression of MMPs, which can degrade the extracellular

matrix and eventually causes tendinopathy (Chen, Yu et al. 2011). However, loading

deprivation of tenocyte can cause direct upregulation of MMP-1(Choi, Kondo et al.

2002). It has been reported that mechanical stimulation is able to reduce MMP-1

expression (Lavagnino, Arnoczky et al. 2003). Due to the different protocols that have

been adopted in tendon studies on the effect of mechanical stimulation on tendon,

simple comparison is often difficult.

29

Compared to the tendon-muscle junction and the mid-tendon, the biomechanical

environment of the bone-tendon junction is more complicated. The attachment of

tendon is through fibrocartilage to mineralized fibrocartilage to bone in relatively a

short distance (<2mm), which is called as ‘the enthesis organ’ (Benjamin, Moriggl et al.

2004). Between the tendon and the bone is a bursa, and fibrocartilage is expressed on

the opposing bone and tendon surface to absorb the compression of the tendon against

the bone, as shown in Figure 1.4 (Cook and Purdam 2012). In most of the clinical cases,

the abnormal imaging findings and the reports of tendinopathy on tendon-bone insertion

are more likely at the site of compression proximal to the tendon insertion, which

indicate that abnormal compression might be able to trigger tendinopathy (Ohberg and

Alfredson 2003; Kong, Van der Vliet et al. 2007). The study of Grigg et al. suggests that

excessive loading on the tendon will cause the loss of bound water normally presented

in the transitional zones (Grigg, Wearing et al. 2009), which further increases the loads

carried by tenocytes located in these regions. To respond to the excessive compression,

tenocyte start to synthesis large water binding proteoglycans to reduce the permeability,

thereby protecting against further insult. Further loading on this swollen region may

aggravate this situation and potentially result in tendinopathy (Hamilton and Purdam

2004; Parkinson, Samiric et al. 2010).

30

Figure 1.4 Normal tendon enthesis, figure taken from Cook et al. (Cook and Purdam 2012)

1.1.4 Pathology of tendon injuries

1.1.4.1 Pathology of tendinopathy

“Tendinitis” has been used for definition of tendon degeneration in tendinopathy for

over 20 years (Puddu, Ippolito et al. 1976). Although the fundamental problem of

tendinopathy is collagen degeneration instead of inflammatory, many clinicians still

use the term of “tendinitis”. Most of the scientists nowadays have advocated the term

of “tendinopathy” for description of the clinical degeneration condition in and around

tendons caused by overuse, while the term “tendinitis” is only used after

histopathological examination (Maffulli, Khan et al. 1998).

31

Repetitive overloading on tendon was commonly considered as a cause of

tendinopathy. Due to repetitive loading during daily activities, Achilles tendon is one

of the most common tendons affected by tendinopathy. Histologically, tendinopathy

is mainly characterized by collagen fiber disorientation and thinning, scattered

vascular ingrowth, cell rounding, changes of cell density and increased expression of

type III collagen, in some cases, glycosaminoglycan (GAG) accumulation, lipid

droplets accumulation, ossification and increased cell apoptosis (Leadbetter 1992;

Khan and Maffulli 1998).

Various types of degeneration can be found in different tendons, however, in AT,

mucoid and lipoid are the most common types. In mucoid degeneration, vacuoles and

mucoid patches characterized by accumulation of proteoglycan/GAG are found

between the collagen fibers(Maffulli, Sharma et al. 2004). However, in lipoid

degeneration, the collagen structure of tendon is disrupted by the abnormal

accumulation of lipid(Maffulli, Sharma et al. 2004). An abnormally high expression

of type III collagen is found in degenerated tendon (Riley, Harrall et al. 1994).

1.1.4.2 Healing process

Tendon injuries can be divided into two categories, traumatic injury and chronic

tendinopathy. The tissue response to these two types of injuries are not quite the same

(Leadbetter 1992). The healing process of tendon has been studied in both human and

animals; however, most are concerned with traumatic injuries healing, whilst the

32

(inadequate) healing response in degenerative tendon is still poorly understood.

1.1.4.2.1 Acute traumatic injury healing

Tendon healing after acute traumatic injury requires reestablishment of the

connection between collagen fibers and the gliding mechanism between tendon and

its neighboring structure (Schneewind, Kline et al. 1964; Abrahamsson and

Gelberman 1994). Formation of scar tissue provides initial repair at the injury site

(Dunphy 1967). However, a lack of mechanical stimulation is thought to lead to an

excess of scar tissue, adhesions, and so compromises the normal tendon function.

Therefore, the rehabilitation procedure after initial immobilization of the injury site is

critical, as the mechanical loading is able to guide organized collagen fiber formation,

decrease the formation of postoperative adhesions and increase tendon

strength(James, Kesturu et al. 2008).

After acute tendon injury, the body initiates the healing process, which includes three

overlapping stages including: (1) acute inflammation; (2) proliferation and (3)

remodeling (Goodship, Birch et al. 1994).

The acute inflammation phase begins right after the injury and lasts 1 to 2 week

depending on the severity of the injury. Clinically, the inflammation phase is

characterized by well-known signs of inflammation such as heat, pain and swelling

33

(Goodship, Birch et al. 1994). Histologically, injuries on the tendon cause the

formation of hematomas in the tendon sheath, which release various chemotactic

factors and pro-inflammatory molecules. These molecules attract inflammatory

cells such as neutrophils, monocytes and macrophages to migrate from surrounding

tissue to the wound site where the cellular debris and foreign body matter are

engulfed and resorbed by phagocytosis(Maffulli, Sharma et al. 2004). Meanwhile,

tenocytes are recruited to the site and start to synthesize various extracellular matrix

components and reestablish vascular networks (Lindsay and Birch 1964; Myers and

Wolf 1974).

During the proliferation stage, tenocytes continue to be recruited and proliferate to

accelerate the repair process at the wound site. However, the newly synthesized ECM

is mainly a disorganized type III collagen(Garner, McDonald et al. 1989). An

extensive vascular network is formed and the wound exhibits scar-like tissue.

During the remodeling stage, 6-8 weeks after injury, cellular proliferation, matrix

synthesis, type III collagen expression start to decrease, and collagen type I synthesis

increases to replace the type III collagen (Liu, Yang et al. 1995). Type I collagen

fibers provide the long-term mechanical support of the regenerated tissue. However,

it is uncertain as to whether the repaired tissue achieves the strength of the original

tissue(James, Kesturu et al. 2008).

34

1.1.4.2.2 Chronic healing

Chronic healing processes in chronic tendinopathy are different from the acute

healing case described above. In normal tendon, type I collagen, the predominant

collagen, is well organized along the axis of the tendon. However, in tendinopathy,

the collagen structure is disorganized and type III collagen is highly expressed. In the

normal healing process, type III collagen is synthesized by tenocytes as a temporary

“band aid”(Maffulli, Khan et al. 1998). During the remodeling stage, type III

collagen is replaced by collagen type I which is more resistant to mechanical loading.

However, the repetitive and chronic damage in chronic tendinopathy keeps the high

expression of collagen type III in tendon without shifting to collagen type I (Riley,

Harrall et al. 1994). Due to the failure to complete this final remodeling stage, the

tendon is gradually weakened; and eventually leads to rupture even at ‘low’ daily

activities(Hamada, Okawara et al. 1994).

1.1.4.3 Molecules involved in tendon remodeling

1.1.4.3.1 Inflammatory mediators in tendon degeneration

Although tendinopathy is considered to be a degenerative process, rather than

inflammation, inflammatory mediators still play an important role. Interleukin (IL)-6

and IL-1 have been identified as two of the most important inflammatory mediators

and their functions are similar (Chen, Yu et al. 2011). IL-6 and IL-1 are both able to

induce Prostaglandin-endoperoxide synthase 2 (COX2), which stimulates the expression

of PGE2 and the acute phase of an inflammatory response in tendon (Chen, Yu et al.

35

2011). The inflammatory response alters the cell homeostasis causing apoptosis of

tenocytes which initiates the pathogenesis of tendon. IL-6 and IL-1have also been

shown to stimulate MMP1 and 3 expression (Archambault, Tsuzaki et al. 2002; Chen,

Yu et al. 2011), which can degrade collagen type I and collagen–associated small

proteoglycans respectively. Therefore, the upregulation of these inflammatory mediators

can cause matrix degradation.

The induction of IL-6 and IL-1 can be divided into chemical and mechanical

stimulation. Macrophage inhibitory factor (MIF) and Substance P (SP) have been

reported to upregulate IL-6 and IL-1 expression in tenocytes (Hart, Archambault et al.

1998; Nguyen, Lue et al. 2003; Morand, Leech et al. 2006). Tenocytes subjected to

cyclic strain increase the production of IL-6 and IL-1 (Skutek, van Griensven et al.

2001; Archambault, Tsuzaki et al. 2002; Tsuzaki, Bynum et al. 2003). The regulatory

network between inflammatory cytokines and tenocyte apoptosis remains unclear. It is

possible that in a routine process, normal loading induces cytokines in tissues that

trigger apoptotic cell death to remove the damaged cells. However, over-loading may

cause excessive production of cytokines and lead to excessive apoptosis and ultimate

tissue degeneration (Millar, Wei et al. 2009).

1.1.4.3.2 Enzyme in tendon degeneration

There are two kinds of metalloproteinase involved in tendon degeneration, which are

Matrix metalloproteinase (MMPs) and A Disintegrin and Metalloproteinase with

36

Thrombospondin Motifs (ADAMTS).

MMPs are zinc-dependent endopeptidases that are able to degrade all the components of

the extracellular matrix. MMPs can be classified into four main groups: collagenase,

gelatinases, stromelysins and membranes type MMPs(Bramono, Richmond et al. 2004).

Tendon degradation is initiated by MMPs (Riley, Curry et al. 2002). ADAMTS family,

which are also known as “aggrecanases”, are able to degrade proteoglycans, however,

the precise ADAMTSs that are involved in the degradation of tendon proteoglycans

remains unclear.

The regulation network of metalloproteinases is very complicated, but includes

transcription, activation and inhibition by tissue inhibitors of metalloproteinase (TIMPs)

(Nagase, Visse et al. 2006). There are 23 MMPs and 19 ADAMTS in humans, almost all

of which can be detected in Achilles tendon, although the expression levels vary widely.

These enzymes are essential regulators in cellular activities, matrix remodeling, and

pathologic processes (McCawley and Matrisian 2001). The main members of the MMP

family involved in tendon remodeling are summarized in Table 1.1(Magra and Maffulli

2005). In injured tendon, including acute tears and chronic tendinopathy, increased

expression of MMP1, 2, downregulation of MMP3 and TIMP2,3,4 are found(Ireland,

Harrall et al. 2001; Choi, Kondo et al. 2002; Riley, Curry et al. 2002; Lo, Marchuk et al.

2004). However, TIMP1, which is upregulated in acute tendon tears, is inhibited in

tendinopathy (Ireland, Harrall et al. 2001; Choi, Kondo et al. 2002; Lo, Marchuk et al.

37

2004). The balance of MMPs and TIMPs is important to tendon homeostasis. Over

expression of MMPs will cause the pathogenesis of tendinopathy(Riley, Curry et al.

2002; Lo, Marchuk et al. 2004; Jones, Corps et al. 2006).

38

Tabl

e 1.

1 T

he f

unct

ion

and

targ

ets

of th

e m

atri

x m

etal

lopr

otea

se (

MM

P)

fam

ily

Nam

eS

ynon

ym

Deg

rade

s O

ther

fun

ctio

nsR

efer

ence

MM

P1

Col

lage

nase

-1, I

nter

stiti

al c

olla

gena

se,

Fi

brob

last

col

lage

nase

C

olla

gen

type

I, I

I, I

II, V

II, V

III,

and

X

(G

oupi

lle,

Ja

yson

et

al

. 19

98;

Bra

mon

o,

Ric

hmon

d et

al

. 20

04;

Mag

ra

and

Maf

full

i 200

5)

MM

P2

72 k

Da

gela

tina

se A

ty

pe I

V g

elat

inas

e C

olla

gens

typ

e IV

, V, V

II, X

, and

XI,

Fib

rone

ctin

,ela

stin

, pro

teog

lyca

ns,

gela

tin

Syne

rgis

tic

wit

h M

MP

1

MM

P3

Str

omel

ysin

-1,T

rans

in

Col

lage

ns I

II, I

V, V

, and

IX

,pro

teog

lyca

ns, l

amin

in, f

ibro

nect

in, g

elat

in

Bro

adsu

bstr

ate

spec

ific

ity,

Act

ivat

es p

ro-M

MP

s

MM

P7

Pum

p-1

Gel

atin

, pro

teog

lyca

ns, f

ibro

nect

in, e

last

in, c

asei

n A

ctiv

ates

pro

-MM

P1

MM

P8

Neu

trop

hil c

olla

gena

se

Col

lage

ns ty

pe I

, II,

and

III

, agg

reca

n

MM

P9

92 k

Da

gela

tina

se-B

C

olla

gens

type

IV

, V, X

, XI,

gela

tin

MM

P10

S

trom

elys

in-2

,Tra

nsin

-2

Col

lage

ns ty

pe I

II, I

V, a

nd V

,gel

atin

, fib

rone

ctin

, A

ctiv

ates

pro

-MM

Ps

MM

P11

Stro

mel

ysin

-3

Agg

reca

n, f

ibro

nect

in, l

amin

in

MM

P12

M

acro

phag

e m

etal

loel

asta

se

C

olla

gen

type

s I

and

IV, a

ggre

can,

fib

rone

ctin

, lam

inin

, ent

acti

n, g

elat

in

type

I, v

itro

nect

in, f

ibri

llin,

e la

stin

, c

MM

P13

C

olla

gena

se-3

C

olla

gens

type

I, I

I an

d II

I, g

elat

in

39

1.1.4.3.3 Growth factor in tendon repair

PDGF- is another important cytokine in tendon and ligament repair. It is able to

promote chemotaxis, cell proliferation, ECM production, surface integrin expression,

and revascularization in tendon and ligament (Nakamura, Shino et al. 1998;

Nakamura, Timmermann et al. 1998; Harwood, Goomer et al. 1999). Treatment of

PDGF- in a small animal injury model of tendon significantly improved the enthesis

structure and biomechanical properties (Hildebrand, Woo et al. 1998; Chan, Fu et al.

2006). In a sheep rotator cuff injury model, scaffolds infused with PDGF-achieved

a better histologic score and mechanical properties than an empty scaffold group (Hee,

Dines et al. 2011).

The Transforming growth factor- (TGF-family is essential for tendon

development. It has been shown to stimulate collagen synthesis, cell proliferation and

migration. Furthermore, it is a key growth factor that modulates the scar tissue

following a wound (Chang, Most et al. 1997; Kashiwagi, Mochizuki et al. 2004;

Galatz, Rothermich et al. 2007; Kim, Galatz et al. 2011; Kovacevic, Fox et al. 2011).

In tendon healing, TGF-expression elevates during the inflammatory phase and

promotes collagen production and cell proliferation(Sporn, Roberts et al. 1986;

Kannus and Jozsa 1991), while TGF-is able to reduce scar tissue formation during

tendon repair(Galatz, Rothermich et al. 2007; Kim, Galatz et al. 2011; Kovacevic,

Fox et al. 2011). In a rat rotator cuff injury model, the tendon-bone junction displayed

better mechanical properties in the group supplemented with TGF-compared to

40

the TGF-group(Kim, Kang et al. 2007; Kim, Galatz et al. 2011; Manning, Kim et

al. 2011).

The Bone morphogenetic proteins (BMP) family is part of the TGF-superfamily.

Some BMPs are able to stimulate bone and cartilage formation and induce

mesenchymal stem cell to differentiate into the cartilage or bone linage(Ducy and

Karsenty 2000). However, instead of osteogenic and chondrogenic induction,

BMP-12 is able to induce tendon formation, and is important for tendon healing (Lou,

Tu et al. 2001). Treatment of BMP-12 in a tendon injury model showed a better

organized tendon tissue, higher volumes of collagen type I and improved mechanical

properties(Forslund, Rueger et al. 2003; Majewski, Betz et al. 2008)

The insulin-like growth factor IGF-1 expression increases during the initial

inflammatory phase of tendon healing and is able to stimulate the migration and

proliferation of fibroblasts and inflammatory cells to the wound site. Several small

animal studies have suggested that IGF-1 has the ability to accelerate the tendon

healing by promoting cell proliferation and ECM production (Abrahamsson 1991;

Abrahamsson, Lundborg et al. 1991; Dahlgren, Nixon et al. 2001; Dahlgren, van der

Meulen et al. 2002).

VEGF is not a common cytokine in human tendon, but is expressed in the areas with

41

high microvascular density during tendon healing (Petersen, Pufe et al. 2003). VEGF

treatment has been shown to increase the vessel formation from 1 to 8 weeks after

tendon rupture (Hou, Mao et al. 2009); however, the mechanical properties

improvement was not significant except for in the first week (Zhang, Liu et al. 2003).

bFGF is secreted by tenocytes and inflammatory cells at the injury site of tendon. It

can stimulate cell proliferation, cell migration, collagen synthesis and angiogenesis

(Chan, Chan et al. 1997; Chan, Fu et al. 2000; Thomopoulos, Harwood et al. 2005;

Tang, Cao et al. 2008). Increased cellularity associated with reduced collagen type I

and III expression in injury tendon was observed with additional bFGF in the early

healing process(Chan, Fu et al. 2008; Thomopoulos, Das et al. 2010), while collagen

synthesis levels increased after two weeks treatment, suggesting a prolonged phase of

proliferation can potentially lead to better tendon repair(Sahoo, Toh et al. 2010).

42

Table 1. 2 Summary of cytokines involved in tendon healing and the main function

Name Function Reference

PDGF- Promote chemotaxis, cellproliferation, extracellular matrix production, surfaceintegrin expression, and revascularization

(Nakamura, Shino et al. 1998; Nakamura, Timmermann et al. 1998; Harwood, Goomer et al. 1999)

TGF- Stimulate collagen synthesis, cell proliferation and migration

Modulate scar tissue after wound, ligament, and tendon-to-bone healing.

(Chang, Most et al. 1997; Kashiwagi, Mochizuki et al. 2004; Galatz, Rothermich et al. 2007; Kim, Galatz et al. 2011; Kovacevic, Fox et al. 2011)

BMP-12 Induces tendon and ligament tissue formation (Fu, Wong et al. 2003)

IGF-1 Stimulate the migration and proliferation of fibroblasts and inflammatory cells to the wound site

Increasing cellular proliferation

Enhance matrix synthesis, improving tendon mechanical properties

(Abrahamsson 1991; Abrahamsson, Lundborg et al. 1991; Dahlgren, Nixon et al. 2001; Dahlgren, van der Meulen et al. 2002)

VEGF Essential for initial vascular plexus formation and granulation tissue development

Stimulate vascular bud formation and endotheliocyte migration during neo-vascularization

(Molloy, Wang et al. 2003)

bFGF Promote cell proliferation, cell migration, collagen production, and angiogenesis

(Chan, Chan et al. 1997; Chan, Fu et al. 2000; Takahasih, Nakajima et al. 2002; Thomopoulos, Harwood et al. 2005; Tang, Cao et al. 2008)

43

1.1.5 Treatments

Although our understanding of tendon biology and the healing mechanism has

improved tremendously recently, and various treatments have been adopted for

tendon injury, the outcomes remain unsatisfactory. The goal of treatment for Achilles

tendon diseases is to reduce pain, promote healing and restore its functional

mechanical properties. Surgical and so-called conservative treatments (e.g.

physiotherapy) are the most common procedures. However in recent years

achievements in tissue engineering in animal studies has begun to offer hope for an

alternative solution to tendon injury.

1.1.5.1 Surgical treatment

1.1.5.1.1 Chronic tendinopathy

Surgery is often considered as a last choice for treating tendinopathy. Patients who fail

to improve after a period of conservative treatments tend to be subjected to surgical

treatment, but the results are not uniformly satisfactory. Various surgical procedures

have been described in different studies. They can be divided into four categories: open

tenotomy with abnormal tissue removal, paratenon stripped (Paavola, Orava et al. 2000);

open tenotomy with abnormal tissue removal, paratenon not stripped(Leach, Schepsis et

al. 1992); open tenotomy with longitudinal tenotomy(Rolf and Movin 1997); and

percutaneous longitudinal tenotomy(Maffulli, Testa et al. 1997). In the study of Tallon

et al., of the1648 cases of surgically treated Achilles tendinopathies the success rate was

77.4% (Tallon, Coleman et al. 2001)

44

1.1.5.1.1 Acute rupture

In acute traumatic injuries such as Achilles tendon rupture, early surgery is preferred.

Reconstructive surgery is the common option for athletes, young people, and patients

with chronic ruptures (Christensen 1953; Krueger-Franke, Siebert et al. 1995).

However there is still no agreed protocol for the management of ruptured Achilles

tendons(Maffulli 1995; Nyyssonen and Luthje 2000) or which operative technique

gives the best outcomes (Bugg and Boyd 1968; Hogsaa, Nohr et al. 1990). Ruptured

Achilles tendons can be classified as open operative, percutaneous operative and

non-operative based on the modalities of management. In open surgery, some studies

reported high complication and skin healing problems (Gillespie and George 1969;

Nistor 1981; Bomler and Sturup 1989), while others show little or a low rerupture

rate(Goldman, Linscheid et al. 1969; Inglis and Sculco 1981; Cetti and Christensen

1983; Beskin, Sanders et al. 1987; Zell and Santoro 2000). Percutaneous repair is

another technique that is able to minimizes skin-healing problems under local

anesthetic(Ma and Griffith 1977). However, compared to open repair, percutaneously

repaired Achilles tendons had higher rerupture rates and thinner diameter(Bradley and

Tibone 1990; Cretnik, Kosanovic et al. 2005).

1.1.5.2 Conservative treatment

Conservative treatments are commonly adopted when the clinical symptoms,

including pain and restricted motion, first appear. Current conservative treatments

include rest, non-steroid anti-inflammation drug (NSAIDs), glucocorticosteroid

45

injection, platelet-rich plasma therapy, physiotherapies and physical modalities.

1.1.5.2.1 Rest

In the early stage of injury, rest with proper protection including cast is an effective

treatment to prevent repetitive injury caused by mechanical overload and allow the

self-repair of tendon (Darlington and Coomes 1977).

1.1.5.2.2 NSAIDs

NSAIDs are commonly used to reduce the pain and inflammation in soft-tissue

injuries. NSAIDs are able to block the acute inflammatory response through

non-specific cyclo-oxygenase inhibition. Although pain can be reduced by this

treatment, the use of NSAIDs in tendon injury remains controversial. Various studies

showed that they are potentially deleterious to tissue healing and might decrease the

mechanical properties of tendons (Kulick, Smith et al. 1986; Weiler, Unterhauser et

al. 2002; Forslund, Bylander et al. 2003).

1.1.5.2.3 Glucocorticosteroid Injection

Glucocorticosteroid injection has been used since the 1950s as a pain management

method for tendinopathy. The glucocorticosteroid is injected in and around the

chronic tendon injury. They are reported to reduce pain and recover the range of

46

motion to prevent stiffness (Darlington and Coomes 1977; Blair, Rokito et al. 1996).

Although the improvement following glucocorticosteroid injection is significant, the

effects are often temporary (Coombes, Bisset et al. 2010). There are also several

cases reported that tendon rupture followed corticosteroid injection, including

Achilles tendon rupture (Ford and DeBender 1979; Kleinman and Gross 1983).

Moreover, complications include weakened mechanical properties (Wiggins, Fadale

et al. 1995; Tillander, Franzen et al. 1999) and aggravated pain (Goldfarb, Gelberman

et al. 2007). In chronic tendinopathy, the absence of inflammation provides no basic

target to adopt steroid injection. It is worth mentioning that the study of Kabata et al.

showed steroid injection induced osteonecrosis-like lesions and cell apoptosis around

the lesions (Kabata, Kubo et al. 2000).

1.1.5.2.4 Platelet-rich Plasma (PRP) Therapy

Platelet-rich plasma (PRP) therapy is a technology that aims to deliver bioactive

agents to the injury site, to enable the activation of proliferative and anabolic cellular

response to then enhance the repair mechanism of the tissue(Anitua, Sanchez et al.

2006). PRP has been proposed as a treatment for various orthopaedic disorders and

conditions. Clinical use of PRP on tendon-related injuries and disorders is over a

decade, but the outcomes remained conflicting. There are 2 different Food and Drug

Administration (FDA) approved PRP extraction methods, SMARTPEP (SmartPREP,

Harvest Technologies Corp., Norwell, MA) and Platelet Concentrating Collection

Systems (3i/Implant Innovations, Palm Beach Gardens, FL)(Arora, Ramanayake et al.

2009). To obtain the PRP clinical trial data on human Achilles tendon, key words

47

‘PRP’ and ‘Achilles tendon’ were used in Pubmed, initially, 32 articles were

identified. Of these, 9 were clinical studies, which were summarized in Table 1.3. The

clinical outcome of PRP on Achilles tendon varied widely. In clinical studies,

Sanchez et al. reported that PRP injection followed by surgical repair of ruptured

Achilles tendon showed better and faster recovery at 6 months(Sanchez, Anitua et al.

2007), whilst Schepull et al. reported no significant improvement in biomechanical

tests and a worse clinical outcome compared to the non-injection group over 12

months(Schepull, Kvist et al. 2011). Several case studies suggested that patients

receiving PRP treatment for Achilles partial rupture and tendinopathy showed

significant reduction in pain and clinical improvement(Filardo, Presti et al. 2010;

Gaweda, Tarczynska et al. 2010; Finnoff, Fowler et al. 2011; Owens, Ginnetti et al.

2011; Monto 2012), whilst in randomized trials, de Vos et al. reported clinical

improvement in both groups but no significant different between the PRP injection

group and saline injection group(de Vos, Weir et al. 2010). Despite the fact that the

clinical effect of PRP treatment still remains controversial, no complication has been

reported from its clinical application(Mishra, Randelli et al. 2012). The conflicted

published data implied that the PRP treatment is still immature, but this technique

still has great potential for musculoskeletal medicine and orthopaedic surgery.

48

Tabl

e 1.

3 C

linic

al P

RP

trea

tmen

t in

Ach

illes

tend

on

Aut

hors

&

year

L

evel

of

evid

ence

D

isea

se

Patie

nts

num

ber

Follo

w

up

outc

ome

Ref

.

San

chez

et a

l. 20

07

Cas

e se

ries

, le

vel

3 A

chil

les

Ten

don

ru

ptur

e 6

6 m

onth

s F

aste

r re

cove

ry a

nd b

ette

r cl

inic

al o

utco

me

for

PRP

(San

chez

, Ani

tua

et a

l. 20

07)

Sch

epul

l et

al

. 201

1 R

ando

miz

ed tr

ial,

leve

l 2

Ach

ille

s T

endo

n

rupt

ure

16

PR

P

vs.

14

cont

rol

12

mon

ths

Bet

ter

clin

ical

ou

tcom

e in

co

ntro

l gr

oup,

no

si

gnif

ican

t dif

fere

nce

in m

echa

nica

l pro

pert

ies

(Sch

epul

l, K

vist

et

al.

2011

)

de V

os e

t al

. 20

10

Ran

dom

ized

tria

l, le

vel 1

A

chil

les

tend

inop

athy

27

P

RP

vs

27

sa

line

sol

utio

n 12

m

onth

s C

lini

cal

outc

ome

impr

oved

bu

t no

in

terg

roup

si

gnif

ican

t d

iffe

renc

e (d

e V

os,

Wei

r et

al

. 201

0)

Fil

ardo

et

al.

2010

C

ase

seri

es

Ach

ille

s T

endo

n pa

rtia

l rup

ture

1

18

mon

ths

Qui

ck r

etur

n to

ful

l pre

-inj

ury

spor

t (F

ilar

do,

Pre

sti

et a

l. 20

10)

Gaw

eda

et a

l. 20

10

Cas

e se

ries

A

chill

es

tend

inop

athy

14

18

m

onth

s R

educ

ed p

ain

and

reco

vere

d fu

ncti

on

(Gaw

eda,

T

arcz

ynsk

a et

al.

2010

)

Fin

noff

et

al.

2011

C

ase

seri

es

Ach

illes

te

ndin

opat

hy

14

14

mon

ths

Cli

nica

l im

prov

emen

t

(Fin

noff

, F

owle

r et

al.

2011

)

Ow

ens

et a

l. 20

11

Cas

e se

ries

lev

el

4 A

chil

les

tend

inop

athy

10

24

m

onth

s C

lini

cal

impr

ovem

ent,

no

MR

I as

sess

men

t im

prov

emen

t

(Ow

ens,

Gin

nett

i et

al.

2011

)

49

Tabl

e 1.

4 C

linic

al P

RP

trea

tmen

t in

Ach

illes

tend

on

Mon

to

et a

l. 20

12

Cas

e se

ries

, le

vel

4 A

chil

les

tend

inop

athy

30

24

m

onth

s S

igni

fica

nt c

lini

cal i

mpr

ovem

ent i

n 28

pat

ient

s (M

onto

201

2)

Fer

rero

et a

l

2012

Cas

e se

ries

A

chil

les

and

pate

llar

tend

inop

athy

30

A

chil

les

tend

on,

28

pate

llar

tend

ons

6 mon

ths

Sig

nifi

cant

cli

nica

l im

prov

emen

t (F

erre

ro,

Fab

bro

et a

l. 20

12)

50

1.1.5.2.5 Physical modalities

Physical modalities such as low-energy shock-wave (Rompe, Furia et al. 2008), laser

(Bjordal, Lopes-Martins et al. 2006), therapeutic ultrasound (Baysal, Bilsel et al.

2006) and heat (Giombini, Di Cesare et al. 2006) have also been used as a treatment

for tendon and ligament injuries. These modalities were suggested to be able to

relieve pain by altering the local vascular system and improve the mechanical

properties by simulating collagen production. Studies showed the success rate of

physical modalities varies widely; the range is from less than 50% to about 80%

(Samilson and Binder 1975; Chard, Sattelle et al. 1988; Hoying and Williams 1996;

Morrison, Frogameni et al. 1997; Goldberg, Nowinski et al. 2001). Due to

recurrence on longer follow-ups, the success rate was reported to drop to around 50%

in recent study (Baysal, Bilsel et al. 2006; Bjordal, Lopes-Martins et al. 2006; Rompe,

Furia et al. 2008). It was suggested that better results are achieved in patients with

minor symptoms.

1.1.5.2.6 Cell Therapy

Cell therapy is the procedure to deliver new cells into a tissue to stimulate the tissue

regeneration. Cell therapy for tendon repair has achieved great success in a rabbit model,

however, the argument about the best cell type in tendon therapy remains inconclusive.

Stem cells, dermal fibroblasts and tenocytes are the most commonly used cell types for

tendon healing (Obaid and Connell 2010). Mesenchymal stem cells are a potential

candidate due to their rapid proliferation, hypoimmunogenicity and multilineage

51

differentiation ability(Uccelli, Moretta et al. 2006). In the past decade promising results

have been achieved using bone marrow-derived stem cells (BMSCs) in tendon repair.

BMSCs therapy in a rabbit model have been reported to stimulate tendon regeneration,

achieve a better neo-tendon morphology and improve tendon biomechanical properties

(Awad, Butler et al. 1999; Ouyang, Goh et al. 2004; Chong, Ang et al. 2007;

Hankemeier, van Griensven et al. 2007). However, the potential risk of ectopic bone

formation needed to be considered(Ross, Duxson et al. 1987). Tendon-derived stem

cells (TDSCs) were first discovered by Bi et al. in 2007(Bi, Ehirchiou et al. 2007), and

they have multi-differentiation potential into musculoskeletal tissue. Ni et al. report that

TDSCs-engineered tendon is able to differentiate into tendon-like tissue and repair the

tendon defect (Ni, Lui et al. 2012; Ni, Rui et al. 2013). Adipose-derived stem cells

(ADSCs) have the advantage of wide availability and are easy to obtain (Obaid and

Connell 2010). Uysal et al. reported that the application of ADSCs in a rabbit tendon

repair model exhibited better tendon healing and biomechanical properties (Uysal and

Mizuno 2010; Uysal and Mizuno 2011; Uysal, Tobita et al. 2012). Synovium-derived

stem cells (SDSCs) were first identified by De Bari et al. in 2001 (De Bari, Dell'Accio

et al. 2001), and they have proven effective in a wide range of musculoskeletal

disorders (Fan, Varshney et al. 2009). The therapeutic effect of SMSCs in the

tendon-bone junction has been reported by Tomita et al. and Ju et al.(Tomita, Yasuda et

al. 2001; Ju, Muneta et al. 2008). Dermal fibroblasts have also been shown to form

tendon tissue (Liu, Chen et al. 2006; Deng, Liu et al. 2009; Woon, Kraus et al. 2011),

although further research has suggested that the healing process using skin-derived

fibroblasts is suppressed with a lack of tenocyte markers and histopathologic

correlations (Chen, Wang et al. 2010; Obaid and Connell 2010). Clinical trials of dermal

52

fibroblasts on lateral epicondylitis (LE) suggested that it a safe and effective treatment

and no significant complication in the majority of patients (Connell, Datir et al. 2009).

Being the native cell source, tenocytes and in situ fibroblasts are arguably the most ideal

cell sources for tendon and ligament therapy respectively. Preclinical and early clinical

studies using these native cell sources are promising, however, the potential morbidity

to the donor site needs to be considered(Cooper, Lu et al. 2005; Lee, Shin et al. 2005;

Webb, Hitchcock et al. 2006; Androjna, Spragg et al. 2007; Chen, Willers et al. 2007;

Freeman, Woods et al. 2007; Moffat, Kwei et al. 2009; Saber, Zhang et al. 2010). In the

most recent clinical trial of autologous tenocyte injection (ATI), patients with chronic

LE showed significant improved function and structural repair after ATI, and no adverse

event was reported at the biopsy sites(Wang, Breidahl et al. 2013).

53

Tabl

e 1.

5 A

sum

mar

y of

cel

l the

rapy

cel

l typ

e fo

r te

ndon

hea

ling

.

Cel

l Typ

e S

ourc

e A

dvan

tage

D

isad

vant

age

Mes

ench

ymal

S

tem

Cel

ls

(MS

C)

Bon

e m

arro

w-d

eriv

ed

1.

Mul

ti-d

iffe

rent

iati

on

abil

ity

incl

udin

g te

nocy

te

2.

Hig

h pr

olif

erat

ion

rate

3.

Hyp

oim

mun

ogen

icit

y

1.

Acc

eler

ate

tend

on h

ealin

g

2.

Impr

ove

mec

hani

cal p

rope

rtie

s of

tend

on

Dif

ficu

lt

to

man

ipul

ate

the

diff

eren

tiat

ion

of

stem

ce

ll

into

de

sira

ble

cell

type

. T

endo

n-de

rive

d 1.

B

ette

r te

noge

nic

diff

eren

tiat

ion

pote

ntia

l

2.

Ten

don-

like

tiss

ue f

orm

atio

n

Adi

pose

tiss

ue-d

eriv

ed

1.

Wid

e av

aila

bili

ty

2.

Eas

y to

ext

ract

3.

No

dam

age

to d

onor

sit

e

Syn

oviu

m-d

eriv

ed

Indu

ce b

one-

tend

on h

eali

ng

Fib

robl

asts

S

kin

1.

Wid

e av

aila

bili

ty

2.

No

dam

age

to d

onor

sit

e

3.

Non

-inv

asiv

e pr

oced

ure

for

cell

har

vest

1.

Dif

fere

nt

cell

type

2.

Unc

erta

in

beha

vior

in

tend

on

54

nich

e

Tabl

e 1.

6 A

sum

mar

y of

cel

l the

rapy

cel

l typ

e fo

r te

ndon

hea

ling

.

Ten

ocyt

e T

endo

n N

ativ

e ce

ll s

ourc

es

Pot

enti

al

mor

bidi

ty

to

dono

r si

te

55

1.1.5.2.7 Physiotherapies

Physiotherapy has been accepted as one of the mainstays of conservative treatment

for chronic tendon injuries. As a force transferring tissue, the biomechanical

environment is essential for tendon homeostasis. Eccentric overloading is a common

treatment for chronic Achilles tendinopathy. There are two types of eccentric

exercises used, maximize the loading on the calf muscle with a straight knee or

eccentrically load the soleus muscle with the knee bent (Alfredson, Pietila et al.

1998). Several studies reported positive effect on symptom relief(Stanish, Rubinovich

et al. 1986; Alfredson, Pietila et al. 1998; Silbernagel, Thomee et al. 2001; Alfredson

and Lorentzon 2003; Fahlstrom, Jonsson et al. 2003; Roos, Engstrom et al. 2004;

Shalabi, Kristoffersen-Wilberg et al. 2004), however, the study of Woodley et al.

showed no effect(Woodley, Newsham-West et al. 2007) and Rompe et al. report an

inferior results(Rompe, Furia et al. 2008). Although the eccentric overloading may be

effective for pain relief and tendon repair, the magnitude of loading and the

underlying mechanism remain unclear.

1.1.5.2.8 Tissue engineering

Tissue engineering’s aim is to grow tissue in the laboratory and usually involves a

procedure that cultures the engineered tissue using different cell sources combined with

various bioscaffolds in vitro. The development of tissue engineering is based on the idea

that instead of repairing the damaged tissue, it may be better to replace it. The detail is

discussed in 1.2.1

56

1.1.6 Summary

Achilles tendinopathy is a degenerative disease that results in the loss of the

functional mechanical properties of tendon and adversely affects the daily activities

of those afflicted. Despite the prevalence of those suffering from this condition, our

knowledge about tendinopathy remains relatively poor. Although the basic

characteristics of tendinopathy that we do understand enable diagnosis, the aetiology

of this degenerative process remains unclear. Moreover, none of the current

treatments, either conservative or surgical, are able to completely or reliably repair

the tendon integrity and restore the tendon mechanical properties. Therefore, further

studies are required to both better understand the disease and to develop effective

treatment strategies.

1.2 Introduction to Bioreactor Design for Tendon/Ligament

Engineering

(This section has been published in ‘Tissue Engineering Part B’ titled “Bioreactor

design for tendon/ligament engineering” PDF refer to appendix 1)

Autograph and allograft transplantations are a common surgical treatments for tendon

and ligaments injured or degenerated. However, the risks of damage to the donor site

from which the autograft are taken, and the potential immune reaction for allografts are

major concerns (Coupens, Yates et al. 1992; Cerullo, Puddu et al. 1995; Harner, Olson

et al. 1996). A promising translational approach to the treatment of tendon/ligament

57

injury or degeneration is through the use of engineered autologous grafts made available

through the development of bioreactors that generate tendon/ligament tissue in vitro.

One common view is that the key to a successful bioreactor is being able to recreate, in

vitro, the cell microenvironments that are experienced by cells in vivo. The cell

microenvironments can be defined using cell morphological information with data from

molecular biology, biochemistry and biomechanics. This review aims to clarify the

requirements for a ‘successful bioreactor’ that may be used for tendon/ligament

engineering, and to provide an overview of the range of components found in

tendon/ligament bioreactors, including custom-made and commercial products. We will

also discuss the studies that have involved the application of tendon/ligament

bioreactors.

1.2.1 Common key elements for tendon/ligament tissue engineering

Based on the composition and function of tendon/ligament tissue one must consider the

four basic elements for their successful regeneration: the cell source, the characteristics

of the scaffold matrix and establishing an appropriate chemical and physical cellular

microenvironment.

An ideal cell source should meet the following three requirements: availability, rapid

proliferation and the ability to differentiate into in situ cells(Arnsdorf, Jones et al. 2009).

The details refer to 1.1.5.2.6.

58

Apart from the selection of the cell source, cell seeding also plays an essential role in

the development of engineered tendon and ligament. Several reports indicate that

sufficient cell number and a uniform distribution throughout the scaffold is desirable for

achieving a homogeneous ECM deposition in vitro (Freed, Marquis et al. 1993; Freed,

Vunjak-Novakovic et al. 1993). Compared to lower initial cell seeding density, high

seeding density has been shown to result in increasing ECM deposition rate, a higher

final cell number and better cellular morphology (Awad, Butler et al. 2000; Wang,

Seshareddy et al. 2009). However, over-seeding also has potential for negative effects

on nutrient delivery, and consequently cellular metabolism and cell viability. Nutrient

depletion at high cell seeding densities could lead to spatially in homogenous ECM

production (Zhang, Gardiner et al. 2008). A study by Issa et al. showed

mechano-stimulated human umbilical veins seeded with 3 million cells/ml had better

cellular proliferation rates than other groups(Issa, Engebretson et al. 2011). However,

optimal seeding density will vary depending on the cell source and the bioscaffold

physical and chemical properties, as these properties will affect nutrient transport and

rates of consumption, as well as the mechanical and chemical stimuli provided to the

attached cells(Zhang, Gardiner et al. 2008).

The construct scaffold plays an important role in engineering the new tendon/ligament

tissue. Ideally the scaffold should be able to provide substantialinitial mechanical

strength for its immediate post-implantation functional role, while providing a suitable

biological environment for cell migration and proliferation. Furthermore, the

degradation rate of the biomaterial needs to be comparable with the rate of tissue

59

synthesis, to allow the eventual replacement of the starting scaffold with neo-tissue.

Both synthetic and natural biomaterials are commonly used in tendon/ligament

engineering. Synthetic polymer scaffolds have the advantage of reproducible

mechanical and chemical properties, and they are relatively easy to fabricate into

different sizes(Lee, Shin et al. 2005; Pham, Sharma et al. 2006; Sahoo, Ouyang et al.

2006; Sahoo, Cho-Hong et al. 2007; Moffat, Kwei et al. 2009). However, their rapid

degradation rate and potential risk of releasing acidic by-products or toxic polyesters

during degradation have limited their application in clinical trials. Given these

disadvantages, more researchers have turned their focus on exploring natural

biomaterials (Derwin, Baker et al. 2006; Juncosa-Melvin, Shearn et al. 2006; Chen,

Willers et al. 2007; Gilbert, Stewart-Akers et al. 2007; Nirmalanandhan, Rao et al. 2008;

Nirmalanandhan, Shearn et al. 2008; Fleming, Spindler et al. 2009; Kinneberg,

Nirmalanandhan et al.). Being the main component of native tendon, collagen type I is

the most obvious choice of material. Although the biocompatibility is excellent, the

poor mechanical properties of reconstituted type I collagen scaffolds has limited their

further development as a load bearing material. Silk fibroin, on the other hand, has

similar biocompatibility as collagen scaffolds and comparable mechanical properties as

native tendon/ligament. In several in vivo studies, silk fibroin-based engineered

ligaments have been proven their ability to restore the function of injured ligament

(Chen, Qi et al. 2008; Fan, Liu et al. 2008; Fan, Liu et al. 2009). Another option is the

decellularised tendon/ligament construct. Although the mechanical and biological

properties are a better match to native tissue than any other currently available scaffolds,

60

donor cells may remain in the allograft, even with strict sterilization and cleaning, and

thus they can potentially cause inflammatory responses (Malcarney, Bonar et al. 2005;

Zheng, Chen et al. 2005).

After choosing an appropriate cell source and scaffold type, the tissue needs to be

encouraged to develop the properties of native tissue by providing an appropriate

biochemical and biomechanical environment to stimulate ECM synthesis. Regarding the

biochemical environment, several growth factors have been found to play an important

role in tendon/ligament formation and healing. These include Insulin-like growth

factor-I (IGF-I), vascular endothelial growth factor (VEGF), platelet-derived growth

factor (PDGF), basic fibroblast growth factor (bFGF), transforming growth factor

(TGF) and growth differentiation factor 5 (GDF-5).The roles of TGFand GDF-5

seem to be particularly prominent. TGFremains active throughout tendon/ligament

healing(Chang, Thunder et al. 2000), and is able to regulate cell migration, proteinase

expression, fibronectin binding interactions, cell proliferation and collagen production

(Bennett and Schultz 1993; Wojciak and Crossan 1994; Marui, Niyibizi et al. 1997;

Chang, Thunder et al. 2000). Recent studies demonstrated that tendon and ligament

formation was impaired in TGF2 and TGF3 knockout mouse embryos, reinforce the

importance of TGFs in tendon development and homeostasis(Pryce, Watson et al.

2009). GDF-5 regulates cell growth and differentiation, with a lack of GDF-5 causing

delayed tendon healing, irregular collagen type I fibrils and weakened fibril mechanical

properties(Clark, Johnson et al. 2001; Mikic, Schalet et al. 2001; Chhabra, Tsou et al.

2003).

61

How these various biochemical factors should be introduced into the tissue bioreactor

system to shape the chemical environment is an extremely challenging and open

question. Specifically, at what concentrations, in what combinations and at what

sequence or timing should they be made available? Studies have shown that the

expression of these factors in tendon repair change over periods of days (Dahlgren,

Mohammed et al. 2005; Kobayashi, Itoi et al. 2006; Chamberlain, Crowley et al. 2009).

Presumably as the engineered tissue progresses, cells begin to control their own

biochemical environment, and the role of the bioreactor is to now provide the ‘building

block’ nutrients and expected systemic signals, along with mechanical stimulus. The

early stage in the tissue-engineered tendon is likely to be the most critical in

establishing tendon development along a pathway to resemble native tendon. Finding

the correct combination of factors is daunting due to the complexity arising from the

multitude of possible combinations and interactions. Systematically varying

experimental conditions, coupled with computational modeling of transport processes

and signaling molecule pathways leading to cell responses, provide the only

conceivable means to both understanding and efficiently optimizing the tissue bioreactor

system.

Tendon’s primary function is mechanical. It operates in a varying load environment,

both on short timescales (e.g. walking, running) and on longer timescales (e.g. changes

in body size with age). Tendon responds to its mechanical environment through changes

in ECM biosynthesis and degradation. Unsurprisingly, given its functional role, a

suitable mechanical stimulus is vital for tendon/ligament homeostasis. In fact, it has

62

been shown that after 4 weeks in a load-free culture environment tenocytes lose their

native elongated morphology, become increasingly rounded, and the collagen fiber

becomes more crimped (Hannafin, Arnoczky et al. 1995). In cell-free reconstructed

collagen fibril network systems, a tensile load has been shown to be protective to

degradation by MMP-8 (Flynn, Bhole et al. 2010). Furthermore, cyclic stretching has

been shown to produce an up to 9 fold increase in the cell number of an engineered

tendon compared with a static culture over a 2 week period(Abousleiman, Reyes et al.

2009). Finally, an appropriate mechanical environment could help guide collagen fiber

formation, i.e. along the direction of loading, which is able to enhance or optimize the

mechanical properties including stiffness, elastic modulus, maximum tensile stress and

maximum force(Juncosa-Melvin, Shearn et al. 2006; Saber, Zhang et al. 2010; Woon,

Kraus et al. 2011).

Rather than being two separate signals, there is crosstalk between mechanical and

chemical signals. Recent studies have shown that gradual and temporary loss of tensile

loading leads to reversible loss of Scleraxis (Scx) expression, which is a transcription

factor specific for tenocytes and their progenitors. In addition, it has been shown that

TGF directly induced the expression of Scx in cultured tenocytes isolated from mice

(Maeda, Sakabe et al. 2011). In Scx-/- mice, a disordered limb tendon phenotype was

observed (Lejard, Brideau et al. 2007), and similar phenomenon happened in TGF

type II receptor gene knock out (Tgfbr2-/-) mice with dramatic loss of Scx

expression(Pryce, Watson et al. 2009).

63

Providing a suitable biomechanical signal is clearly an important component for the

success of tendon/ligament engineering. Generating a suitable mechanical signal within

the bioreactor system is critically important for tendon/ligament tissue engineering.

1.2.2 Bioreactor design specific to tendon/ligament engineering

Despite the increasing appreciation of tendon/ligament biology and function,

conventional culture methods do not seem to meet the biochemical and biomechanical

requirements to generate bioengineered tendon/ligament in vitro. A bioreactor system

that subjects the ‘cell culture’ to programmable mechanical stimulation (PMS),

mimicking the physiological conditions of tendon/ligament in vivo, while allowing

cellular proliferation, differentiation and matrix production in a mechanical environment,

may provide a solution.

Bioreactors for tendon/ligament engineering are different to the systems that have been

used in various other tissue engineering fields in the past decades, i.e. systems for

muscle(Dennis, Smith et al. 2009), liver(Catapano, Patzer et al. 2010) and bone(Rauh,

Milan et al. 2011; Salter, Goh et al. 2012). Compared to other bioreactors, the main task

of the bioreactor for tendon/ligament engineering is to provide the proper biomechanical

and biochemical environment specific to tendon/ligament formation. In order to achieve

this, certain basic components are required, i.e. the actuating system and the culture

chamber, which can provide the construct’s PMS and controlled culture environment

respectively. Furthermore, the bioreactor may also include a medium circulation system,

64

monitoring system, feedback system, and a medium analysis system, depending on the

operational requirements (Figure 1.1). With these facts in mind, several custom-made

bioreactors have been developed for tendon/ligament engineering (see Table 1.1), and

the aforementioned components of these are now are discussed in detail.

Figure1.1 Schematic demonstration of connection between different components of tendon/ligament bioreactor system

65

Table 1.1 Components of custom-made bioreactor systems

Reference Actuating system

Culture chamber

Monitor system

Feedback system

Medium circulation system

Medium analysis system

(Altman, Lu et al. 2002)

Biaxial Multiple

(Webb, Hitchcock et al. 2006)

Uniaxial Single chamber, multiple samples

(Juncosa-Melvin, Shearn et al. 2006),(Nirmalanandhan, Rao et al. 2008),(Nirmalanandhan, Shearn et al. 2008),(Butler, Gooch et al. 2010)

Uniaxial Multiple Displacement

(Androjna, Spragg et al. 2007)

Uniaxial Multiple Force

(Nguyen, Liang et al. 2009)

Uniaxial Single Force

(Abousleiman, Reyes et al. 2009)

Uniaxial Multiple

(Butler, Hunter et al. 2009)

Uniaxial Single chamber, multiple samples

Displacement Force

(Chen, Yin et al. 2010)

Uniaxial Multiple

(Doroski, Levenston et al. 2010)

Uniaxial Multiple Displacement

(Parent, Huppe et al. 2011)

Uniaxial Multiple Displacement Force

66

1.2.2.1 Actuator and culture chamber design

The actuating system is the main component for providing different PMS to engineered

tissue. Pneumatic actuators(Juncosa-Melvin, Shearn et al. 2006; Nirmalanandhan, Rao

et al. 2008; Nirmalanandhan, Shearn et al. 2008), linear motors(Nguyen, Liang et al.

2009; Doroski, Levenston et al. 2010) and step motor-ball screws(Altman, Lu et al.

2002; Webb, Hitchcock et al. 2006; Chen, Yin et al. 2010) are the most common

actuators used in tendon/ligament bioreactors. Pneumatic actuators have several

advantages, including the ease of maintenance, cleanliness, low cost and high

power-to-weight ratio(Varseveld and Bone 1997). Unfortunately, by using air as a media,

pneumatic actuators are subject to high friction. The sensitivity and response to an input

signal is relatively slow because of the “dead band” and “dead time” caused by stiction

and air compressibility. Due to these nonlinearities, it is difficult to achieve accurate

position control with pneumatic actuators(Varseveld and Bone 1997). Typically, the

accuracy of pneumatic actuators is approximately0.1mm, which is not insignificant

when typical tendon bioreactors require <5% strain on 1-5cm tissues. Compared with

pneumatic actuators, electrical actuators are more expensive, but the level of accuracy in

their positional control is much higher. A direct-drive linear motor is able to provide

high-speed/high-accuracy linear motion by eliminating mechanical transmission(Alter

and Tsao 1994; Alter and Tsao 1996; Braembussche, Swevers et al. 1996), and the

accuracy is the highest of all three actuators; i.e. ~1µm. Step motor-ball screw

transmission systems (SMBS) are based on a ball-screw linkage and a crank-slider

mechanism. In general, SMBS transfer the rotation of the crank to a reciprocating

motion of the screw(Liu, Hsu et al. 2001). By choosing different ball screws, the

optimal speed range and output force can be selected using Equation (1), i.e.

67

vFlT2

(1)

whereT is torque applied to screw, F is linear force, l is ball screw lead, and ν is ball

screw efficiency. Although SMBS is not as accurate as a linear motor, a positional

accuracy of around 5µm can still be achieved. In a multi-chamber shared loading

system, SMBS is widely used given its high accuracy and relative high loading capacity

(Altman, Lu et al. 2002; Webb, Hitchcock et al. 2006; Androjna, Spragg et al. 2007). In

independent loading multi-chamber systems, linear motors are more popular due to their

small size and relatively simple mechanical arrangement(Nguyen, Liang et al. 2009;

Parent, Huppe et al. 2011).

In the addition to the actuator, the connection between the mechanical input and the

tissue is vital and often presents a major challenge. During PMS of the tissue in culture

an even distribution of force throughout the entire sample is critical, otherwise tissue

integrity is compromised by overloading the mechanical connection regions and/or by

inhomogeneous mechanical stimulation. Different strategies of applying PMS have been

adopted based on different construct dimensions. In a study by Chen et al., cell-seeded

knitted silk-collagen sponge scaffolds were fixed on stainless rings and connected to

sample hooks(Chen, Yin et al. 2010). The bioreactor system by Juncosa-Melvin et al.

applied two posts to fix the construct, punching through the scaffold as shown in Figure

1. 2(Juncosa-Melvin, Shearn et al. 2006). However, non-uniform construct deformation

is clearly apparent. Tensile force is focused on the side of constructs, which causes

uneven distribution of mechanical stimulation. Although tissue clamps are the most

68

popular and relatively effective method for holding tendon constructs(Webb, Hitchcock

et al. 2006; Nguyen, Liang et al. 2009; Parent, Huppe et al. 2011), the potential for

damage at the clamping region needs to be considered. The method of reproducibly

applying uniform loads to soft tissue without tissue damage or slippage is a critical

problem in need of a satisfactory solution. It is our opinion that a robust clamping

region should be designed along with the artificial bioscaffold to ensure the proper

connection between the sample and the PMS.

69

Figure 1.2 Stem cell seeded Collagen sponge deformation during mechanical stimulation. Modified

from reference 156

The culture chamber is an essential part of whole bioreactor system. The high humidity

of culturing conditions (99% humidity, 37℃ and 5% CO2) and chemistry of the culture

medium are corrosive to many materials. Corrosion products may in turn be toxic to the

tissue. Therefore, noncorrosive and autoclavable material such as stainless steel,

polymethylemethacrylate (PMMA), polyoxymethylene, polycarbonate, glass and silicon

are preferred and widely used in culture chamber design (Altman, Lu et al. 2002;

70

Abousleiman, Reyes et al. 2009; Butler, Hunter et al. 2009; Chen, Yin et al. 2010;

Doroski, Levenston et al. 2010).

The chamber structure is an important consideration in the bioreactor design, with most

currently available culture chambers divided into two groups: integrated(Webb,

Hitchcock et al. 2006; Butler, Hunter et al. 2009) and separated

chambers(Juncosa-Melvin, Shearn et al. 2006; Chen, Yin et al. 2010; Doroski,

Levenston et al. 2010) (Altman, Lu et al. 2002; Androjna, Spragg et al. 2007; Parent,

Huppe et al. 2011) (Butler, Gooch et al. 2010). In integrated chambers, multiple

samples are cultured while sharing the same culture medium. Conversely, separated

chambers can provide separate culture environments for each sample. Although the

complexity of design and manufacturing costs may be higher in a separated chamber

system, reduced cross-contamination and the option of independent environmental

control are distinct advantages.

During tissue culture, sufficient air exchange within the culture chamber is critical.

Air exchange in conventional cell culture incubators is through the integrated

hydrophobic filter of the culture flask and the gap between the leak and the culture

dish/well plate. Therefore, the ideal design for the culture chamber should be similar.

Like conventional cell culture, an unsealed chamber bioreactor connects to the

outside environment through various ways (Juncosa-Melvin, Shearn et al. 2006;

Androjna, Spragg et al. 2007; Nguyen, Liang et al. 2009; Doroski, Levenston et al.

71

2010), such as the hydrophobic filter leak(Webb, Hitchcock et al. 2006) and the

labyrinth channel (Figure 1.3)(Parent, Huppe et al. 2011). In a study by Webb et al.

(Webb, Hitchcock et al. 2006), a modified tissue culture flask was used as an

integrated culture chamber, and could culture up to four samples simultaneously.

Although there are potential risks of cross-contamination from different samples and

toxicity from autoclaving the culture flask, the integrated hydrophobic filter leak can

ensure adequate gas exchange without inducing contamination. In the bioreactor

system used in the research of Parent et al.(Parent, Huppe et al. 2011), a labyrinth

channel was added to improve the air exchange and eliminate contamination as

shown as Figure 1.3. However, in closed chambers air exchange mostly depends on

medium circulation, which is now discussed in the following section.

Figure 1.3 Schematic drawings of air exchange through the labyrinth channel in the culture chamber. Modified

from reference (Parent, Huppe et al. 2011)

1.2.2.2 Environmental control and medium circulation systems

Although to the best of our knowledge, no contamination has been reported in any

bioreactor study, transportation of the bioreactor and opening of chamber for medium

exchange every three days, especially for multi-chambers system, is still a potential

72

contamination risk. Therefore, a medium circulation system can be introduced to

improve the efficiency and minimize these risks. In addition to the advantages of

reduced contamination, a circulating medium may be better able to infiltrate into

cultured tissue. For example, perfusion bioreactors used in bone engineering circulate

culture medium for better nutrient delivery and subsequent improved cell numbers

(Cartmell, Porter et al. 2003; Uemura, Dong et al. 2003; Meinel, Karageorgiou et al.

2004; Holtorf, Jansen et al. 2005; Janssen, Oostra et al. 2006; Olivier, Hivart et al.

2007). Similarly, human umbilical vein cultured under a circulating medium had

approximately three times the cellular number as those cultured using a quiescent

medium(Abousleiman, Reyes et al. 2009).

A traditional incubator based bioreactor and/or independent bioreactor can be used for

tendon/ligament engineering. In an incubator bioreactor system, air exchange is through

the hydrophobic filter leak of the medium reservoir, and proper CO2 and temperature

level is controlled by the incubator. Then by circulating the culture medium, suitable

conditions can be applied to engineered tendon/ligament, as shown in Figure 1.4

(Abousleiman, Reyes et al. 2009) . However, an independent bioreactor system, as

shown Figure 1.5, does not rely on an incubator to control the culture environment. The

percentage of different gases (PO2, CO2 and N2) are control by air valves (Altman, Lu et

al. 2002), and the mixed gas is humidified first, and then passed into a medium heater.

Waste gas emission is transported through a filter in the case of contamination. The

prepared, warm culture medium is circulated through the bioreactor culture chamber.

For certain tissues such as cartilage, some specific culture conditions are required. For

73

example, under a low oxygen environment, engineered cartilage displays faster matrix

glycosaminoglycan deposition rate, and better cellular morphology but with less

dedifferentiation (Hansen, Schunke et al. 2001; Domm, Schunke et al. 2002; Saini and

Wick 2004). The environmental chamber allows researchers to manipulate different

culture conditions thereby enabling a systematic study of cell growth and differentiation

into functional tissue.

74

Figure 1.4 Schematic illustration of an incubator bioreactor system. Suitable temperature,

humidity and CO2 level of culture medium are maintained by the incubator in the medium reservoir.

The medium is circulated by a pump. The waste valve is closed normally and it will open during

medium exchange.

75

Figure 1.5 Schematic diagram of an independent bioreactor system. Suitable temperature, humidity

and CO2 level of the culture medium are control by the environmental chamber. Cold culture

medium is pumped to the environmental chamber for heating and then circulated through the culture

chamber. The waste valve is closed normally and it will open during medium exchange.

1.2.2.3 Monitoring and feedback systems

The biomechanical properties of the final engineered tendon/ligament should be a

central concern of bioreactor design, as the tendon/ligament’s functional role in the

body is primarily mechanical. The maximum load and elastic modulus are essential

mechanical properties to evaluate the suitability of engineered tendon/ligament.

However, in most studies, mechanical tests are performed only at the end of tissue

culture. For example, the stiffness of the constructs at different time points during

76

culture is rarely recorded or assessed. The correlation between stiffness and tissue

maturation may provide a better understanding about how the cell differentiates into

functional tissue, and for this reason ‘on-line’ stiffness monitoring is likely to be

invaluable.

In tendon/ligament bioreactors force measurement is enabled using sensors called load

cells (Butler, Juncosa-Melvin et al. 2008; Nguyen, Liang et al. 2009; Parent, Huppe et

al. 2011), which are located differently in various bioreactor designs. Load cells located

between the actuator and sample clamps(Butler, Hunter et al. 2009), shown in Figure

1.6A, requires high manufacturing accuracy, as the friction between the shaft and

culture chamber can induce error. For fragile material like biological tissues, this

friction might be higher than the applied load. Conversely, a load cell placed at the end

of the culture chamber, shown in Figure 1.6B, can minimize the fiction between the

actuator and culture chamber and also the fluid resistance during stimulation. In order to

acquire accurate data, load cell selection should be based on the initial mechanical

properties of the constructs. Working with the sensitivity of the load cell, extra attention

is needed when manipulating the construct so as to avoid causing damage to the load

cell through overloading.

77

Figure 1.6 Load cells position in different bioreactor systems. A. load cell is located between the actuator and sample clamps. B. a separate load cell placed at the end of culture chamber system.

In addition to force monitoring, tissue displacement is another important variable that

requires monitoring. Linear variable differential transformers (LVDT) and optical

decoders are commonly used position sensors(Juncosa-Melvin, Shearn et al. 2006;

Parent, Huppe et al. 2011). Although the motion of actuating systems is preprogrammed,

overload of the actuator and manual mis-operation can cause desynchronization

between the program and actual stimulation. The real time displacement monitor can

produce a full record of stimulation position, which then allows researchers to track if

there are any unusual features in the results. Another function of the position sensor is to

provide feedback of the displacement information to the actuator control system to

correct for any desynchronization(Parent, Huppe et al. 2011). Critically, by monitoring

78

load and displacement, real time stiffness during engineered tendon/ligament culture

can be estimated.

Lastly, an imaging component might be useful in the bioreactor system to monitor the

formation of the tissue construct at high resolution and possibly provide information to

the feedback system. Two imaging modalities, including the confocal microscope and

optical coherence tomography (OCT), are potential candidates (Drexler, Morgner et al.

1999; Goetz, Thomas et al. 2008). Both of these techniques are able to acquire

high-resolution image and might be adopted in the dynamic culture environment for

observation of cell and tissue morphology (Drexler, Morgner et al. 1999; Goetz, Thomas

et al. 2008). The magnification of the confocal microscopy may range from 500 to 2400

fold at a dynamic environment. A miniaturized confocal laser scanning probe has been

tested in rodents to provide the visualization of the vascular circulation and cell perfusion

Compared to the confocal microscope, OCT has a higher penetration depth up to 1~2

mm and longer working distance (the distance from the objective lens to the sample

surface) up to 10 mm, which is sufficient for distant imaging of the tissue constructs in

the bioreactor system(Smithpeter, Dunn et al. 1998; Drexler, Morgner et al. 2001;

Burkhardt, Walther et al. 2012) . However, there are two major issues to be addressed

before these systems could be integrated into the bioreactor design. Firstly, for distant

imaging, the optical light needs to penetrate3 layers of media (chamber cover, air and

culture medium) with different refraction index to obtain imaging. Secondly, sample

deformation induced by dynamic mechanical loading can cause image shift resulting in

poor focus or loss of the image. Imaging dynamic sample is extremely difficult in

79

conventional bioreactors, as features of interest (e.g. cells) may leave the field of view.

Techniques need to be developed to move the microscope alone with the sample to

achieve optimal focus on the view of interest. Other than distant imaging, minimally

invasive imaging technique might provide another potential strategy for this obstacle.

Recent studies show that confocal fluorescence microendoscopy and OCT could be

integrated into a hypodermic needle (Pillai, Lorenser et al. ; Quirk, McLaughlin et al.),

which might be inserted into the tissue constructs and thus creating a relatively static

imaging environment. However, the technique is invasive and could damage the tissue

integrity. In all, the current imaging technology is not mature enough to acquire the

dynamic sub-micrometer image, butan imaging system allows dynamic imaging at high

resolution is certainly worth to explore. Cell morphological and tissue structural change

coping to the PMS might provide a different respective for studying the effect of PMS

on tendon/ligament biology and tendon/ligament engineering.

1.2.3 Commercial bioreactor systems for tendon/ligament engineering

Recently, commercial tendon/ligament engineering bioreactor systems have become

available. As discussed above, these basic principles are applied to customized

bioreactors; however, with greater design input and advanced manufacturing techniques,

commercial products are able to provide more accurate and complex environments for

tendon/ligament culture. To our knowledge, two relatively complete commercial

bioreactor systems have been developed recently: The Bose® ElectroForce® BioDynamic®

system and the LigaGen system.

80

The Bose® ElectroForce® BioDynamic® test instrument provides an accurate

programmable uniaxial stretching stimulation and a controllable medium circulation

environment to engineered tendon/ligament, which, theoretically, can be adjusted to

mimic the in vivo biomechanical environment. With a load cell and optional laser

micrometer, this bioreactor system is able to monitor the force/strain curve of

engineered tendon/ligament during the culture period. Two different force and

displacement ranges are available. Single chamber and multiple chambers systems with

shared or independent loading are optional. As there are culture chambers compatible to

the Bose® testing devices such as ElectroForce 3200, biomechanical tests can be done

at different time points without disruption of the tissue culture

(www.bose-electroforce.com).

The LigaGen system is a lightweight (<3kg) incubator compatible bioreactor. It is

capable of applying a maximum force of 40N to the tissue sample, and simulate

complex, and presumably more physiologically realistic, loading patterns. Two systems

are available from LigaGen. L30-1X is a single culture model, which has a 23ml

internal volume chamber for single tissue culture, and the L30-4C is a multiple culture

model with an 80ml chamber for shared dynamic culture on two or four samples. The

standard medium circulation system can reduce contamination risk during medium

exchange. Rather than being a comprehensive system, the bioreactor is extensible to

suite individual needs, by adding various components to the (universal) basic model as

required. With an accessory tissue monitoring sensor, this system is able to achieve

real-time measurements of the sample stiffness during culture. If flow control is

necessary, extra control systems can be installed (http://www.tissuegrowth.com/).

81

However, there are some disadvantages in the commercial bioreactors. Firstly, the

fixation mechanism to stabilize the tendon tissue in the bioreactor cannot be adjusted.

Tissue clamps provided in the commercial bioreactor systems can become less effective

when it comes to the use of a cylindrical scaffold. Secondly, capacities of the chamber

and mechanical input are limited and they can only host small animal tissues. This may

restrict clinical development. Moreover, some ligaments, such as the anterior cruciate

ligament, are not only subjected to tensile force, but also to rotational loading, and none

of the commercial bioreactor systems provide the addition of torsional loading. Lastly,

full-scale commercial bioreactor systems are expensive, and for most of the time, not all

of the functions they provide are commonly used in every study.

1.2.4 Ideal bioreactors for tendon/ligament engineering

The ideal bioreactor should be able to culture tendon- and ligament-like constructs,

which are well-organized, cell-seeded, assemblies of collagen bundles with mechanical

properties functionally similar to the native tissue. Autologous cell-seeded constructs

are biological compatible, and able to provide mechanical support similar to native

tissue, and consequently they are widely studied in tendon/ligament engineering.

However, induction of cell directed collagen fiber reorganization and assembly of

collagen bundles are two important impediments to this approach. PMS can help

provide the necessary signals to cells to increase collagen synthesis, spatially organize

of the collagen along the primary stress direction, and stabilize collagen from

collagenase degradation (Lavagnino, Arnoczky et al. 2003; Yang, Im et al. 2005;

Nguyen, Liang et al. 2009). Moreover, proper PMS can upregulate different

82

proteoglycans, such as decorin, biglycan, fibromodulin and fibronectin, which help the

cells organize the parallel collagen fibrils forming bundles (Scott 1984; Cribb and Scott

1995; Vogel 2004; Webb, Hitchcock et al. 2006; Franchi, Trire et al. 2007).

The host body is in many ways the ultimate bioreactor for all engineered tissues. The

study of Juncosa-Melvin et al. indicated that the maximum force of engineered patellar

tendons increased more than 3000 times after 2 weeks implantation in rabbits

(Juncosa-Melvin, Shearn et al. 2006), and to date none of the bioreactors have been able

to accomplish this outcome. Therefore, an ideal bioreactor should aim to mimic the

dynamic biochemical and biophysical environments in vivo. In musculoskeletal tissue

engineering, various bioreactors have been developed. For instance, muscle tissues are

not only subjected to mechanical stretching but also able to receive the electrical

impulses to simulate inputs from the central nervous system (Ross, Duxson et al. 1987),

and mechanical and electrical stimulation bioreactors have been developed based on

mimicking the in vivo environment (Donnelly, Khodabukus et al. 2010; Sharifpoor,

Simmons et al. 2011). Compared to muscle, the in vivo environment is comparatively

less complex in tendon/ligament and appears to require only passive mechanical input.

Summarizing, the ideal tendon/ligament engineering bioreactor that enables systemic

research should integrate all aforementioned components (Figure 1.1). The bioreactor

should have culture chambers and an actuating system, but also be fitted with a medium

circulation system, an environmental system, a monitoring system, a feedback system,

83

and a waste medium analysis system. This bioreactor should first be able to provide not

only a multiple, suitably sized and sterilized chambers for tissue culture, but also

accurate and programmable mechanical stimulation. Tensile strain and rotation are

needed to mimic different in vivo tendon/ligament loads. Second, the circulated medium

should infiltrate into tissue better than a static medium configuration, and the circulation

system needs to reduce the risks of contamination during medium exchange and drug

delivery. Moreover, environmental control, such as PO2, CO2 and pH level, allows

researchers to explore the impact of different culture conditions on tissue maturation.

Third, a monitoring system should provide the real-time status of cultured

tendon/ligament, such as force and displacement, and based on these data, adjustment of

PMS by use of a feedback system. Fourth, through analysis of the waste medium,

nutrient consumption needs to be evaluated, which may enable the changing of the

medium base at different stages as required, rather than fixed, regular medium exchange

every 3 days. Finally and importantly, the best patterns of PMS for culturing engineered

tendon/ligaments need to be defined.

1.2.5 Previous work in this field

Since the first 3D engineered tendon/ligament bioreactor system published by Altman et

al. in 2002, the effect of PMS on the engineered tendon/ligament has drawn a lot of

research attention. In the past decade, dynamic loading of culture in bioreactor systems

has been proven to have been a significant development in producing an engineered

tendon/ligament. Indeed, various studies have been performed using bioreactor systems

and have met with considerable success, and these are summarized in Table 1.2.

84

Compared with static culture, the tissue produced using PMS has better cell morphology,

including elongated cellular morphology and increased cell density. The mechanical

properties of engineered tendon/ligament, such as tensile strength and elastic modulus,

are also greatly improved by cyclic loading of the tissue culture, as is the microstructure

of the extracellular matrix, such as collagen fiber alignment. Gene expression is also

positively influenced by PMS. For instance, collagen type I expression under dynamic

loading is three times higher than static culture in 2 week(Butler, Hunter et al. 2009).

Although cyclic stretching has been proven to be an effective way to stimulate the

engineered tendon/ligament culture, the optimal stimulation pattern is still unknown.

Nirmalanandhan et al. revealed that a 2.4% strain cycle consisting of 3000 cycles per

day, produced the best linear stiffness in rabbit MSC seeded in type I collagen sponge

(Nirmalanandhan, Shearn et al. 2008). However from the perspective of Butler et al.,

the stimulation pattern should be adjusted based on maturation of the engineered

tendon/ligament, with higher dose loading applied at the later stage of tissue culture

(Butler, Hunter et al. 2009).

85

Tabl

e 1.

1 S

tudi

es u

sing

bio

reac

tor

syst

ems

for

tend

on/l

igam

ent e

ngin

eeri

ng

Fir

st a

utho

r B

iore

acto

r ty

pe

(Com

pany

) P

aram

eter

s of

m

echa

nica

l st

imul

atio

n

Sca

ffol

d m

ater

ial

(d

imen

sion

s)

Cel

l sou

rce

Eff

ect

Ref

eren

ce

Alt

man

(20

02)

Cus

tom

-mad

e st

ep

mot

or

bior

eact

or

wit

h en

viro

nmen

tal

cham

ber

Cyc

lic

stre

tchi

ng

2mm

, 90

o ro

tati

on,

0.01

67H

z,

21

days

Col

lage

n ty

pe

I ge

ls,

Bom

byxm

ori

silk

wor

m s

ilk

fibe

r m

atri

ces

(len

gth:

30 m

m)

Hum

an b

one

mar

row

st

rom

a ce

lls

(hB

MS

C)

Elo

ngat

ion

of

hBM

SC

, cr

oss-

sect

ion

cell

de

nsit

y

(Alt

man

, L

u et

al.

2002

)

Junc

osa-

Mel

vin

(200

6)

Cus

tom

-mad

e pn

eum

atic

cy

lind

er

bior

eact

or

wit

h LV

DT

fo

r di

spla

cem

ent

mon

itor

ing

Cyc

lic

stre

tchi

ng 2

.4%

st

rain

, 0.

0033

Hz,

8h

/day

fo

r 2

wee

ks

Type

I

coll

agen

sp

onge

(2

3

0.8

mm 9

0.

8mm 3

0.

1mm

)

Rab

bit M

SC

M

axim

um

forc

e

, li

near

st

iffn

ess

,

max

imum

st

ress

li

near

mod

ulus

(Jun

cosa

-Mel

vin,

S

hear

n et

al.

2006

)

Web

b (2

006)

C

usto

m-m

ade

step

m

otor

bio

reac

tor

Cyc

lic

stre

tchi

ng 1

0%

stra

in,

0.25

Hz,

8h

/day

fo

r 7d

ays

Pol

yure

than

e co

nstr

uct

(20 1

0 2

mm

)

Hum

an

trac

heal

fi

brob

last

Type

I

coll

agen

,

TG

F

-1

, C

TG

F,

Ela

stin

, al

pha

I ,

Pro

coll

agen

,

Fib

rone

ctin

,

MM

P-1

,

elas

tic

mod

ulus

(Web

b,

Hit

chco

ck

et a

l. 20

06)

86

Tabl

e 1.

2 S

tudi

es u

sing

bio

reac

tor

syst

ems

for

tend

on/l

igam

ent e

ngin

eeri

ng

And

rojn

a (2

007)

C

usto

m-m

ade

bior

eact

or

wit

h lo

ad-d

ispl

acem

ent

mea

sure

sys

tem

Cyc

lic

stre

tchi

ng

9%

stra

in,

twic

e da

ily

for

30m

in e

ach

peri

od s

epar

ated

by

8h r

est f

or 2

wee

ks

Sm

all-

inte

stin

e su

bmuc

osa

(3cm

5c

m 9

5

m)

Dog

te

nocy

te

Cel

l de

nsit

y

, st

iffn

ess

(A

ndro

jna,

Spr

agg

et a

l. 20

07)

Nir

mal

anan

dhan

(2

008)

C

usto

m-m

ade

pneu

mat

ic

cyli

nder

bi

orea

ctor

Cyc

lic

stre

tchi

ng

2.4%

an

d 1.

2%

stra

in,

1Hz,

st

imul

atio

n pe

riod

: 8h

/day

, 10

0 an

d 30

00 c

ycle

s/da

y fo

r 12

day

s

Type

I

coll

agen

sp

onge

R

abbi

t il

iac

cres

t M

SC

The

st

imul

atio

n pa

tter

n of

2.

4%

stra

in,

3000

cy

cles

/day

, 1H

z ha

s be

st e

ffec

t on

in

crea

sing

st

iffn

ess.

(Nir

mal

anan

dhan

, S

hear

n et

al.

2008

)

Nir

mal

anan

dhan

(2

008)

C

usto

m-m

ade

pneu

mat

ic

cyli

nder

bi

orea

ctor

Cyc

lic

stre

tchi

ng

2.4%

st

rain

, 0.

0033

Hz,

8h

/day

fo

r 12

day

s

Type

I

puri

fied

bo

vine

co

llag

en

gel,

type

I

coll

agen

sp

onge

s (l

engt

h: 1

1mm

and

51

mm

)

Rab

bit

MS

C

stif

fnes

s

(Nir

mal

anan

dhan

, R

ao e

t al.

2008

)

Ngu

yen

(200

9)

Cus

tom

-mad

e li

near

ac

tuat

or

bior

eact

or

wit

h lo

ad

cell

fo

r fo

rce

mea

sure

men

t

Pre

load

ed

wit

h 0.

05N

, cy

clic

st

retc

hing

10

%,

0.5H

z,

2h

stim

ulat

ion-

2h

rest

-2h

stim

ulat

ion-

18h

rest

fo

r 5

days

Por

cine

sm

all

inte

stin

e su

bmuc

osa

(len

gth:

2c

m

and

wid

th: 1

cm)

Rab

bit

MC

L

fibr

obla

st

Impr

oved

fi

ber

orie

ntat

ion,

fi

ber

angu

lar

disp

ersi

on

,

bett

er

orga

nize

d co

llag

en

fibe

r, el

onga

ted

cell

m

orph

olog

y,

(Ngu

yen,

Lia

ng e

t al

. 200

9)

87

Tabl

e 1.

3 S

tudi

es u

sing

bio

reac

tor

syst

ems

for

tend

on/l

igam

ent e

ngin

eeri

ng

Abo

usle

iman

(2

009)

C

usto

m-m

ade

line

ar

actu

ator

bi

orea

ctor

wit

h m

ediu

m

circ

ulat

ion

syst

em

Cyc

lic

stre

tchi

ng

2%

stra

in,

1h/d

ay,

0.01

67H

z fo

r 1

and

2 w

eeks

.

Hum

an u

mbi

lica

l vei

ns

(wal

l th

ickn

ess:

0.

75m

m,

oute

r di

amet

er:

6.75

0.

25m

m,

leng

th:

8.5c

m )

Wis

tar

Rat

bo

ne

mar

row

M

SC

bett

er c

ell

dist

ribu

tion

, m

ore

elon

gate

d ce

ll

mor

phol

ogy,

C

ell

prol

ifer

atio

n

, ul

tim

ate

stre

ss

, el

asti

c m

odul

us

(Abo

usle

iman

, R

eyes

et

al

. 20

09)

But

ler

2009

) C

usto

m-m

ade

pneu

mat

ic

cyli

nder

bi

orea

ctor

wit

h LV

DT

fo

r di

spla

cem

ent

mon

itor

ing

Cyc

lic

stre

tchi

ng 2

.4%

st

rain

, 0.

0033

Hz,

8h

/day

for

2 w

eeks

Type

I c

olla

gen

spon

ge

(94%

por

e vo

lum

e; 6

2 m

m

mea

n po

re

diam

eter

)

Mou

se

Mes

ench

ym

al

Ste

m

Cel

l

Type

I

coll

agen

,

line

ar s

tiff

ness

(B

utle

r, H

unte

r et

al.

2009

)

Che

n(20

10)

C

usto

m-m

ade

step

m

otor

bi

orea

ctor

Cyc

lic

stre

tchi

ng 1

0%

stra

in,

2h/d

ay,

1Hz

for

14 d

ays

Kni

tted

si

lk-c

olla

gen

spon

ge s

caff

old

(5cm

0.

5cm 0

.2cm

)

Hum

an

embr

yoni

c st

em c

ell

Col

lage

n I

, C

olla

gen

III

, E

pha4

,

Scx

,

Sox

9

, M

yosi

n

, In

tgri

n

1

, In

tgri

n

2

, In

tgri

n

1

,

Col

lage

n co

nten

t

, C

olla

gen

diam

eter

,

bett

er c

ell a

lign

men

t

(Che

n,

Yin

et

al

. 201

0)

Dor

oski

(2

010)

C

usto

m-m

ade

Lin

ear

mot

or

bior

eact

or

Cyc

lic

stre

tchi

ng 1

0%

stra

in (

5% o

ffse

t, 5%

am

plit

ude)

, 1H

z,

3h/d

ay,

1 7

14 a

nd 2

1 da

ys

Pol

y (e

thyl

ene

glyc

ol)-

base

d hy

drog

el

mat

eria

l ol

igo

(pol

y(et

hyle

ne

glyc

ol)

fum

arat

e)

(12.

5

9.5

1.6m

m)

MS

Cs

(PT-

2510

; L

onza

)

Col

lage

n I

, C

olla

gen

III

,

TN

C

, Te

nasc

in-C

(Dor

oski

, L

even

ston

et a

l. 20

10)

88

Tabl

e 1.

4 S

tudi

es u

sing

bio

reac

tor

syst

ems

for

tend

on/l

igam

ent e

ngin

eeri

ng

Sab

er (

2010

)

Lig

agen

L

30-4

C

(Tis

sue

Gro

wth

Te

chno

logi

es)

Cyc

lic

stre

tchi

ng

1.25

N,

1 cy

cle/

min

, 1h

/day

for

5 d

ays

Ace

llul

ar

rabb

it

hind

paw

te

ndon

(l

engt

h: 5

cm)

Rab

bit

teno

cyte

U

ltim

ate

tens

ile

stre

ss

(cl

osed

to

fl

eshl

y ha

rves

ted

tend

on)

, el

asti

c m

odul

us

(Sab

er,

Zha

ng

et a

l. 20

10)

Issa

(20

11)

C

usto

m-m

ade

bior

eact

or

Cyc

lic

stre

tchi

ng

2%

stra

in,

1h/d

ay,

0.01

67H

z fo

r 1

and

2 w

eeks

.

Hum

an u

mbi

lica

l ve

ins

(wal

l th

ickn

ess:

0.

41m

m)

wit

h W

hart

on’s

je

lly

mat

rix

as c

entr

al p

orti

on (

tota

l th

ickn

ess:

0.7

5mm

)

Wis

tar

rat

bone

m

arro

w

MS

C

Cel

l pr

olif

erat

ion

, te

nsil

e st

reng

th

(I

ssa,

E

ngeb

rets

on

et

al. 2

011)

Woo

n (2

011)

L

igag

en

L30

-1C

&

Lig

agen

L

30-4

C

(Tis

sue

Gro

wth

Te

chno

logi

es)

L30

-1C

: dy

nam

ic

load

ing,

10N

, 1h

/day

, 0.

0167

Hz

for

5 da

ys

L30

-4C

: dy

nam

ic

load

ing,

0.

625N

1.

25N

2.

5N,

1h/d

ay,

0.01

67 H

z fo

r 3

5 an

d 8

days

Ace

llul

ar h

uman

fle

xor

tend

on s

caff

olds

H

uman

D

erm

al

fibr

obla

st

Ult

imat

e te

nsil

e st

ress

, e

last

ic m

odul

us

(W

oon,

K

raus

et

al.

2011

)

89

1.2.6 Conclusion

The goal of tendon/ligament bioreactor design should be to create a construct with

similar microstructure and mechanical properties, as native tissue, using bioscaffolds

and autologous tenocytes. However, a tendon-like uni-orientated collagen structure

cannot be achieved without a mechanical stimulus within the culture environment.

Traditional culture techniques do not provide this PMS. The use of a bioreactor system

is able to bridge the gap between in vitro and in vivo systems by creating suitable

biochemical and biomechanical environments. Although it is clearly very difficult to

reproduce the in vivo microenvironmental conditions exactly within the bioreactor, the

goal is to mimic the in vivo biomechanical condition as closely as possible. The essential

components of a tendon/ligament engineering bioreactor are the actuating system and

culture chamber, which are responsible for the PMS and providing sterilized

environment for tissue culture. For better manipulation of the culture environment,

accurate stimulation and more precise reporting of mechanical maturation of engineered

tendon/ligament, an environmental control system, medium circulation system, monitor

and feedback system need to be included. As bioreactors nowadays are becoming more

focused on pre-clinical research, using these clinically in the future still presents a

substantial challenge. Tissue engineering requires substantial system optimization to

achieve a reproducible functional engineered tendon/ligament consistently. This process

is difficult and expensive in animal models, let alone in humans. However, there are

additional problems in moving from small animal models to humans related to the

physical size of human tissue samples. Issues of nutrient transport, cell source and

spatial heterogeneity of scaffold properties and cell stimulation become more prominent

in these larger tissues.

90

However, all the evidence suggests that by using bioreactors as described above it will

be possible to successfully produce engineered tendons and ligaments. When this occurs

we will be able to manufacture basic multi-chambers bioreactors with accurate and

appropriate patterns of PMS, which are affordable, and have low maintenance.

As one of the key elements of tissue engineering, biomechanical signal has been

introduced to generate engineered tendon for over a decade, and cyclic tensile strain is

commonly used. Positive effects, as summarized as figure 1.1, have been reported based

on various loading regimes with tensile strain ranged from 2% to 10% associated with

different frequencies and loading duration. These studies have demonstrated that

mechanical loading can improve cell proliferation, cell morphology, collagen I synthesis

and mechanical properties of engineered tendon. However, most of the studies only

examine one strain; the comparison of the effect caused by different mechanical strains

is poorly studied. The precise physiological levels of strain, frequency and duration to

affect such a response are not well understood.

1.3 Hypothesis and aims

1.3.1 Hypothesis

1.3.1.1 Overall hypothesis

I hypothesize that the biomechanical environment is essential for tendon

homeostasis. Further, the application of different mechanical stimulation on tendon

91

in a bioreactor system can induce different pathological changes and that the

biomechanical environment can help manipulate the repair mechanism of tendon

tissue.

1.3.1.2 Specific hypothesis

i. Different mechanical environments in a bioreactor have different impact on

tissue.

ii. There exists a mechanical strain regime that will maintain tendon homeostasis.

iii. There exists a mechanical strain regime that will lead to adverse changes in a

tendon

iv. A strain regime can be found which restores the functional and histological

properties of a tendon.

.

1.3.2 Aims

i. To design and construct a bioreactor system that is able to provide a

programmable biomechanical environment for in vitro tendon culture. (Study I,

Chapter 3)

ii. To evaluate the effect of different mechanical stimulation on tendon

homeostasis in a bioreactor system. (Study II, Chapter 4)

92

iii. To evaluate if the mechanical stimulation optimized in the previous study is

able to activate the self-repair mechanism to repair the tendon degeneration

caused by loading deprivation culture. (Study III, Chapter 5)

93

Chapter 2

Materials & Methods

94

CHAPTER 2 MATERIALS AND METHODS

There are a number of histological and biochemical analysis used throughout this thesis.

This chapter presents these techniques and methodology.

2.1 Material

All chemical reagents used during the course of this experiment were obtained through

the following manufacturers.

2.1.1 Tissue Culture reagents

REAGENT MANUFACTURER/SUPPLIER

DMEM/F-12 Gibco Life Technologies, USA

L-Glutamine 200mM 100X Gibco Life Technologies, USA

FBS (Fetal Bovine Serum) Invitrogen, Australia

Pen-Strep (Penicillin-streptomycin) Gibco Life Technologies, USA

Gentamicin Invitrogen, Auckland, New Zealand

95

2.1.2 Chemical reagents

REAGENT MANUFACTURER/SUPPLIER

Agarose LE, analytical grade Promega Corporation, USA

Baxter water Baxter Healthcare Pty.Ltd., Australia

Chloroform MERCK, Germany

Ethanol Hurst Scientific Pty. Ltd., Australia

Ethanol (molecular grade) Sigma-Aldrich, USA

Formaldehyde Sigma-Aldrich, USA

Paraformaldehyde (PFA) MERCK, Germany

TRIzol® reagent Invitrogen, Australia

Xylene Sigma-Aldrich, USA

Tissue embedding medium (Paraplast®

Regular) Paraffin wax Sigma-Aldrich, USA

Haematoxylin Sigma-Aldrich, USA

Eosin Sigma-Aldrich, USA

Diethyl pyrocarbonate Sigma-Aldrich, USA

proteinase K Qiagen, Germany

96

DPX Mountant Sigma-Aldrich, USA

2.1.3 Molecular products

PRODUCT MANUFACTURER/SUPPLIER

1kb DNA ladder Promega Corporation, USA

6x blue/orange loading dye Promega Corporation, USA

100bp DNA ladder Promega Corporation, USA

dNTPs (5nM) Promega Corporation, USA

Go-Taq® Green Master Mix 2x Promega Corporation, USA

M-MLV RT RNase (H-) Promega Corporation, USA

Oligo dT (100mM) Promega Corporation, USA

RT-PCR buffer (5x M-MLV RT) Promega Corporation, USA

SYBR® Safe DNA gel stain Invitrogen, USA

iQ™ SYBR® Green Supermix Bio-Rad Laboratories, USA

2.1.4 Oligonucleotide Primers

All oligonucleotide primers used in real-time polymerase chain reaction (PCR) were

97

purchased from GeneWorks Pty. Ltd. (Australia). The primers were reconstituted in

DEPC water to a concentration of 100M and stored at -20℃ until required.

PRIMER SEQUENCE (5’ to 3’) PRODUCT SIZE

COL1A2 Forward TTGCCCTTCCTTGATATTGC 241

COL1A2 Reverse CCTCTTTCGCCCACAATTTA

COL3A1 Forward AAGCCCCAGCAGAAAATTG 160

COL3A1 Reverse TGGTGGAACAGCAAAAATCA

MMP1 Forward GGCTCAGTTCGTCCTCACTC 246

MMP1 Reverse CAGGTCCATCAAAGGGAGAA

MMP3 Forward GCTTTGCTCAGCCTATCCAC 189

MMP3 Reverse ACCTCCAAGCCAAGGAACTT

MMP12 Forward ATGCCAGGGAGACCAGTATG 248

MMP12 Reverse AAAGCATGGGCTATGACACC

TIMP1 Forward CTTGTCATCAGGGCCAAGTT 213

TIMP1 Reverse TCCAGCGATGAGAAACTCCT

TIMP2 Forward GTGGACTCTGGGAACGACAT 223

98

TIMP2 Reverse AGTCACAGAGCGTGATGTGC

TGFB Forward TGCTTCAGCTCCACAGAGAA 181

TGFBReverse CCTTGCTGTACTGGGTGTCC

36B4 Forward ACCCAAAATGCTTCATCGTG 150

36B4 Reverse CAGGGTTGTTTTCCAGATGC

2.1.5 Antibodies

PRODUCT MANUFACTURER/SUPPLIER

Monoclonal anti-collagen type III Sigma-Aldrich, USA

Biotinylated goat anti-mouse IgG Sigma-Aldrich, USA

2.1.6 Commercial kits

PRODUCT MANUFACTURER/SUPP

LIER

RNeasy® minikit(250) QIAGEN, Germany

Liquid DAB+ Substrate Chromogen System DAKO, Denmark

99

In Situ Cell Death Detection Kit, POD Roche, USA

2.1.7 Other materials and equipment

MATERIAL/EQUIPMENT MANUFACTURER/SUPPLIER

AND EK-300i scientific weighing scale

A&D Mercury Pty. Ltd., China

Centrifuge 5810R Eppendorf, Germany

Corning high speed microcentrifuge Corning Incorporated, USA

Corning mini microcentrifuge Corning Incorporated, USA

CentriStarTM Cap centrifuge tubes:15ml and 50ml

Corning Incorporated, Mexico

Easypet® pipette dispenser Eppendorf, Germany

Forma Series II water jacker CO2 incubator

Thremo Scientific, Australia

Grant W-14 water bath Grant Instruments, UK

Homo-polymer, boil-proof microtubes: 0.6ml, 1.5ml and 2.0ml

Axygen® Scientific, USA

Hot plate model 209-1 IEC Australia Industrial Equipment & Control Pty. Ltd., Australia

100

KaterMaster catering foil Bunzl Limited, Australia

Laborlux S microscope Leitz, Germany

LabServTM incubator Biolabs (Aust) Pty. Ltd., Australia

Olympus CKX41 inverted microscope

Olympus Corporation, Japan

Nikon digital sight DS-5Mc Nikon Corporation, Japan

Nikon Eclopse E200 reflective microscope

Nikon Corporation, Japan

Parafilm “M” ® laboratory film Parafilm, USA

PCR-tube 0.5ml, thin wall Eppendorf, Germany

pH211 microprocessor pH meter Hanna Instruments, Australia

Ratek vortex mixer Rowe Scientific Pty. Ltd., Australia

Revco Elite Plus -80℃ freezer Thremo-Fisher Scientific, USA

Serological pipette: 2, 5, 10ml Sarstedt, France

SmartSpecTM 3000 Bio-Rad Laboratories, USA

SmartSpecTM 3000 cuvette quartz Bio-Rad Laboratories, USA

Tissue culture dish 100x20 mm Sarstedt, USA

Tissue culture flask: 25cm2 and Corning Incorporated, USA

101

75cm2

Leica TP 1020 tissue processor Leica, Germany

Sratfrost® microscope slides Knittel-Glaser, Germany

Superfrost ultra plus® slide Thremo Scientific, Australia

Microscope cover glass 22x50mm Hurst scientific Pty. Ltd., Australia

Leica 819 low profile microtome Blades

Leica, Germany

Jung Biocut 2035 Leica, Germany

Transferpipette 3.5ml Sarstedt, Germany

Veriti 96 well thermal cycler Applied BiosystemsTM, Singapore

CFX ConnectTM Real-time PCR detection system

Bio-Rad Laboratories, USA

Hard-Shell® PCR plates 96 well WHT/CLR

Bio-Rad Laboratories, USA

Labnet ExcelTM single electronic pipette

Labnet International, Inc.

2.1.8 Software

SOFTWARE MANUFACTURER/SUPPLIER

102

GraphPad Prism GraphPad Software Inc., USA

Image J National Institutes of Health, USA

NIS-C Elements Nikon Instruments Inc., USA

NIS-Elements BR 3.2 64bit Nikon Instruments Inc., USA

Solidworks 3D CAD Design Software

Solidworks Corp., USA

2.1.9 General solution

All general solution and buffers were prepared using MilliQ double distilled water,

unless otherwise indicated.

SOLUTION COMPOSITION

PBS 10x PBS stock solution:

30mM NaH2PO4, 70mM Na2HPO4, 1.3M NaCl

dissolved in MilliQ H2O

1x PBS pH 7.4

Dilution of 10x PBS with MilliQ H2O and adjusted to

pH 7.4

103

Paraformaldehyde 4% paraformaldehyde:

40g of Paraformaldehyde dissolved in prewarmed

0.75L MilliQ H2O ~ 60°C with stirring. Cover with foil

and stir with low heat, add 5-10 drops of 10M NaOH till

the solution become clear. Remove from heat and add

100mL 10x PBS and adjust pH to 7.4. Filter solution

through a 0.45m membrane filter before aliquoting and

storing at -20℃.

Hematoxylin working

solution (Harris)

Hematoxylin solution (Harris): 100g Postassium or

ammonium heat dissolves in 1L MilliQ H2O.

Add 50 ml of 10% alcoholic hematoxylin sol. and

heat to boil for 1min. Remove from heat and slowly add

2.5 g of mercuric oxide (red). Heat to the solution and

until it becomes dark purple color. Cool the solution in

cold water bath and add 20 ml of glacial acetic acid

(concentrated). Filter before use.

Eosin working solution Eosin stock solution: 1g Eosin Y dissolved in 100ml

MilliQ H2O.

Phloxine stock solution: 1g Phloxine B dissolved in

104

100ml MilliQ H2O

Eosin-Phloxine B working solution: Combine 100ml

Eosin stock solution, 10ml Phloxine stock solution, 780ml

95%[v/v] Ethanol and 4ml Glacial acetic acid

2.2 Tissue culture procedures

2.2.1 Isolation of rabbit Achilles tendon

Female New Zealand White rabbits (Oryctolagus cuniculus) aged between 14 and 15

weeks with body weight 3-5kg were sacrificed. The lower legs were wetted and

sterilized by 70%[v/v] ethanol. A transection was made between the Achilles tendon and

tibia to calcaneus using a sterilized scalpel. The Achilles tendon was then completely

exposed. A full length Achilles tendon was removed with part of the calcaneus and

approximately 1cm of the soleus muscle.

2.2.2 Culture of rabbit Achilles tendon

The Achilles tendons were briefly rinsed in 70%[v/v] ethanol and temporally stored in

serum-free DMEM-F12 supplemented with 50µg/ml Gentamicin after isolation from

rabbit. Before the tissue culture procedure, the Achilles tendons were washed in 40ml

1x PBS three times. In the loading deprivation culture group, tendons with muscle and

calcaneus were cultured in a T25 culture flask. In the dynamic culture group, the

105

Achilles tendons were cultured with muscle and calcaneus fixed in the bioreactor using

tissue clamps and hook. The whole culture procedure was processed in a standard

incubator at 37℃ and 5% CO2, and the culture media was DMEM/F-12 supplemented

with 10% fetal bovine serum and 50 µg/mL Gentamicin and changed every three days.

2.3 RNA extraction

2.3.1 Extraction of total RNA

Approximately 200mg of tendon tissue was dissected from the mid-substance of the

Achilles tendon, then washed with ice-cold 1x PBS for three times. The tissue was

placed into 5ml tube and snap frozen using liquid nitrogen and pulverized using mortar

and pestle. The tissue debris was then lysed using 1ml Trizol and homogenized by

pipetting in several times followed by 5 minutes incubation at room temperature. The

lysate was collected in a microcentrifuge tube, shaken by hand with 200m chloroform

for 15 seconds and incubated at room temperature for 2-3 minutes. The mixture was

spin at 12,000 × g for 15mins at 4℃, the RNA was separated in the aqueous phase. The

supernatant was transfer to a microcentrifuge tube, vortex mixed thoroughly with 1

volume of 70%[v/v] ethanol and transferred up to 700L to the spin cartridge (RNeasy®

minikit), and centrifuge at 12,000 × g for 15 seconds at room temperature. This

procedure was repeated several times till the entire samples has been processed. The

spin cartridge was then centrifuged at 12,000 × g for 15 seconds at room temperature

with 700L Wash Buffer I (RNeasy® minikit), and the flow-through and collection tube

106

were discarded. 500L Wash Buffer II (RNeasy® minikit) with ethanol was added into

the cartridge and centrifuged at 12,000 × g for 15 seconds at room temperature, and the

flow-through was discarded, this procedure was repeated once. The cartridge was

centrifuged at 12,000 × g for 1-2 minutes at room temperature to dry the membrane

with bound RNA; the collection tube was then discarded. Add 30–100L RNase-free

water to the center of the spin cartridge, incubated at room temperature for 1 minute

followed by 2 minutes centrifuge at ≥12,000 × g at room temperature to elute the RNA

from the membrane into the recovery tube. The purified RNA was stored at -80℃ until

required.

2.3.2 RNA concentration reading

200L of 1/50 dilution of purified RNA with Baxter water was prepared. The

SmartSpec™ 3000 (Bio-Rad) was blanked using 200L RNase-free water first, and the

samples was then injected into the Quartz Microvolume Cuvette, and read by the

Spectrophotometer. The absorbance ratio (A260/A280) provides an estimation of the

purity of the RNA. Generally, 1.8-2.0 is the acceptable range, where 2.0 represent

“pure”. All the purified RNA used in this study has A260/A280 ratio above 1.8.

107

2.4 Quantitative polymerase chain reaction (QPCR)

2.4.1 Reverse transcription of mRNA

A total 300ng of RNA was used to convert to equal amounts of cDNA using the

following protocol:

RNA 300ng

Oligo dT 0.25L

dNTP (5mM) 2L

RNase-free water final volume of 15L

Total volume 15L

The samples were incubated at 75℃ for 3 minutes, and then the following reagents

mixture was added:

5x MMLV RT-PCR Buffer 4L

RNase 0.5L

MMLV RT RNase 0.5L

Total volume 5L

108

The samples were then incubated at 42℃ for 60 minutes, followed by incubation at 92℃

for 10 minutes. The cDNA was stored at -20℃ until required

2.4.2 Quantitative polymerase chain reaction (QPCR)

Each of the QPCR reaction was performed using CFX ConnectTM Real-time PCR

detection system (Bio-Rad Laboratories, USA) and prepared according to the following

protocol:

iQ™ SYBR® Green Supermix 5L

cDNA 1L

Primer Forward (20M) 0.5L

Primer Reverse (20M) 0.5L

RNA free water 3L

Total volume 10L

The QPCR reaction primers were specified in 2.1.4 Oligonucleotide primer. The QPCR

conditions were carried out as per manufacturer’s instructions, each with 40 cycles.

Housing keeping gene 36B4 was used as internal control. A cycle threshold (Ct) value

109

was obtained from each reaction. The comparative 2-△△Ct method was used to calculate

the relative expression of each target gene. The RT-PCR were repeated in 3 individual

samples from each group and performed in triplicate in every RT-PCR setting. The

average expression levels were calculated and used for comparison.

2.5 General histology

2.5.1 Tissue preparation

The Achilles tendon was first washed by 1x PBS for three time after culture, and then

fixed using 35mL 4% PFA for overnight at 4℃ followed by 45 minutes wash with 1x

PBS. The samples were then placed in tissue cassettes, and put into the Leica TP 1020

processor, processed using the following protocol:

CHEMICAL CONDITION TIME

70%[v/v] Ethanol Vacuum 2 h

80%[v/v] Ethanol Vacuum 2 h

95%[v/v] Ethanol Vacuum 2 h

100%[v/v] Ethanol Vacuum 2 h

100%[v/v] Ethanol Vacuum 2 h

Xylene Vacuum 3 h

110

Xylene Vacuum 3 h

Xylene Vacuum 3 h

Paraffin Wax Vacuum 2h

Paraffin Wax Vacuum 2h

Paraffin Wax Vacuum 2h

TOTAL: 25h

After processing, the sample cassettes were removed from the tissue processor, and the

samples were taken out from the cassettes. A little wax was placed in the embedded tray,

allowing it to slightly cool. The sample then was placed down, and the tray was filled

with wax. The sample and tray was first cool down on the cooling station for 20 minutes

and then the embedded block was removed from the tray.

2.5.2 Haematoxylin and Eosin Staining

The paraffin-embedded tissue sample were cut into 5m thick section, placed onto the

microscope slider and processed by following protocol:

CHEMICAL TIME

111

Xylene 5min

Xylene 5min

100%[v/v] Ethanol 2min

90%[v/v] Ethanol 2min

70%[v/v] Ethanol 2min

Distill water 2min

Haematoxylin 2min

Washed in running water till clean

Lithium Carbonate 1min

Distill Water 1min

Eosin 30s

Washed in running water till clean

70%[v/v] Ethanol 2min

90%[v/v] Ethanol 2min

100%[v/v] Ethanol 2min

100%[v/v] Ethanol 3min

Xylene 5min

112

Xylene 5min

Mount in DPX

2.5.3 Histological scoring

Three tissue sections from different depth of the paraffin block of each sample were

selected. The H&E stained sections were viewed under the microscope, with at least 5

fields (20x magnifications) of each slice randomly selected for histological assessment

by 3 individuals blinded to group allocation. Four parameters, including fiber

arrangement, fiber structure cell roundness and cell density, were evaluated using a 4

point scoring scale from 0 to 3, where 0 being normal and 3 being severely abnormal.

The represented Figures of each score are shown as Figure 2.1 . The average score of

each parameter was calculated and used for comparison.

113

Figure 2.1 Histological scoring system modified from the study of Chen et al.(Chen, Yu et al. 2011)

2.6 Type III collagen immunohistochemistry

Paraffin-embedded sections were dewaxed and rehydrated using graded ethanol as

described in 2.5.2 Haematoxylin and Eosin Staining, and processed by the following

procedure.

CHEMICAL CONDITION TIME

3% hydrogen peroxide (H2O2)

Room temperature

10min

PBS 4min

5% fetal bovine serum in 1xPBS 30min

114

PBS 4min

monoclonal anti-collagen type III 1:100 damp chamber 4℃ overnight

PBS

Room temperature

3 x 10min

Biotinylated goat antimouseIgG 1:200 1h

PBS 3 x 10 min

DAKO Dab stain 3 min

Running water 30 min

Haematoxylin 1 min

running water 1min

70%[v/v] Ethanol 2min

90%[v/v] Ethanol 2min

100%[v/v] Ethanol 2min

100%[v/v] Ethanol 3min

Xylene 5min

Xylene 5min

Mount in DPX

115

2.7 TUNEL assay

Paraffin-embedded sections were dewaxed and rehydrated using graded ethanol as

described in 2.5.2 Haematoxylin and Eosin Staining, and processed by following

procedure.

CHEMICAL TIME

20g/mL proteinase K 15min

PBS 4min

TUNEL reaction mixture 1h

PBS 3 x 4min

convert-AP 30min

PBS 4min

DAKO Dab stain 5min

Running water 30 min

Haematoxylin 1 min

Running water 1min

70%[v/v] Ethanol 2min

90%[v/v] Ethanol 2min

116

100%[v/v] Ethanol 2min

100%[v/v] Ethanol 3min

Xylene 5min

Xylene 5min

Mount in DPX

Three tissue sections from different depth of the paraffin block of each sample were

selected. The TUNEL stained sections were viewed under the microscope, with at least

5 fields (20x magnifications) of each slice randomly selected for histological assessment

by 3 individuals blinded to group allocation. The positive staining cell and normal cell

were counted. The overall cell apoptotic rate was calculated and used for comparison.

2.8 Statistics analyses

All statistics have been analyzed using the prism 5.03 software package (GraphPad

Software 5.0, USA). Data are presented as mean ± SEM. Statistical analysis was

performed by one-way analysis of variance (ANOVA). P values less than 0.05 were

considered significant.

117

Chapter 3

Bioreactor Design and

Fabrication

This chapter is partly based on the paper

Programmable mechanical stimulation influences tendon homeostasis in a bioreactor system

Tao Wang, Zhen Lin, Robert E Day, Bruce Gardiner, Euphemie Landao-Bassonga, Jonas Rubenson, Thomas B. Kirk, David W. Smith, David G. Lloyd, Gerard Hardisty,

Allan Wang, Qiujian Zheng, Ming H Zheng

Biotechnology and Bioengineering

Volume 110, Issue 5, pages 1495-1507, May 2013

118

CHAPTER 3 (RESULTS): DESIGN AND CONSTRUCTION OF

BIOREACTOR

In this study we focus on the design of a tendon-engineering bioreactor, which is able to

provide programmable mechanical stimulation and a sterile culture environment. The

specific objective of the current study is to apply the design principles of mechanical

and electrical engineering to construct a functional bioreactor for tendon culture. To

fulfill this objective, a new bioreactor was designed using Solidworks 3D CAD software

(Solidworks Corp., Concord, MA), and constructed by an external mechanical vendor

under the supervision of the PhD candidate. This bioreactor system will assist in

evaluating the effect of various programmable mechanical stimulation regimes on

tendon homeostasis and repair in vitro, and ultimately enable optimization of the

mechanical stimulus for engineered tendon culture

3.1 Design criteria

To develop a functional bioreactor several design criteria were formulated:

1. The bioreactor must be able to process more than three independent culture

specimens simultaneously.

2. The system should have transparent covers, allowing visual monitoring.

3. The mechanical stimulation must be programmable.

4. The extension resolution should be better than ±100m with the load varying as

119

required to produce the set strain

5. The minimum expected load was 20N per sample, which is minimum 120N for the

bioreactor system.

6. The system must incorporate a means of holding the specimen ends to apply

stimulation without either damaging the specimen or allowing the construct to slip

relative to the grips.

7. The bioreactor must be able to operate inside a standard incubator, so all of the

electrical and mechanical components have to be able to function in a warm and

humid environment.

8. The parts in contact with the tissue culture should be autoclavable.

3.2 Material selection

In our design, the bioreactor is divided into two main components, the frame and the

culture chambers. The bioreactor is operated inside an incubator, and the culture

medium must be changed regularly. As this would require the operator to move the

bioreactor in and out of the incubator, the overall weight was a problem that needed to

be considered. For local sterilization, 70%[v/v] ethanol and UV treatment are

commonly used; therefore, both parts had to be chemical resistant and UV stable.

Moreover, the material used for culture chamber had to withstand a higher level of

sterilization such as autoclaving and also be resistant to humidity and chemical attack

from the culture medium.

120

In summary, the material used in the bioreactor had to meet the following requirements:

Good machinability.

Corrosion resistant (culture medium and warm humid air)

UV stable

Compatible with alcohol (up to 70%[v/v])

For the frame: as light as possible

For the culture chamber: compatible with autoclave sterilization

From the criteria listed above, a corrosion resistant metal would be preferred to other

materials due to its good machinability, high strength, UV stable and availability in the

market. We chose 6063 aluminum for the frame. Although its corrosion resistance is not

as good as stainless steel, it is relatively light compared to other metals. For the culture

chamber, AISI 316 stainless steel was chosen due to its good resistance to chemical

attack and humidity, moreover, it is suited to high pressure steam autoclaving, which is

the main sterilization method in our study.

3.3 Overall bioreactor design

After considering all the criteria and reviewing the published literature on both

custom-made and commercialized bioreactors, the design was finalized. The overall

121

dimensions of this bioreactor are 260mm×334mm×55mm, which is able to fit into a

standard incubator easily, and the overall weight is 6.9kg. Programmable mechanical

stimulation was generated with a single stepper motor and the motion transferred to the

cultured tissue through a ball-screw. This design is able to process dynamic cultures for

6 specimens simultaneously; and frequency, displacement and culture period can be

preprogrammed through a computer. Figure 3.1 shows the schematic illustration of the

bioreactor system, and Figure 3.2 shows the overview of the bioreactor.

Figure 3.2 Schematic illustration of the bioreactor system

122

Fig

ure

3.3

Des

ign

of 6

cha

mbe

rs b

iore

acto

r

123

Figure 3.4 Top view of bioreactor.

124

3.3.1 The Actuating System

The actuating system is the key of the whole bioreactor system, providing mechanical

stimulation to the cultured tissue. After reviewing the various designs of published

bioreactors, it was found that pneumatic actuators (Juncosa-Melvin, Shearn et al. 2006;

Nirmalanandhan, Rao et al. 2008; Nirmalanandhan, Shearn et al. 2008), linear

motors(Nguyen, Liang et al. 2009; Doroski, Levenston et al. 2010) and stepper

motor-ball screw combinations(Altman, Lu et al. 2002; Webb, Hitchcock et al. 2006;

Chen, Yin et al. 2010) are the most common actuating systems. Considering the position

accuracy and force output required in our study, a stepper motor-ball screw system was

chosen. There are two common types of stepper motors available in the market, 2-phase

and 5-phase stepper motors. A 2-phase motor only has 8 magnetic poles, whereas a 5

–phase motor has 10. The extra poles provide higher resolution, lower vibration, higher

acceleration and less synchronization loss than a 2-phase motor for little extra expense.

Therefore, a 5 phase stepper motor (PMC33A3, Oriental Motor, Japan) with 0.72°

resolution was selected for this bioreactor. To protect the stepper motor, an aluminum

cover was designed, which can eliminate the heat caused by the motor due to its good

heat conduction (Figure 3.3).

In order to transfer the rotation of the stepper motor to linear movement, a ball-screw is

necessary. Two common forms of ball-screw are commercially available, individual

ball-screws and guide actuator assemblies. Further machining and base design would be

required if a single ball-screw was chosen. The guide actuator consists of a

pre-assembled ball-screw, guide track and moving platform, which reduces the error

125

caused by further machining. Therefor the KR15 (THK, Japan) guide actuator was

chosen in our design. The lead of the KR15 is 1mm, and given the 500 steps per

revolution resolution of the stepper motor, this result in 1/500 or 2 micrometer per step

linear resolution.

3.3.2 The Culture Chamber

AISI 316 stainless steel was used for culture chamber construction due to its good

machinability and good resistance to chemical and high temperature. Multiple chambers

are necessary, considering that triplication is the minimum requirement for each

experiment. To balance the force distribution, the 6 chambers were uniformly located on

the base as shown as Figure 3.3. In order to simplify the operation, the culture chamber

was a separable individual design with 30mL capacity, shown as Figure 3.4. Each

culture chamber can be separated from the base and autoclaved. A removable glass

cover was designed to prevent contamination and enable better observation during the

culture period. A 1mm gap was reserved between the shaft of tissue clamps and the

glass cover, which allow the sufficient air exchange to the culture medium.

126

Figure 3.5 Design of culture chamber

3.3.3 Tissue Fixation

The method of tissue fixation is an important aspect of bioreactor designing needing

further consideration, as during mechanical stimulation in culture medium, sample

slippage can occur. Most bioreactor designs only have flat jawed tissue clamps that,

while suitable for flat membranes, cannot provide effective fixation with thick

cylindrical tissues such as tendon, leading to increased slippage. Therefore here two

fixation methods were designed for different sample types. The thin tissue or scaffold

can be clamped by the clamp cover and tightened using screws shown as figure (Figure

3.5A). Big and slippery tissues can be fixed by the tissue hook or tied to the tissue hook

using suture (Figure 3.5B).

127

Figure 3.6 Exploded view of the tissue clamps (A) and tissue hooks (B) assembly.

3.3.4 The Control System

The two main choices for control systems in other bioreactors are direct computer

control or a dedicated controller. Although programming a dedicated controller is more

complicated than using a computer, the dedicated controller is much more reliable.

Therefore, the selected control system was a programmable logic controller (MELSEC

FX1S, Mitsubishi, Japan) and a stepper motor driver (PMD03CA, Oriental Motor,

128

Japan). This driver is compatible with the PMC33A3 stepper motor. The motion control

program was written by the candidate himself using GX Developer (Mitsubishi, Japan),

and then downloaded to the PLC via RS232 cable. The PLC is able to provide

independent control on the stepper motor without any external requirements except

power.

3.4 Bioreactor setup

Before each experiment, the frame of the bioreactor was first sprayed using 70%[v/v]

ethanol and air dried in a laminar flow hood followed by sterilization under UV light for

30 minutes. The culture chambers and tissue clamps were double wrapped in paper bags

separately and autoclaved followed by drying in a 60℃ oven for 2 hours.

The bioreactor was placed in a standard incubator at 37℃and 5% CO2shown as Figure

3.6A. A 5-pin plug and cable connected the bioreactor to the control box outside of the

incubator through the standard hole at the back of the incubator (Figure 3.6B). Figure

3.7 illustrates the bioreactor setup in the incubator.

129

Figure 3.7 (A) Bioreactor in the incubator. (B) A five pin plug and cable connects the bioreactor to the control box through the hole at the back of the incubator.

130

Figure 3.8 illustration of bioreactor system setup in the incubator; the bioreactor is connected to the control box through a 5-core cable.

3.5 Discussion

In recent years, many studies of tendon engineering have been conducted using

custom-made bioreactor system, and the results have been promising(Juncosa-Melvin,

Shearn et al. 2006; Webb, Hitchcock et al. 2006; Androjna, Spragg et al. 2007;

Nirmalanandhan, Rao et al. 2008; Nirmalanandhan, Shearn et al. 2008; Abousleiman,

Reyes et al. 2009; Nguyen, Liang et al. 2009; Chen, Yin et al. 2010; Doroski, Levenston

131

et al. 2010; Saber, Zhang et al. 2010; Issa, Engebretson et al. 2011; Woon, Kraus et al.

2011). Several commercial tendon engineering bioreactors are available now, such as

Bose® ElectroForce® BioDynamic® and LigaGen system. To develop a functional

bioreactor system, several factors needed to be considered, including material,

mechanical design, tissue fixation method and operating environment. Compared with

commercial bioreactors in Bose and LigaGen, the current design has several advantages.

Firstly, our device can condition multiple samples simultaneously but separately. It can

be easily cleaned and sterilized in standard autoclaves. Being constructed from AISI 316

stainless steel it can go through many sterilization cycles without corrosion or

degradation. The device can produce a maximum force of 198N, which is sufficient

for most tendon scaffold and tendon tissue from small scale animals such as are

commonly used in tissue engineering research. Most commercial bioreactor systems

have problems in holding thick and slippery tissue during PMS, where as our novel

tissue hook design demonstrated excellent tissue fixation. Finally, the actuation system

is designed to operate inside a standard incubator. It is able to function in a humid

environment and the horizontal design is easy to move and will only occupy a small

space in the incubator. As the tendon’s main function is mechanical, this bioreactor

system is able to provide high-accuracy programmable mechanical stimulation (PMS),

enabling different loading regimes including displacement, frequency and stimulation

period. Once programmed it is able to operate reliably for weeks without computer

support.

132

Chapter 4

Programmable mechanical

stimulation influences tendon

homeostasis in a bioreactor

system

This chapter is partly based on the paper

Programmable mechanical stimulation influences tendon homeostasis in a bioreactor system

Tao Wang, Zhen Lin, Robert E Day, Bruce Gardiner, Euphemie Landao-Bassonga, Jonas Rubenson, Thomas B. Kirk, David W. Smith, David G. Lloyd, Gerard Hardisty,

Allan Wang, Qiujian Zheng, Ming H Zheng

Published in

Biotechnology and Bioengineering

Volume 110, Issue 5, pages 1495-1507, May 2013

133

CHAPTER4 (RESULTS): PROGRAMMABLE MECHANICAL

STIMULATION INFLUENCES TENDON HOMEOSTASIS IN A

BIOREACTOR SYSTEM

4.1 Abstract

Identification of functional programmable mechanical stimulation (PMS) on tendon not

only provides the insight of the tendon homeostasis under physical/pathological

conditions, but also guides a better engineering strategy for tendon regeneration. The

aims of the study are to design a bioreactor system with PMS to mimic the in vivo

loading conditions, and to define the impact of different cyclic tensile strains on tendon.

Rabbit Achilles tendons were loaded in the bioreactor with/without cyclic tensile

loading (0.25Hz for 8 hours/day, 0~9% for 6 days). Tendons without loading lost their

structural integrity, as evidenced by disorientated collagen fiber, increased type III

collagen expression, and increased cell apoptosis. Tendons with 3% of cyclic tensile

loading had moderate matrix deterioration and elevated expression levels of MMP-1, 3

and 12, whilst loading regime of 9% caused massive rupture of collagen bundles.

However, 6% of cyclic tensile strain was observed to maintain the structural integrity

and cellular function. Our data indicated that an optimal PMS exists to maintain tendon

homeostasis, although this is only a narrow range of tensile strain that can induce this

action. The suggested clinical impact of this study is that optimized eccentric training

program, based around targeting specific tendon strain, could be developed to achieve

beneficial effects on chronic tendinopathy management.

134

4.2 Introduction

Tissue engineering methods have great potential in the repair of damaged tendons.

However, these methods are still very much in their infancy, due to the lack of an

integrated understanding of the key conditions required for successful tendinopathy

intervention and repair. Recent tendon research has focused on the development of

tendon tissue engineering, with the aim of constructing a bio-substitute material for

tendon/ligament repair(Juncosa-Melvin, Shearn et al. 2006; Sahoo, Ouyang et al. 2006;

Webb, Hitchcock et al. 2006; Chen, Willers et al. 2007; Chen, Qi et al. 2008;

Nirmalanandhan, Rao et al. 2008; Saber, Zhang et al. 2010; Stoll, John et al. 2010;

Peach, James et al. 2012; Wang, Gardner et al. 2012; Woon, Farnebo et al. 2012); This

strategy involves an integration of technologies from cell biology, material science and

mechanical engineering.

Engineered bioreactor systems have been demonstrated to be effective tools for

constructing various biological tissues and organs such as muscle(Dennis, Smith et al.

2009), liver (Catapano, Patzer et al. 2010) and bone(Rauh, Milan et al. 2011; Rauh,

Milan et al. 2012). Several bioreactor systems have been developed specifically for

engineered tendon/ligament formation (Altman, Lu et al. 2002; Juncosa-Melvin, Shearn

et al. 2006; Webb, Hitchcock et al. 2006; Nirmalanandhan, Rao et al. 2008;

Nirmalanandhan, Shearn et al. 2008; Issa, Engebretson et al. 2011). Owing to force

transmission being the primary function role of tendon/ligament, mechanical

stimulation plays an essential role in the maintenance of tendon structure and

135

function(Hannafin, Arnoczky et al. 1995). Achilles tendons endure tensile loading in

vivo during daily physical activities. More complex structures, like the anterior cruciate

ligament, undergo both tensile and torsion loading during weight-bearing knee

flexion(Li, Defrate et al. 2005). To restore tendon function after injury, engineered

tendon constructs should be compatible with the in vivo environment both biologically

and mechanically(Calve, Lytle et al. 2010). Cell-scaffold constructs may be ‘exercised’

in an artificial mechanical environment before implantation. Indeed, programmable

mechanical stimulation (PMS) is considered to be one of the key components for

tendon/ligament forming bioreactor system (Altman, Lu et al. 2002; Juncosa-Melvin,

Shearn et al. 2006; Webb, Hitchcock et al. 2006; Nirmalanandhan, Rao et al. 2008;

Nirmalanandhan, Shearn et al. 2008; Issa, Engebretson et al. 2011). In this study, a

bioreactor system was designed to apply different cyclic tensile strain on isolated rabbit

Achilles tendon to establish an ex vivo model (see previous Chapter). The present aim is

to evaluate the effect of different mechanical stimulation on tendon tissue using this

developed bioreactor system.

The mechanical stimulation of engineered tendon constructs can result in different

structures. Non-loading of engineered neo-tendons results in loose, disorganized

matrices with reduced collagen quantity(Jiang, Liu et al. 2011). Static mechanical stress

has been shown to improve the fiber arrangement of engineered tendon constructs, but

did not produce a compacted tendon-like collagen structure in vitro(Chen, Willers et al.

2007). More promisingly, engineered tendon under cyclic tensile strain exhibited a

well-organized collagen fiber arrangement and structure and enhanced tensile strength,

136

suggesting that PMS is preferable for engineered tendon formation(Jiang, Liu et al.

2011). Even though the importance of providing PMS in tendon bioreactor has been

recognized previously (Brown and Carson ; Wang, Gardner et al. 2012), the optimal

cyclic tensile stimulation parameters have not been defined. It is generally assumed that

mimicking the dynamic mechanical loading created by daily physical activity will give

the optimal loading regime in tendon bioreactor system, but this is not yet proven.

The goals of this study were to investigate the effect of different mechanical stimulation

regimes on tendon integrity. I hypothesized that there is an optimum range of

mechanical stimulation required for tendon homeostasis and that under or

over-stimulation would reduce tendon integrity.

4.3 Material and Method

4.3.1 Tissue harvest and distribution

Achilles tendons were dissected from the hindlimbs of female New Zealand White

rabbits (Oryctolaguscuniculus) aged between 14 and 15 weeks with body weight 3-5kg.

The tissue was briefly rinsed in 70%[v/v] ethanol, washed in 1x phosphate buffered

saline (PBS) 3 times, and submerged in serum free Dulbecco’s modified Eagle’s

medium (DMEM)/F-12 (GIBCO, Invitrogen, Auckland, New Zealand) supplemented

with 50 µg/mL Gentamicin before further processing.

137

In total 24 Achilles tendons from 12 rabbits were allocated to 3 groups (Figure 4.1A): a

loading deprivation group (8), a dynamic culture (cyclic tensile strain) group (12) and a

control group (4). In the loading deprivation group, the Achilles tendons were cultured

in petri dish in growth medium (DMEM/F-12, 10% fetal bovine serum and 50 µg/mL

Gentamicin) without mechanical loading for either 6 days (4 tendons) or 2 weeks (4

tendons). In the dynamic strain group, the tendons were fixed on the tissue hooks using

suture and cultured under 3%, 6% or 9% cyclic tensile strain for 6 days (8 hours per day,

0.25Hz, Figure 4.1A). The stimulating cycle was as demonstrated in Figure 4.1B.This

regimen was based on previously published studies and data on rabbit activity levels

(West, Juncosa et al. 2004). The samples were sliced into two half, ones was fixed in 4%

paraformaldehyde overnight for histological analysis and the other half directly

subjected to RNA extraction. The control group was made up of 4 tendons directly

dissected from the animal and immediately fixed.

Figure 4.1 (A) Number and distribution of samples for each of the 12 rabbits. (B) Daily

mechanical stimulation pattern used in this study. (C) Mechanical input signal from the control

system

138

4.3.2 Histological preparation and assessment

Refer to 2.5

4.3.3 TUNEL assay

Refer to 2.7

4.3.4 Immunostaining for type III collagen

Refer to 2.6

4.3.5 Quantitative Real-time polymerase chain reaction (Q-PCR)

Refer to 2.3 and 2.4

139

Table 4.5 Rabbit specific primer sequences for the genes used.

Target Genes Forward (5’ to 3’) Reverse (3’ to 5’)

COL1A2 TTGCCCTTCCTTGATATTGC CCTCTTTCGCCCACAATTTA

COL3A1 AAGCCCCAGCAGAAAATTG TGGTGGAACAGCAAAAATCA

MMP1 GGCTCAGTTCGTCCTCACTC CAGGTCCATCAAAGGGAG

AA

MMP3 GCTTTGCTCAGCCTATCCAC ACCTCCAAGCCAAGGAACTT

MMP12 ATGCCAGGGAGACCAGTATG AAAGCATGGGCTATGACACC

36B4 ACCCAAAATGCTTCATCGTG CAGGGTTGTTTTCCAGATGC

4.3.6 Statistical analyses

Refer to 2.8

4.4 Results

4.4.1 Histological examination of tendon undergoing different cyclic tensile

strains

Compared to the native tendon (Figure 4.2 A), the collagen fiber arrangement changed

from parallel to moderately wavy after 6 days in the absence of applied loading. Cell

density was increased with slightly rounded nuclei. After 2 weeks without loading,

progressive extracellular matrix (ECM) disruption was observed. Collagen fiber

140

arrangement had no discernible pattern. Fragmentation of collagen fibers and rounded

tenocyte nuclei were observed microscopically (Figure 4.2C). Histological assessment

showed a time dependent increasing grading in fiber arrangement, fiber structure and

cell roundness, indicating progressive divergence from the native (control) tissue. Cell

density was also increased after 6 days without loading (Figure 4.2D-G).

141

Figure 4.2 Loading deprivation culture group. (A) Native Achilles tendon structure (day 0). (B) Achilles tendon histology after 6 days loading deprivation culture with minor wavy collagen structure. (C) Achilles tendon histology after 2 weeks culture displays disorganized collagen

structure with round tenocytes. (D) Fiber arrangement, scale of 0-3, 0 represents compacted and parallel, and 3 represents no identifiable pattern. (E) Fiber structure, scale of 0-3, 0 represents

continuous, long fiber, and 3 represents severely fragmented. (F) Cell roundness, scale of 0-3, 0 represents long spindle shape, and 3 represents severe rounding. (G) Cell density, scale of 0-3, 0

represents normal pattern, and 3 represents severely increase Result are expressed as the mean±S.D. (*p<0.05, **p<0.01, **p<0.001). Scale bar=50 micrometers.

142

In the dynamic culture (cyclic tensile strain) group, histology revealed that the 3% strain

stimulated group had slightly disrupted extracellular matrix structure and significant

cellular morphology changes (Figure 4.3B). These tendon samples graded 1.5 in fiber

arrangement, 1.1 in fiber structure and 1.3 in cell roundness (Figure 4.3E-G). However,

in the tendons that underwent 6% strain, the extracellular matrix orientation and cellular

morphology appeared normal (Figure 4.3A, C). That is, there was no statistical

difference of the histology scoring between the 6% group and native tissue (Figure

4.3E-H). In the 9% strain stimulated group, the tendon samples were found to be

partially torn and graded 1.45 in fiber structure and 0.8 in fiber arrangement (Figure

4.3E, F). Increased cell number and slightly rounded tenocyte nuclei were observed

around the rupture site (Figure 4.3D, G, H).

143

Figure 4.3 6 days dynamic culture group. (A) Native Achilles tendon structure. (B) 3% stimulated group with wavy collagen structure and minor rounded tenocytes. (C) 6% stimulated group with similar

morphology compared with the native group. (D) 9% stimulated group with severe tear collagen structure, and rounded nuclei around the tear region (arrow). (E) Fiber arrangement, scale of 0-3, 0 represents compacted and parallel, and 3 represents no identifiable pattern. (F) Fiber structure, scale of 0-3, 0

represents continuous, long fiber, and 3 represents severely fragmented. (G) Cell roundness, scale of 0-3, 0 represents long spindle shape, and 3 represents severe rounding. (H) Cell density, scale of 0-3, 0

represents normal pattern, and 3 represents severely increase. Result are expressed as the mean±S.D. (*p<0.05, **p<0.01, **p<0.001). Scale bar=50 micrometers.

144

4.4.2 Impact of cyclic tensile strain on apoptosis of tenocytes

To investigate the impact of mechanical force on the apoptosis of tenocytes, I used the

TUNEL assay to examine the cellular apoptosis rate in specimens. In the native tendons,

the average cell apoptosis fraction (as a proportion of total cell number) was less than

10% (Figure 4.4A, G). The apoptosis fraction progressively increased over time when

the tissues were cultured without loading (35% in 6 days and 95% in 2 weeks) (Figure

4.4G). Cell viability in 3% and 6% stimulated Achilles tendon was similar to that in

native tendon (Figure 4.4D, E) without statistical significance (Figure 4.4G), whilst cell

apoptosis rate increased to 45% when the stimulation strain reach 9% (Figure 4.4F, G).

145

Figure 4.4TUNEL Assay (A) Native Achilles tendon with less than 10% cell apoptosis. (B) 6 days loading deprivation group has increased apoptotic tenocytes pointed out by arrows. (C) In 2 weeks

loading deprivation group, almost all tenocytes are under apoptosis (D) 3% stimulated group with minor increased apoptotic cells (arrow). (E) 6% stimulated group with similar cell apoptosis rate as native

sample (F) 9% stimulated group with massive cell apoptosis around the damaged site (pointed by arrows). (G) Quantified cell apoptosis fraction of each group. Result are expressed as the mean±S.D. (*p<0.05,

**p<0.01, **p<0.001). Scale bar=50 micrometers

146

4.4.3 Type III collagen turnover in Achilles tendon under dynamic culture

To investigate the extracellular matrix turnover in tendon undergoing different cyclic

tensile strain, type III collagen immunohistochemistry was done. No positive type III

collagen was observed in the core of healthy tendon using immunochemistry (Figure

4.5A). After 6 days of either no loading or 3% strain, the Achilles tendons expressed

only minor and moderate collagen type III respectively (Figure 4.5B,C), whilst

relatively higher positive staining were found in the 9% strain group (Figure 4.5E). No

type III collagen expression was observed in the 6% cyclic strain stimulated group

(Figure 4.5D), which was similar to the native tendon.

Figure 4.5Collagen type III immunohistochemistry (A) Native Achilles tendon with minor collagen type III expression. (B) Loading deprivation group has minor increased collagen type III expression and some positive staining tenocytes pointed out by arrow. (C) 3% stimulated group with increased collagen

type III expression. (D) 6% stimulated group with similar collagen type III expression as native sample (E) 9% stimulated group with massive increased collagen type III expression and positive staining tenocytes

(pointed by arrows). Scale bar=50 micrometers

147

4.4.4 Impact of cyclic tensile strain on ECM remodeling gene expression

Q-PCR was carried out to examine the gene expression of type I (COL1A2), type III

collagen (COL3A1) and MMP1, 3, 12 (matrix metalloproteinases). As shown in Figure 8,

the tissues subjected to 6% cyclic tensile strain had the highest expression of COL1A2

(Figure 8A), whereas 3% stimulated group had the lowest expression. Type III collagen

(COL3A1) was suppressed by minor and moderate loading (3% and 6%), yet

upregulated in the 9% strain stimulated group (Figure 8B). Gene expression of MMP1,

3 and 12 were highly upregulated by 3% strain stimulation compared to the other

groups.

Figure 4.6Quantitative PCR results. A: COL1A2 expression in 6% stimulated group is significantly higher than other groups. B: COL3A1 expression in 9% stimulated group is

significantly higher than other groups. Three percent and 6% group had much lower COL3A1 expression compared to the other groups. C–E: Three percent stimulated group has significantly high

expression on MMP1 MMP3 and MMP12, respectively. Result is expressed as the mean±S.D (*p<0.05, **p<0.01, **p<0.001)

148

4.5 Discussion

A previous study has suggested that engineered neo-tendons in a non-loading

environment result in loose, disorganized and less collagen-rich matrices (Jiang, Liu et

al. 2011). Static mechanical loading has been shown to improve the fiber arrangement

of engineered tendon constructs, but did not produce a compacted tissue structure in

vitro (Chen, Willers et al. 2007). Engineered neo-tendons with PMS had a

well-organized fiber arrangement and structure and enhanced tensile strength,

suggesting that dynamic mechanical loading is preferable for engineered tendon

formation (Jiang, Liu et al. 2011). Although PMS has been recognized as a critical

regulator for tendon formation (Brown and Carson et al. 1999), the optimal PMS

parameters are yet to be defined.

In this study we designed a uni-axial bioreactor system with a programmable

stimulation regimen. Achilles tendons from rabbits were used to evaluate the impact of

different PMS conditions on tendon tissue. Our result showed that 3% cyclic tensile

strain in the bioreactor did not prevent matrix deterioration, whilst at 6% cyclic tensile

strain structural integrity and cellular function was maintained. At 9% cyclic tensile

strain there was massive rupture of the collagen bundles. However, these results are for

an already mature tendon and the loading regimes may be different for engineered

constructs at various stages of growth. Furthermore, the frequency of loading may also

have an important role in maintaining tendon homeostasis and this need to be better

qualified by in vivo measurements of Achilles tendon loading. Nevertheless, the current

149

results demonstrate that PMS can be achieved in this bioreactor, and that an optimal

cyclic tensile loading (around 6% cyclic tensile strain) is required to maintain healthy

tendon homeostasis, at least over the time period of the cultures used in this study.

It has been well known that tensile mechanical loading is essential for tendon integrity,

and loading deprivation can cause tendinopathy-like morphology such as disorientated

collagen fibers and rounded cell nuclei(Hannafin, Arnoczky et al. 1995). Yet the specific

effect of different loading conditions on tendon has not been well elucidated. To study

the effect of PMS on tendon/ligament, both in vitro and in vivo models have been used.

The Flexcell® system has been applied to examine the in vitro gene expression profiles

of tenocytes under tensile or shear force (Skutek, van Griensven et al. 2001; Wang, Jia

et al. 2003; Yang, Im et al. 2005). These systems can precisely control the mechanical

regimen and monitor the cell behaviors, but only use a monolayer of cells, not an

organized tissue. Cells isolated from their extracellular niche and cultivated in a

monolayer may not necessarily replicate their original behavior in vivo. In vivo models,

such as Achilles tendon in rats or rabbit, have been commonly used as study models.

Tendon loading is mainly induced by electrical stimulation on muscle(Nakama, King et

al. 2005; Asundi, King et al. 2008), treadmill training on animal and partial tendon

transaction(Smith, Sakurai et al. 2008). However, mechanical stimulation on the tendon

varies between individuals, making precise quantification of load next to impossible. In

this study, we have demonstrated a bioreactor system that not only can be used as an

device for tendon/ligament engineering, but also serves as an ex vivo culture model for

the study of tendon biology. Although the biomechanics of Achilles tendon is more

150

complicated under in vivo environment (West, Juncosa et al. 2004), by applying

programmable mechanical stimulation, the effect of different cyclic tensile strain on

tendon tissue could be systematically analyzed on a tissue level.

The present study has shown for the first time that only a narrow range of PMS can

stimulate the anabolic effect of tendon tissue. This suggests a proposed model

describing the effect of mechanical loading, from under-load to over-load, on tendon

homeostasis. This model is illustrated in Figure 4.7 in which the black curve presents

the typical mechanical response caused by increasing tensile strain in tendon. The

breakage of collagen fibers increases with increasing strain, leading to eventual rupture

at the failure point. The red curve, on the other hand, proposes the biological repair

response of the tendon induced by tensile strain. Mechanical loading is required for the

biological activity of tenocytes, yet the regime needs to be within a certain range to

promote matrix product and net tissue repair. By combining these two curves, three

crossover areas are seen, which can be divided according to their metabolic status into:

A (anabolic) zone and C (catabolic) zones. In the A zone, where the tensile loading is

within a certain range, the biological repair response caused by PMS exceeds the

induced damage. In the C zone, there is either too little load to stimulate matrix

production or too much mechanical damage caused by excess tensile loading. In our

study the 3% strain stimulated group (under-loaded) was located in C zone. When the

tendons were subjected to insufficient tensile strain, early tendinopathologic changes

(including increased cellular number, type III collagen expression and disorientated

collagen fibers) could be observed. These morphological changes may be attributed by

151

the up-regulation of MMPs expression in tendon under 3% tensile loading. MMPs are

zinc-dependent endopeptidases that able to degrade all the components of the

extracellular matrix.MMP-1 is primarily responsible for the cleavage of collagen,

including type I, II and III(Krane, Byrne et al. 1996). MMP-3 and MMP-12 are other

members of the MMP family that have a wide range of substrate specificities. Moreover,

MMP-12 is able to activate MMP-3, and MMP-3 has the ability to activate other MMPs

such as MMP-1, MMP-7 and MMP-9 (Knauper, Lopez-Otin et al. 1996; Chen 2004).

Studies have shown that decreased mechanical force on tendon fiber would lead to the

deprivation of cytoskeletal tension in tenocytes and alter the mechanostransduction in the

cells (Arnoczky 2008). Insufficient loading of tenocyte can initiate the degenerative

cascade by expressing a pattern of catabolic gene, including MMPs(Lavagnino,

Arnoczky et al. 2003; Arnoczky, Tian et al. 2004; Lavagnino, Arnoczky et al. 2006;

Arnoczky, Lavagnino et al. 2008; Egerbacher, Arnoczky et al. 2008).In this study, we

have confirmed that under the underloading condition, the phenotype of the tenocytes

changed from spindle to round, and the expression levels of catabolic genes, such as

MMP-1, 2, and 12, were upregulated. The elevated catabolic factors may contribute to the

further micro-damage of the collagen fiber. Conversely, although the tendons in the 6%

strain stimulated group were subjected to more mechanical damage, the repair

mechanism was able to maintain tissue homeostasis (i.e. there is a narrow ‘net anabolic

window’ between two net catabolic zones). In the 9% group (overloaded), mechanical

damage overwhelmed the tendon tissue repair and caused partial rupture, which is

consistent with our proposed model.

152

The identification of an optimal PMS range for tendon will not only guide us to a better

tissue-engineering pathway for tendon regeneration, but also provide an insight into

tendon homeostasis under both physiological and pathological conditions. Clinically,

chronic tendinopathy of Achilles tendon is usually conservatively managed by eccentric

overload training. The calf muscle is eccentrically loaded with the knee held straight or

the soleus muscle is maximally loaded with the knee bent(Alfredson, Pietila et al. 1998).

Although positive impacts of eccentric overload training on tendinopathy treatment

have been recorded, the outcomes have varied widely (Stanish, Rubinovich et al. 1986;

Alfredson, Pietila et al. 1998; Silbernagel, Thomee et al. 2001; Alfredson and Lorentzon

2003; Fahlstrom, Jonsson et al. 2003; Roos, Engstrom et al. 2004; Shalabi,

Kristoffersen-Wilberg et al. 2004). In the study of Alfredson et al.,(Alfredson, Pietila et

al. 1998) the overall improvement in pain score was up to 94%, though in the study of

Silbernag el et al. reported only 29% (Alfredson and Lorentzon). Although the

mechanism of eccentric overload training on tendinopathy management remains unclear,

the present study has provided a possible explanation of the mechanism of eccentric

loading on chronic tendinopathy. The high variation of clinical outcomes might be due

to poor control of the loading regimen. Given that there is only a narrow range of tensile

strain where eccentric exercise could provide an anabolic effect on tendon, clinical

treatment protocols will need to accurately control the magnitude of load applied

through the injured tendon.

Finally, it is noted that nutrient supply might be altered by advective transport

associated with mechanical stimulation. Although a previous study on canine flexor

153

tendons suggested that mechanical stimulation did not alter small nutrient uptake

compared to static culture, and the tendon repair improvements derived from passive

motion is not related to an increase of nutrient transport (Hannafin and Arnoczky 1994),

under different loading regimes, or in different tendon types, this may not be the case.

Moreover, the transport of large molecules, such as growth factors and matrix

components, may be affected by mechanical stimulation, as has been shown in studies

in related tissues (Bonassar, Grodzinsky et al. 2000; Zhang, Gardiner et al. 2007).

Therefore, the tendon stimulated by dynamic mechanical loading in this study might be

attributed to the enhanced nutrient transport, compared to the static control.

Consequently it is difficult to isolate the direct effect of mechanical stimulation on

tendon response from potential enhanced advective transport processes. Further study

would be required to measure the nutrition transport in tendon under PMS

154

Figure 4.7 Matrix damage (black curve) and production (red broken) affected by mechanical stimulation. A zone indicates a small range of anabolic effect caused by mechanical stimulation. Matrix

production stimulated by mechanical loading overcomes the matrix damage that the tendon is able to maintain its structural integrity. C zone indicates two catabolic zones that matrix damage caused by

mechanical stretching overcomes the matrix production result in tendinopathy.

4.6 Conclusion

We have developed and validated a uniaxial bioreactor system. The results in our study

showed tendons subjected to 6% of cyclic tensile strain were able to maintain their

structural integrity and cellular function. Our data suggested that tendon is a

mechano-sensing tissue where the net anabolic action of cyclic tensile strain only occurs

over a narrow range. In our study, mechanical underloading (3% cyclic tensile strain)

and overloading (9% cyclic tensile strain) produced a net negative impact on the rabbit

Achilles tendon tissue.

155

Chapter 5

Cyclic mechanical stimulation

rescues rabbit Achilles tendon in

a bioreactor system

This chapter is based on the paper

Cyclic mechanical stimulation rescues rabbit Achilles tendon from degeneration in a bioreactor system

Tao Wang, Zhen Lin, Ming Ni, Robert E Day, Bruce Gardiner, Euphemie Landao-Bassonga, Jonas Rubenson, Thomas B. Kirk, David W. Smith, David G. Lloyd,

Yan Wang, Qiujian Zheng, Ming H Zheng

(Submitted to Tissue Engineering Part A)

156

CHAPTER 5 (RESULTS): CYCLIC MECHANICAL

STIMULATION RESCUES RABBIT ACHILLES TENDON FROM

DEGENERATION IN A BIOREACTOR SYSTEM

5.1 Abstract

Mechanical stimulation has been identified as an essential factor for maintaining tendon

homeostasis, and various studies have reported that Achilles tendons subjected to

long-term in vivo immobilization display pathological changes while mechanical

properties degrade. Our previous study showed that 6% strain cyclic mechanical

stimulation is able to maintain the tendon homeostasis in the bioreactor system.

Therefore, in this study, we hypothesize that 6% strain cyclic mechanical stimulation

can rescue morphological and mechanical degeneration similar to tendinopathy ex vivo.

New Zealand white rabbits and divided into 4 groups: native control group, 6 days and

12 days loading deprivation groups, and stimulation group. In loading deprivation

groups, the Achilles tendons were cultured without loading for either 6 days or 12 days.

In stimulation group, tendons were subject to 6% cyclic strain at 0.25Hz for 8 h/d for 6

days after 6 days no loading culture. Tendons in both loading deprivation groups

developed progressive histological degradation, increased tenocytes apoptosis and

weakened mechanical properties after loading deprivation. However, in stimulation

group, the histological and mechanical properties substantially improved. Our results

indicate that a 6% strain cyclic mechanical stimulation was able to rescue early-stage

pathological changes from loading deprivation and restore Achilles tendon ex vivo.

157

5.2 Introduction

Clinical findings have demonstrated tendon degeneration and compromised

biomechanical properties occurs after load deprivation, including in situations ranging

from bedridden patients to astronauts(Reeves, Maganaris et al. 2005). Studies have

reported that early-stage pathological alterations in Achilles tendon following long term

immobilization in both human and rat (Larsen, Forwood et al. 1987; Nakagawa, Totsuka

et al. 1989; Kannus, Jozsa et al. 1997; Trudel, Koike et al. 2007). The previous study

(Chapter 4) observed that rabbit Achilles tendon after load deprivation exhibited

pathological changes that included: (i) wavy and ruptured collagen fibers, (ii) rounded

cell morphology and (iii) increased cell apoptosis (Wang, Lin et al. 2013). Accumulated

evidence, including Chapter 4, shows that in order to maintain normal tendon

homeostasis, mechanical stimulation is required(Arnoczky, Lavagnino et al. 2007;

Egerbacher, Arnoczky et al. 2008; Smith, Sakurai et al. 2008; Thornton, Shao et al.

2008; Wang, Gardiner et al. 2013). The key question is can functional tendon be

restored after pathological changes due to unloading?

In the previous study, appearing in Chapter 4, it was found that a relatively narrow

range of mechanical stimulation is able to maintain rabbit Achilles tendon in the

bioreactor, whereas too little strain results in protease degradation while too much strain

results in mechanically induced fatigue damage. For Achilles rabbit tendon it was found

that a 6% strain set in a cyclic mechanical stimulation device is able to maintain tendon

integrity in a bioreactor system. This current study examined if it was possible to rescue

158

the early pathological changes caused by 6 days of loading deprivation, similar to that

observed in tendinopathy, by then applying 6% strain cyclic loading for a further 6 days.

The positive effects of mechanical loading following load deprivation reported here

have implications for all kinds of situations involving tendon degeneration, ranging

from physical treatment following immobilization to the tendon healing process

following surgery, and even to planning the most effective exercises for astronauts.

5.3 Material and method

5.3.1 Tissue preparation and programmable mechanical stimulation

Full-length Achilles tendons, including the bone-tendon and tendon-muscle junction

were dissected from the hindlimbs of 15 weeks old female New Zealand white rabbits

(Oryctolagus cuniculus). Animals were obtained immediately post-euthanasia.

By-products from tendon samples were washed by 1x phosphate buffered saline (PBS)

3 times, and submerged in serum free Dulbecco’s modified Eagle’s medium

(DMEM/F-12; GIBCO, Invitrogen, Auckland, New Zealand) supplemented with 50

µg/mL Gentamicin (Invitrogen, Auckland, New Zealand) before further processing.

In total, 72 Achilles tendons from 36 rabbits were evenly allocated to 4 groups (see

Figure 1). In the native control group 18 Achilles tendons were dissected from the

rabbits and stored for further experiments. In the 6 and 12 day load deprivation groups

(LD6, LD12), 18 tendons each were cultured in the growth medium (DMEM/F-12, 10%

159

fetal bovine serum and 50 µg/ml Gentamicin) on a petri dish without mechanical

loading at 37℃ in a humid environment with 5% CO2 for either 6 or 12 days. In the

stimulation group an additional 18 rabbit Achilles tendon were cultured on petri dish

without mechanical loading for 6 days, followed by 6 days in the bioreactor chambers

undergoing 6% cyclic tensile loading.

Figure 5. 1 Flowchart of experimental plan and rabbit Achilles tendon allocation (LD: loading

deprivation)

In the bioreactor chambers, the Achilles tendons were secured to the tissue hook using

surgical sutures, as shown in Figure 5.2A, and then placed into the bioreactor chamber,

which was in turn placed in a standard incubator at 37℃ in a humid environment with

160

5% CO2 (Figure 5.2B). During the dynamic culture process, the tendons were subjected

to 6% cyclic strain mechanical stimulation at 0.25Hz for 8 hours followed by 16 hours

rest (7200 cycles/day) for total 6 days (see Figure 5.2C).

Figure 5. 2 (A) Rabbit Achilles tendon was fixed in a bioreactor culture chamber; (B) Bioreactor setup in the incubator; (C) mechanical stimulation regime applied in the stimulation group; (D) Biomechanical testing with a custom cryogrips. The arrow indicates the rupture of

Achilles tendon happened at the middle portion.

In each group, 6 samples were fixed in 4% paraformaldehyde overnight for

histological analysis, 6 samples were directly subjected to RNA extraction, and the rest

were wrapped by PBS soaked gauze and stored at -20℃ for further mechanical testing.

161

5.3.2 Histological preparation and assessment

Refer to 2.5

5.3.3 TUNEL assay

Refer to 2.6

5.3.4 Biomechanics testing

To examine the effect of mechanical stimulation on the biomechanical properties of

tendon, we performed mechanical testing using a modified dual frozen technique (Chen,

Yu et al. 2011). Nine tendons from each group were used for the testing. The samples

were thawed at room temperature on the day of testing. Both bone-tendon and

tendon-muscle ends were placed in a pair of custom-designed cry grips. To prevent

tissue sliding, both ends were snap-frozen by liquid nitrogen. A thermometer was used

to monitor the temperature of the tissue between two cryogrips, to avoid the testing

substrate freezing. The grips were then fixed into the Instron mechanical testing system

(model 5566, Instron, China), and the middle portion of the Achilles tendon were

subjected to tensile loading to failure (Chen, Yu et al. 2011). The samples were

preloaded at 2N, then gradually pulled at a speed of 2mm/second until it reach the

ultimate force and dropped to failure force. Failure force was preprogrammed as 70% of

ultimate force. During the experiment, the samples were closely monitored. To confirm

that each tested tendon was broken at the middle portion but not the point near the grip

162

(Figure 2A). A force versus displacement curve was recorded, and the ultimate load was

recorded. Mean stiffness was calculated based on the linear region of the loading curve.

5.3.5 Real-time polymerase chain reaction

Refer to 2.3 and 2.4

5.3.6 Immunohistochemistry

Refer to 2.6

5.3.7 Statistical analyses

Refer to 2.8

5.4 Results

5.4.1 Cyclic mechanical stimulation improved the histological structure and cell

viability of cultured tendon

The Achilles tendons cultured under loading deprivation exhibit minor wavy and loose

collagen structure (Figure 5.3B), compared to native tendon (Figure 5.3A), more gaps

between the collagen fiber were observed, and the tenocytes started to lose their

163

elongated morphology (Figure 5.3F). When the culture period extended to 12 days,

disorientated fibers with progressive extracellular matrix (ECM) disruption and rounded

cell nuclei were observed (Figure 5.3C, G). These morphological changes are consistent

with the pathology of tendinopathy. Interestingly, when 6% strain cyclic mechanical

loading was applied on the Achilles tendon after 6 days loading deprivation culture, the

tendon morphology seems to be similar to native tendon. Collagen fibers were found to

be straight and well organized (Figure 5.3D), with no gap was observed between

collagen fibers, while long spindle-shape cell morphology was successfully restored

(Figure 5.3H).

Figure 5. 3 Comparison of morphological features of Achilles tendons. (A) Native Achilles tendon tissue; (B) 6 days loading deprivation culture; (C) 12 days loading deprivation culture; (D) 6 days loading deprivation culture followed by 6% strain dynamic culture; (E) Elongated spindle cell morphology in native tendon tissue; (F) Tenocytes started to lose their original morphology and turn

round after 6 days loading-deprived culture; (G) Round tenocytes; (H) 6% strain stimulation successfully restore the long spindle morphology of tenocytes. Scale bar is 20m.

Using the quantitative scoring system (Figure 5.4), the 6 days loading deprivation group

164

expressed a slight abnormality in fiber arrangement, i.e. fiber structure, cell roundness,

and a moderate increase of cell density. In addition, the overall score further increased

as the loading deprivation culture period was extended to 12 days, whereby every aspect

increased significantly except for cell density. For the stimulation group, the tendon

successfully regained normality, in both the overall score and the individual scores with

no significant differences observed when compared to native tendon.

Figure 5. 4Histological score of three groups. (A) Overall histology score, and the 6% rescue group is significantly better than the other groups; (B) fiber arrangement; (C) fiber structure; (D) cell roundness; (E) cell density. Results are expressed as the mean±SEM. One way ANOVA significance values were *p

< 0.05, **p < 0.01, and ***p < 0.001

In regard to cell apoptosis, the 6 day loading deprivation group exhibited numerous

165

positive TUNEL stained tenocytes, as can be observed under the microscope image

(Figure 5.5B), while for the 12 day loading deprivation group, large numbers of

apoptotic cells were observed (Figure 5.5C). However, there were only a few apoptotic

cells detected in the stimulation group (Figure 5.5D), similar to that seen in the native

tendon (Figure 5.5A). Quantification of these observations showed that cell apoptosis

rate was around 30.6% for the 6 days loading deprivation group, which further

increased to 91.3% in the 12 days loading deprivation group. However, the stimulation

group’s tissue had a 17.8% cell apoptosis rate that was significantly lower than either

load deprivation groups (Figure 5.5E), and only slightly higher than the native group.

166

Figure 5. 5TUNEL assay (A) Apoptotic cell was rarely detected in native Achilles tendon tissue (B) In 6 days loading deprivation group, around 30% of tenocytes were under apoptosis, pointed by arrows; (C) In 12 days loading deprivation group, almost all the tenocytes were undergo apoptosis;

(D) apoptotic tenocytes significantly decreased compared to the other experimental groups; (E) Quantitative cell apoptosis rate. Results are expressed as the mean±SEM. One way ANOVA

significance values were *p < 0.05, **p < 0.01, and ***p < 0.001 Scale bar is 20mm.

167

5.4.2 Cyclic mechanical stimulation improved mechanical properties of cultured

tendon

As the force transferring tissue, mechanical properties are the fundamental characteristic

of tendon, although the pathological changes in the tendon microstructure were

observed, the impact on the overall mechanical properties is still unclear. Therefore,

mechanical testing was performed to examine the strength of the Achilles tendon after

the ex vivo culture. In the mechanical tests, all samples failed at the mid-tendon region,

which was the region of interest in this study (Figure 5.2D). As shown in Figure 5.6A,

the average ultimate tensile force after 6 days of loading deprivation was 613.6±49.7 N,

which was significantly higher than 12 days loading deprivation group (449.8±55.9 N)

but lower than stimulation group (773.3±89.6 N). However, all the culture groups had a

significantly lower ultimate tensile force compared to the native Achilles tendon (896.1

±96.1 N).

The mean stiffness (Figure 5.6B) of the stimulation group was 165.6±39.1 N/mm,

which was significantly higher than 6 days and 12 days loading deprivation groups

(99.2 ± 22.4 N/mm, 52.5 ± 12.8 N/mm), while there was no statistical difference

compared to the native controls (193.1±26.8 N/mm).

168

Figure 5. 6 (A) The maximum load of Achilles tendon tissue. The maximum loads decreased with extended cultured time in static cultured group, however, after 6 days static culture, 6% stimulation group is able to restored part of mechanical properties. Compared to native control tissue, the maximum loads in

cultured group are significantly lower. (B) The mean stiffness of Achilles tendon tissue. The mean stiffness of static cultured group dropped dramatically with extended culture period. However, in 6% rescue group, no significant difference was observed compared to native control tendon. Results are

expressed as the mean±SD. One way ANOVA significance values were *p < 0.05, **p < 0.01, and ***p < 0.001

The force versus displacement curves of native Achilles tendons, 6 days loading

deprivation group and stimulation group are similar (Figure 5.7). In these the force

increased dramatically for the first 4mm displacement, and then moderately increased,

which was followed by a sudden drop in the tensile force as the tendons began to fail.

However, in the 12 days loading deprivation group, the force climbed moderately for

the first 8mm and slowly decreased to 70% of ultimate force, at which point the tests

were terminated.

169

Figure 5. 7 Force versus displacement curves for the native Achilles tendon (Control), rescue group (0% 6d+6% 6d), 6 days static cultured and 12 days static cultured Achilles tendon. 6% rescued cultured tendon displays similar mechanical character with native Achilles tendon. After 12 days static culture,

Achilles tendon lost its original mechanical character became less stiff and soft.

5.4.3 Rescue effect of cyclic mechanical stimulation on ECM remodeling gene

expression

Although the histological and mechanical changes have been identified, the underlying

molecular mechanism still remains unclear. The homeostasis of tendon required the

balance of the ECM generation and degradation. To estimate the generative process, the

gene expression of COL1A1, COL3A1, TGFbwere tested, whiles MMP1, MMP3,

TIMP1, and TIMP2 were considered to be the predominant cytokine in tendon

remodeling. All the gene expression level was examined using RT-PCR. As shown in

Figure. 5.8, although there was a slight decrease on COL1A1 expression of the loading

170

deprivation groups compared to native control, no significant difference was detected;

however, COL1A1 expression increased dramatically in the stimulation group. A similar

trend was observed in TGFband MMP3 expression. For the first 6 days no loading

culture, MMP 1 was highly expressed, yet it dropped with extended culture period. In

the stimulation group, MMP 1 was less expressed compared to 6 days loading

deprivation group, however, no statistical difference was detected compared to the 12

day loading deprivation group. On the other hand, TIMPs expressions of culture tendon

were significant higher than the native control except for the 6 day loading deprivation

group on TIMP2. Extensive increase on TIMPs expressions were detected on

stimulation group, especially TIMP1. However, no significant difference between

stimulation group and 12 days loading deprivation groups on TIMP2.

Figure 5. 8 Comparison of gene expression fold change normalized by native tissue gene expression amount in the 6 day and 12 day loading deprivation group, the stimulation group and the native control group, the gene expression level is normalized against 36B4. Result are expressed as

the mean±SEM. One way ANOVA significance values were *p < 0.05, **p < 0.01, and ***p < 0.001

171

5.4.4 Cyclic mechanical stimulation decreased Type III collagen expression of

cultured tendon

Increased collagen type III content is an index of tendon degeneration and tendinopathy.

Tendon cultured under loading deprivation condition for 6 days started to express

moderate collagen type III turnover and several positive stained tenocytes were detected

(as pointed by arrow in Figure 5.9B). Furthermore, almost all the tenocytes in the 12

days loading deprivation group were expressing type III collagen, and substantial type

III collagen was observed histologically (Figure 5.9C). Interestingly, in the stimulation

group, collagen type III expression was only minimally elevated, but at a level (Figure

5.9D) similar to native tendon (Figure 5.9A).

Figure 5. 9 Type III collagen immunohistochemistry, positive staining is indicated by arrows (A) Type III collagen rarely detected in native Achilles tendon tissue (B) In 6 days loading deprivation group, type III collagen was upregulated, positive staining were pointed by arrow; (C) In 12 days

loading deprivation group with massive increased type III collagen expression ; (D) No obvious type III collagen expression detected in 6% rescue group. Scale bar is 20m.

172

5.5 Discussion

It is well known that mechanical stimulation plays a fundamental role in maintaining

tendon homeostasis, and that the absence of mechanical stimulation can cause

morphological changes such as rounded cell nuclei and disorientated collagen fibers

associated with tissue degeneration (Hannafin, Arnoczky et al. 1995; Wang, Lin et al.

2013). “Tendon exercise” such as eccentric overload training or concentric training

seems to be effective in patient rehabilitation, however, the mechanism still remains

unclear(Kingma, de Knikker et al. 2007). In Chapter 3, a bioreactor system was

developed that could be used to mimic the in vivo tensile mechanical loading on

Achilles tendon and in Chapter 4 it was verified that cyclic mechanical stimulation

indeed was essential for the maintenance of the tissue integrity of tendon. We also

suggested that a narrow range of mechanical stimulation was required to create a net

anabolic effect in Achilles tendon, and identified that 6% strain cyclic mechanical

stimulation was optimal to maintain Achilles tendon homeostasis in the bioreactor

system(Wang, Lin et al. 2013). However, the therapeutic effect of mechanical

stimulation on early-stage of pathological change in Achilles tendon was yet to be

observed.

In this study, this dynamic bioreactor culture system was combined with the unloaded

culture condition to examine the effect of cyclic mechanical stimulation on deteriorated

tissue caused by loading deprivation. Loading deprivation was achieved by culturing

the rabbit Achilles tendons freely on the petri dish with adequate growth medium. After

173

6 days of loading deprivation, early-stage degenerative changes were observed,

including wavy collagen fiber, increased type III collagen expression, and slightly

rounded nuclei. After 12 days of culture without loading, further severe pathological

changes such as disorientated collagen fiber pattern, ruptured fibers, rounded nuclei and

substantial increased type III collagen expression developed evenly throughout the

tendon. The deterioration of the tendon appeared to be no different at the center and the

edge of the tissue, suggesting that inadequate nutrition penetration was not the sole

cause of the damage. In the stimulation group, the tendons were transferred into the

bioreactor system after 6 days loading deprivation, and 6% strain cyclic mechanical

strain stimulation was applied for a further 6 days. The stimulation tendons successfully

regained their original morphology, exhibiting as straight, well-organized collagen

fibers and cells with elongated nuclei, whilst the type III collagen expression had

reduced to the normal level. The result suggested that the previously identified optimal

cyclic tensile loading (6%) was able to repair the early pathological change of tendon

caused by load deprivation.

Programmed cell death (apoptosis) has been identified as a common feature of tendon

degeneration and injury (Yuan, Wang et al. 2003; Hosaka, Teraoka et al. 2005; Lian,

Scott et al. 2007). I previously showed that mechanical loading was essential for

tenocyte viability, where increased cell apoptosis was observed in the tendon tissues that

has no mechanical loading (Wang, Lin et al. 2013). In this study, the cell apoptosis rate

in the 6 and 12 days loading deprivation groups were 30.6±3.9% and 91.3±3.5%

respectively, which might explain why in the 12 day loading deprivation group, no

174

increased cell density was detected. When the tendon developed early degeneration, cell

proliferation increased, presumably to activate repair. However, in the absence of

mechanical stimulation, the observed high cell apoptosis rate inhibited this repair

cascade then reducing the cell population to a “normal” level. The stimulation group,

on the other hand, expressed an apoptosis rate of 17.8±3.9%, which was closed to that

of healthy tendon tissue, suggesting that 6% strain cyclic mechanical stimulation

eliminated cell apoptosis in early-stage degradation and may thus contribute to the

restoration of the homeostasis of tendon.

The microstructural organization of the tendon’s extracellular matrix, particularly

collagen type I, determines the mechanical properties of tendon tissue. Although the

disturbed collagen structure and increased collagen type III turnover were observed in

loading deprivation culture groups, this was rescued by 6% strain cyclic loading

stimulation. Critically mechanical testing confirmed that the mechanical stimulation

restored normal mechanical properties of the Achilles tendon.

The biomechanical characteristics of tendon have not been well documented in

literature as failure often occurs at sites other than the middle portion of tendon, such as

muscle-tendon junction, tendon-bone junction or surgical repair site (Thermann,

Frerichs et al. 2001; Matsumoto, Trudel et al. 2003; Gebauer, Beil et al. 2007; Chen, Yu

et al. 2011). To ensure the efficacy of the testing, a pair of modified cryogrips and the

freezing technique using liquid nitrogen was adopted in the study to avoid sliding of the

175

tissue during testing. During the mechanical loading, the tissue was closely monitored

to ensure it broke at the middle portion rather than the ends near the grips. The peak

load of native control tissue tested for the whole Achilles tendon in this study was 896.1

±96.1 N, which is comparable to the findings reported in the study by Trudel et

al.(Trudel, Koike et al. 2007). The results of biomechanical testing showed significantly

reduced mechanical strength and stiffness of Achilles tendon cultured for 6 days and 12

days load deprivation. The mechanical characteristic could be better described from the

force versus displacement curves. Due to ruptured and disorientated collagen fibers, the

force increased more slowly in 12 days loading deprivation group compared to the other

groups, for a given strain. However, in the stimulation group, the rabbit Achilles tendons

regained their mechanical properties to levels close to that of normal tissue. This was

evidenced by there being no significant difference in the mean stiffness of the

stimulation group compared to that of the native control group, and the

force-displacement curve in stimulation group appearing similar to that of the native

control, apart from a lower ultimate tensile strength. Thus, the testing demonstrated a

positive effect of mechanical stimulation in preventing further tendon degeneration, and

indeed in promoting tendon healing.

The homeostasis of tendon requires constant remodeling of the collagen through the

generation of new collagen in response to the degradation of collagen by proteases and

mechanical loading. Matrix metalloproteinase (MMPs) are zinc-dependent

endopeptidases that are able to degrade all the components of the extracellular matrix.

MMP-1 is primarily responsible for cleavage the collagen type I, the dominant collagen

176

of tendon (Knauper, Lopez-Otin et al. 1996; Krane, Byrne et al. 1996). MMP-3 is

another subset of the MMP family with a wide range of substrate specificity. Moreover,

it has the ability to activate other MMPs such as MMP-1, MMP-7 and MMP-9(Ye,

Eriksson et al. 1996). TIMPs, on the other hand, are the inhibitors of MMPs. The ratio

of MMPs to TIMPs is the key to understanding the rate of protease degradation, which

then impacts on tendon homeostasis. Over expression of MMPs or under expression of

TIMPs will cause the pathogenesis of tendon degradation (Riley, Curry et al. 2002; Lo,

Marchuk et al. 2004; Jones, Corps et al. 2006). In our study, in the first 6 days of

loading deprivation culture, MMP1 and MMP3 were significantly up-regulated, while

TIMP1 and 2 were down-regulated. The shift in balance between MMPs and TIMPs

may explain the early-stage pathological changes observed in the tendon. However, in

the stimulation group, the up-regulation of TIMP1 and down-regulation of MMP1 may

have suppressed the protease degeneration of collagen fibers. Upon mechanical

stimulation, expression of TGF- and COL1A2 promoted the formation of collagen

fiber. The RT-PCR results are consistent with our observation that 6% cyclic tensile

strain mechanical stimulation was optimal for creating a net anabolic effect that could

reverse the early-stage degenerated changes.

Immobilization is a common postoperative treatment for bone fracture and tendon injury.

For example, fracture of the lower end of the tibia or rupture of the Achilles tendon

commonly require immobilization for up to 6 weeks (Cetti, Henriksen et al. 1994)

(Ruedi and Allgower 1979), resulting in the deprivation of physical dynamic loading of

the tendon. Early-stage structural degradation of Achilles tendon has been reported after

177

long-term immobilization in both humans and rats (Larsen, Forwood et al. 1987;

Nakagawa, Totsuka et al. 1989; Kannus, Jozsa et al. 1997; Trudel, Koike et al. 2007).

However, clinical rehabilitation guideline for Achilles tendon weakness after a period of

immobilization has not been well established. In this study, we have demonstrated that

loading deprivation in a bioreactor tendon culture led to tendon structural deterioration

compromised mechanical properties and elevated expression levels of catabolic markers.

By applying 6% strain cyclic mechanical stimulation, the histomorphology and

biomechanical properties of rabbit Achilles tendon were successfully restored. However,

all characteristics of the preferred cyclic strain profiles need to be defined, i.e.

magnitude, frequency and dosage, and these may vary between species. Interestingly, a

recent review suggests that about 5% strain may be the value that affects human

Achilles tendons in vivo(Obst, Newsham-West et al.). Even though all strain profile

characteristics are to be elucidated, the findings of this current study suggest that

precisely calibrated mechanical stimulation is able to reverse early-stage tendon

degradation. It has been reported that compared to long episodes of early loading, short

periods of mechanical stimulation during the inflammatory phase could improve tendon

healing with histological signs of repeated damage in the form of increased bleeding

(Enwemeka, Spielholz et al. 1988; Eliasson, Andersson et al. 2011). Taken together,

accurate and controllable mechanical stimulation on injury tendon may be considered as

a potential therapeutic approach for patient rehabilitation. A specially designed orthotic

device could be developed to apply an optimal titration of strain on tendon tissue and

facilitate tendon repair.

178

5.6 Conclusion

In summary, our findings suggest that loading deprivation induced degenerated

pathogenesis in rabbit Achilles tendon could be rescued by 6% cyclic tensile loading

stimulation. The findings suggest that loading deprivation, such as during long-term

immobilization, should be avoided when possible to reduce the complication of tendon

weakness in patients. In addition, the benefit of dynamic loading may be an inspiration

to develop a rehabilitation device for tendon weakness after long-term immobilization.

179

Chapter 6

General Discussion and Future

Directions

180

CHAPTER 6 GENERAL DISCUSSION AND FUTURE

DIRECTIONS

6.1 General discussion

Tendon is a soft connective tissue between muscle and bone and functions to transfer

forces. Tendons are also responsible for storing elastic energy. Tendons are subjected to

high mechanical loading during daily activities and must undergo continuous repair to

maintain its functional properties. Owing to its specific mechanical function, it is

unsurprising that its biomechanical environment is essential for maintaining tendon

homeostasis.

In 1995, Hannafin et al. demonstrated that even without mechanical loading, tendons

undergo tendinopathy-like morphological changes characterized by disorientated

collagen fibers and rounded cell nuclei(Hannafin, Arnoczky et al. 1995). Manipulation

of loading on the tendon is able to regulate tendon metabolism and activate self-repair

mechanisms. Eccentric overload is one of the known effective interventions for patients

with chronic Achilles tendinopathy. By maximizing the activation of soleus or

gastrocnemius muscles, mechanical loading is able to be applied on tendon(Kingma, de

Knikker et al. 2007). Traditionally, immobilization is a standard recommendation

post-surgical tendon repair. However, studies have shown that short periods of

mechanical stimulation during the inflammatory phase could improve tendon healing

(Enwemeka, Spielholz et al. 1988; Eliasson, Andersson et al. 2011). This is also

consistent with the results present in this thesis (Chapter 5). Various studies on tendon

181

engineering suggest that mechanical stimulation is able to increase cell proliferation,

collagen deposition and improve the mechanical properties of neo-tendon tissue(Altman,

Lu et al. 2002; Juncosa-Melvin, Shearn et al. 2006; Webb, Hitchcock et al. 2006;

Androjna, Spragg et al. 2007; Nirmalanandhan, Rao et al. 2008; Nirmalanandhan,

Shearn et al. 2008; Butler, Hunter et al. 2009; Nguyen, Liang et al. 2009; Chen, Yin et

al. 2010; Doroski, Levenston et al. 2010; Saber, Zhang et al. 2010; Issa, Engebretson et

al. 2011; Woon, Kraus et al. 2011). Previous evidence suggests that mechanical

stimulation is able to confer beneficial effects on tendon regeneration, although the

reported range of improvements is varied. Silbernagel et al. showed that in 12 weeks of

eccentric training, the overall improvement on pain relief was 29%(Silbernagel, Thomee

et al. 2001), however, in a study by Alfredson et al., 94% improvement was reported

(Alfredson, Pietila et al. 1998). Owing to limited understanding of the biomechanical

environment in tendon tissue and the tissues’ repair and remodeling response, the

explanation of the present results within a physiological framework was difficult. It is

well known that sudden and excessive loading on the tendon can cause tear or rupture

and loading deprivation can cause tendon degradation. However, the physiology of

loading and underloading is rarely studied. Therefore, the investigation of various

loading conditions on the tendon is essential for understanding tendon biology and the

development of treatments for tendon repair.

6.1.1 The custom-made bioreactor (Chapter 3)

In the past decade, various methodologies have been employed to study the effect of

mechanical loading on tendon such as in vitro monolayer cellular models and in vivo

182

animal models. In the cellular studies, the Flexcell® system has been widely used

(Skutek, van Griensven et al. 2001; Wang, Jia et al. 2003; Yang, Im et al. 2005).

Although precise control and monitoring systems are allowed at the cellular level, the

application is on a monolayer of cells, not an organized tissue. Cells isolated from their

extracellular niche and cultivated in a monolayer may not necessarily replicate their

behavior in vivo. In vivo models, on the other hand, allow the systemic study of the

biomechanical environment of the tendon. However, mechanical stimulation on the

tendon varies between individuals, making precise quantification of load difficult.

Therefore, in order to study the effect of various mechanical stimulations on tendon, a

bioreactor that can provide a sterilized and programmable biomechanical environment

for tendon culture is necessary.

In the first study, we adopted a relatively simple “incubator-based” design, in which the

culture conditions such as temperature, humidity and CO2 are controlled by the

incubator. An actuation system and a culture chamber are the key components of the

bioreactor system, as they provide the mechanical output and the sterilized culture

environment, respectively. A step motor-ball screw transmission system (SMBS) was

adopted as the actuator in our bioreactor. Compared to the other actuator systems such

as the pneumatic actuators(Juncosa-Melvin, Shearn et al. 2006; Nirmalanandhan, Rao et

al. 2008; Nirmalanandhan, Shearn et al. 2008) and linear motors(Nguyen, Liang et al.

2009; Doroski, Levenston et al. 2010), SMBS has the advantages of high resolution,

highloading capacity and good resistance to humid and warm environment(Liu, Hsu et

al. 2001; Altman, Lu et al. 2002; Webb, Hitchcock et al. 2006; Androjna, Spragg et al.

183

2007). AISI 316 stainless steel, which has good machinability and resistance to

chemicals and high temperature, was used for the construction of the culture chamber.

The labyrinth channel integrated in the chamber ensured sufficient air exchange without

inducing contamination, and the glass covers allowed for visual monitoring. Several

commercial bioreactor systems have been developed such as The Bose® ElectroForce

BioDynamic® system and the LigaGen system. Compared to these products, our system

has an advantage in that it provides two methods for tissue fixation, tissue clamp and

tissue hook. Tissue clamps have been found to be more effective at stabilizing

membrane-like tissue, while tissue hooks are designed for fixing thick and

irregular-shaped tissue.

In the present studies, full-length rabbit Achilles tendons, including the bone-tendon and

tendon-muscle junction, were used as an Achilles tendon tissue model and cultured in

the bioreactor system. Tissue hooks were found to be able to provide excellent fixation

without slippage, which was essential for this study. Slippage can cause inconsistent

mechanical loading and can comprise the results.

6.1.2 Mechanical stimulation in bioreactor system (Chapter 4 and 5)

The present studies utilized the novel approach of applying mechanical stimulation on

fresh-harvest rabbit Achilles tendon in a bioreactor system. The application of the

bioreactor in evaluating the effect of mechanical stimulation on tendon has several

advantages:

184

Unlike the monolayer cellular study, the tenocytes were in their native niche

embedded in the collagen matrix. During mechanical stimulation, tenocytes were

not only subjected to tensile strain, but also compression from adjacent collagen

fibers (Caliari, Weisgerber et al. 2012). Therefore, mechanical stimulation applied

by the bioreactor system can be better replicated in in vivo loading environments

compared to a monolayer system.

Unlike animal studies such as treadmill training, the bioreactor system used in this

study is able to provide accurate, programmable and reproducible mechanical

stimulation, allowing the systematic study of different intensity loading on the

tendon.

The use of a bioreactor system allows the chemical environment in the culture to be

manipulated. Therefore, it offers the potential for systematically studying the effects

of nutrients, growth factors and cytokines under different loading regimes.

6.1.3 Effect of various mechanical stimulation on tendon homeostasis (Chapter 4)

Full-length rabbit Achilles tendons were dissected and cultured under different

mechanical environments: 0% 3% 6% and 9% strain at 0.25Hz 8h/day for 6 days.

Histology, immunohistochemistry, TUNEL assay and RT-PCR were carried out to

evaluate tissue integrity.

Histologically, tendons subjected to 6% strain mechanical loading exhibited similar

morphology as native tendon. However, loading deprivation group, 3% strain and 9%

185

strain groups showed pathological changes.

Loading deprivation

Tendon exhibited minor pathological changes including loose collagen fiber and a

moderately increased cell apoptosis rate after 6 days of loading deprivation culture

(Figure 4.2 B). The morphology further degenerated when the culture period was

extended to 2 weeks. Typical tendinopathy morphologies were observed such as

massive cellular apoptosis, disorientated collagen fiber and rounded tenocyte nuclei

(Figure 4.2C, 5.4C). The severe pathological changes are initiated by high cell

death. The remaining cells then fail to maintain the tissue integrity.

3% strain

In the 3% strain group, wavy collagen fiber, minor rounded cell nuclei, loose ECM

structure and increased collagen type III synthesis were observed (Figure 4.3B,

5.5C). The elevated MMPs expression and downregulation of COL1A2 are

consistent with these pathological changes. In 2007, Arnoczky et al. hypothesized

that pathological development of tendinopathy maybe the result of under

stimulation rather than overloading(Arnoczky, Lavagnino et al. 2007). Our study

provided further evidence in support of this theory. It is suggested in this thesis that

increase protease activity acting on underloaded collagen and reduced anabolic

activity of tenocytes due to reduced mechanical stimulus, despite a reduced

mechanical damage, can explain the observed pathological changes in this 3%

strain group.

186

6% strain

In the 6% loading group, tendons were able to maintain their normal morphology,

type III collagen content and apoptosis rate. RT-PCT results showed an increased

expression of COL1A2. The rate of micro-damage caused by mechanical loading is

matched by the rate of repair such that the structural integrity was able to be

maintained.

9% strain

In this group, an acute injury model was successfully established. Tendon subjected

to excessive loading caused rupture of collagen fiber (i.e. mechanical damage) and

increased cell apoptosis close to the injury site, which are the typical morphology of

acute injury on tendon. Type III collagen was highly expressed in RT-PCR results

and was confirmed by immunohistochemistry.

Although pathological changes were detected in three of four loading conditions, the

underlying causes are different. From the different loading groups and the pathological

changes and protein expressions, we proposed a hypothetical model for tendon damage

and repair, as shown in Figure 4.7. An increased intensity of mechanical stimulation

increases the mechanical damage. Protease (chemical) damage was active in the low

strain groups. Only at intermediate strains of 6% was the repair able to match the rate of

mechanical and proteolytic damage, and hence the ability to maintain structural integrity.

187

Therefore, the conclusion drawn from this study is that a narrow range of mechanical

stimulation in this bioreactor system was able to maintain tendon homeostasis; with

mechanical underloading and overloading both detrimental to tendon.

6.1.4 Rescue of degenerated tendon with mechanical stimulation.

In the previous study, we demonstrated that 6% strain mechanical stimulation was able

to maintain the tendon homeostasis in the bioreactor system. Following this, the next

hypothesis was whether or not 6% strain can also rescue tendon following load

deprivation. Therefore, in this study, rabbit Achilles tendon was first cultured without

loading for 6 days, and then followed by either an additional 6 days without load or 6

days at 6% strain at 0.25Hz 8hr/day for. The tendons then were analyzed by

biomechanical testing and histological assessment.

Compared to native rabbit Achilles tendons, tendons under static culture for 6 days and

12 days showed 30% and 50% reduction in their mechanical properties, respectively.

However, the rescue group showed significant improvement compared to these static

culture groups. Indeed tendons in the rescue group exhibited similar morphology,

apoptosis rate and collagen type III content to native tendons. RT-PCR results showed,

application of 6% strain mechanical stimulation significantly reduced MMPs expression

and increases COL1A2 and TIMPs expression, which was consistent with a suppression

of tissue catabolic processes and an increase in those associated with tissue repair.

188

Immobilization is a common postoperative treatment of bone fracture and tendon

injuries(Ruedi and Allgower 1979; Cetti, Henriksen et al. 1994). In various studies,

early-stage degradation in Achilles tendon was reported after long-term immobilization

in both humans and rats(Larsen, Forwood et al. 1987; Nakagawa, Totsuka et al. 1989;

Kannus, Jozsa et al. 1997; Trudel, Koike et al. 2007). However, there is no clinical

rehabilitation guideline for Achilles tendon weakness after a period of immobilization.

The results of this study suggests that proper mechanical loading may activate a

self-repair mechanism to repair early stage degenerative pathological changes in tendon.

6.2 Limitation of this work

It is important to mention several limitations of the work described in this PhD thesis.

6.2.1 Chapter 3

In chapter 3, the bioreactor designed is only able to provide some basic function; there

are still several limitations in this design, with the benefit of hindsight could be

improved.

1. Failure warning: although the bioreactor system can provide preprogrammed

automatic mechanical stimulation on the tissue sample, some accidents like power

failure and maloperation might still happen. A complete computer monitoring

system would be ideal, but even a basic warning light should be installed and

preprogrammed in the system to report the error.

189

2. Program interface: The operational system was programed using the basic PLC

language, which might not be familiar to the other users. A more user-friendly

‘common’ programming language should be designed for the convenience of the

other users.

3. Bioreactor connection: in the present design, the bioreactor is connected to the

control box through a cable, which required specific incubator model with a cable

hole back panel. However, a cable hole is not a standard configuration in all

incubators; therefore, this bioreactor can only be used in some special incubator.

6.2.2 Chapter4

1. Mechanical stimulation accuracy: Achilles tendons from different individuals

show variations in length. Although the tissue hook of the bioreactor can provide

an effective way to stabilize the tendon and the stimulation distant are fixed at

23mm, some error might still be induced.

2. Medium infiltration: During the in vitro culture, nutrient delivery is depended on

medium infiltration. In loading deprivation group, medium infiltration might not

sufficient for the whole tendon. Although the histological section exhibited

consistent morphology in the center and the edge of the tendon, the nutrient

transport still needs to be considered when interpreting the results.

3. Statistical analyze: In the histological scoring, scores should have been analyzed

using a nonparametric statistical procedure as the data is ordinal, not continuous.

190

6.2.3 Chapter 5

Biomechanical testing: Achilles tendons from different individuals show

variations in initial mechanical properties. Although in this experiment, the

Achilles tendon from same individual we distributed in different groups to

minimize error, it is still needed to be mentioned; especially as the sample size was

relatively small.

6.3 Future directions

6.3.1 Gene profile of tenocyte subjected to 2D or 3D mechanical stimulation

As demonstrated by the results in this thesis, tendons are able to sense and respond to

mechanical loading. Although they are mainly subjected to tensile strain on the

macroscale (i.e. the whole tissue scale), the micromechanical environment is much more

complicated. Tendon is a dense connective tissue that comprised by well-organized

collagen fiber with tenocyte adhered on it. Subjected to overall tensile strain, tenocytes

not only receive tensile but also the shear force and compression from the surrounding

collagen fibers and proteoglycans as well as potential fluid shear stress (Figure 7.1)

(Khan and Scott 2009). However, the in vitro studies that evaluate the effect of

mechanical stimulation on tenocyte or fibroblast mostly focus on tensile strain only. The

Flexcell® system was commonly used in the studies of in vitro gene expression profiles

of tenocytes under mechanical stimulation (Skutek, van Griensven et al. 2001; Wang,

191

Jia et al. 2003; Yang, Im et al. 2005). These systems can precisely control the

mechanical regimen and monitor the cell behaviors, but only can be used for 2D

monolayer cell culture, not the organized tissue. Cells isolated from their extracellular

niche and cultured in a monolayer may not necessarily replicate their original behavior

in vivo. The recent study of Ni et al. successfully generated a scaffold-free engineered

tendon from tendon-derived stem cell by stimulation with connective tissue growth

factor (Ni, Rui et al. 2013). This model allows the better study on the effect of

mechanical stimulation on tenocyte. As the structure of scaffold-free tendon is more

similar to native tendon, when compared to monolayer cell cultures, the in vivo

mechanical stimulation can be better mimicked in our bioreactor system using this

scaffold-free tendon model. The gene profile of tenocyte subjected to 2D and 3D

mechanical stimulation can lead to the better understanding how tenocyte respond to in

vivo mechanical stimulation.

Figure 7. 1 Tenocyte undergoing (A,B) shear and (C) compression during tendon subjected to tensile strain. Figures taken from Khan et al.(Khan and Scott 2009)

192

6.3.2 The effect of mechanical stimulation on tendon nutrient infiltration

Nutrient supply during in vitro culture of tendon mainly depends on passive infiltration.

However, during mechanical stimulation, advective nutrient transport would be induced.

Although Hannafin et al. reported that mechanical stimulation did not alter small

nutrient uptake, compared to static culture, and that the tendon repair improvements

derived from passive motion is not related to an increase of nutrient transport (Hannafin

and Arnoczky 1994), under different loading regimes, or in different tendon types, this

may not be the case. Moreover, the transport of large molecules, such as growth factors

and matrix components, may be affected by mechanical stimulation, as has been shown

in studies in cartilage (Bonassar, Grodzinsky et al. 2000; Zhang, Gardiner et al. 2007).

Although unlike cartilage, the in vivo nutrient supply is dependent on limited blood

vessels instead of synovial fluid, the nutrient infiltration in vitro is essentially the same

as cartilage, especially for large scale animal engineered tendon generation. Therefore,

rabbit Achilles tendon culture in a bioreactor system can be used as a nutrient transport

model for study. Different sizes of dye that are compatible to target cytokines can be

used in the bioreactor system (Day, Megson et al. 2005).Nutrient delivery could then be

assessed based on the dye infiltration. Furthermore, nutrient delivery is often a key

obstacle in tissue engineering. If the nutrient delivery induced by tensile cyclic strain in

the bioreactor is not sufficient, medium circulation system could be designed based on

the current model of bioreactor to enhance nutrient infiltration.

193

6.3.3 The effect of mechanical stimulation on stem cell tenogenic differentiation

Stem cells have been suggested to have great potential in regenerative medicine

application due to its self-renewal capacity, hypoimmunogenicity and multi-lineage

differentiation ability(Uccelli, Moretta et al. 2006). However, the complexity and

reliability of manipulating stem cell to differentiate into a target cell type limits the

clinical application. In vitro induction of stem cell differentiation is commonly achieved

by chemical modulations. Recent studies suggested that mechanical stimulation acted as

an important regulator in stem cell differentiation induced by cytokines (Rui, Lui et al.

2011), and might be essential for lineage selection and cell differentiation. Khayat et al.

suggested low frequency mechanical stimulation inhibited MSC adipogenic

differentiation (Khayat, Rosenzweig et al. 2012), and a similar result was reported by

Sen et al.(Sen, Xie et al. 2008). Continuous cyclic mechanical tension has been

indicated to inhibit osteogenic differentiation of MSC (Shi, Li et al. 2011), whilst short

term cyclic tensile strain has been reported to stimulate osteogenic differentiation (Rui,

Lui et al. 2011). Furthermore, dynamic compression can induce the chondrogenic

differentiation of MSC (Huang, Farrell et al. 2010; Kupcsik, Stoddart et al. 2010). The

study of Chen et al. reported that static mechanical stimulation was able to stimulate

tenogenic differentiation of human embryonic stem cells (Chen, Song et al. 2009).

Although the mechanical stimulation has been suggested as a key regulator of stem cell

differentiates into musculoskeletal cell type, tenogenic differentiation induced by

mechanical loading is relatively rarely recorded. As a mechano-transduction tissue, it is

reasonable to speculate that dynamic tensile strain is able to simulate the tenogenic

differentiation of stem cells, although the magnitude, frequency and duration remain

unclear. The present study showed that 6% strain, 0.25Hz for 8h/day was able to

194

increase collagen type I expression of dissected rabbit Achilles tendon, which was one

of the tenogenic market of tenocytes. Therefore, the mechanical stimulation regime used

in this study should be tested for tenogenic differentiation of stem cells.

6.3.4 Therapeutic guideline

One commonly used conservative treatment for chronic Achilles tendinopathy is

eccentric overload. The basic principle is to maximize the activation of the calf or

soleus muscle using body weight, so generate a tensile strain on Achilles tendon

(Alfredson, Pietila et al. 1998). The treatment is instructed to be continued unless the

patient experience disabling pain, however, if the patient can perform the exercise

without experiencing any pain, extra weight is required(Alfredson, Pietila et al. 1998).

The treatment is normally performed 2 times daily, 7 days/week for 12 weeks, and for

each treatment including 3 sets of loading, 15 repetitions each set(Alfredson, Pietila et

al. 1998). Several studies reported positive effect on symptoms relief(Stanish,

Rubinovich et al. 1986; Alfredson, Pietila et al. 1998; Silbernagel, Thomee et al. 2001;

Alfredson and Lorentzon 2003; Fahlstrom, Jonsson et al. 2003; Roos, Engstrom et al.

2004; Shalabi, Kristoffersen-Wilberg et al. 2004) however, the study of Woodley et al.

showed no effect(Woodley, Newsham-West et al. 2007) and Rompe et al. reported an

inferior results(Rompe, Furia et al. 2008). The present study indicated that there was

only a narrow range of tensile strain stimulation that could achieve a net anabolic or

even therapeutic effect, which was 6% strain in the rabbit Achilles tendon model.

Overloading and underloading would trigger degenerative cascade through different

pathways. This study provides a cue to explain the clinical outcome variety from the

195

eccentric overloading training. As the treatment is based on the sense of patients, the

strain applied on Achilles tendon may not be sufficient to achieve the optimal clinical

outcome; moreover, the treatment is not consistence with different patients. Although

rabbit tissue is different from human tissue, similar principles can be applied in human.

Physiological soleus and calf muscle force during gait, running and jumping can be

measured by the inverse dynamics algorithm, and then the physiological strain on

Achilles tendon can be calculated. Base on the physiological strain range of Achilles

tendon, more experiments are required to evaluate the optimal loading regime for the

rehabilitation program. However, even knowing the optimal loading strain on Achilles

tendon, the monitoring of tendon strain is still a big challenge, which requires various

expertise from sport science.

6.3.5 Robot- driven orthosis

Immobilization is a standard postoperative procedure after tendon surgery to eliminate

the loading on the surgical site then prevent the rerupture during recovery period;

however, several studies have reported that the immobilization clinically caused

degenerative changes in tendon tissue (Amiel, Woo et al. 1982; Loitz, Zernicke et al.

1989). Recent study reported that short-episode of mechanical stimulation at the

inflammatory stage could improve healing with histological signs of repeated damage in

the form of increased bleeding (Enwemeka, Spielholz et al. 1988; Eliasson, Andersson

et al. 2011). Therefore, early intervention seems to be beneficial for the tendon healing;

however, self-initiative training might be difficult to control and might damage the

surgical site. Robot-driven orthosis, in the other hand, can control the activity level and

196

protect the surgical site from overloading. Although no robot-driven orthosis is currently

being used for tendon rehabilitation, a clinical trial on patient suffered from multiple

sclerosis and stroke seems to be promising (Husemann, Muller et al. 2007; Lo and

Triche 2008; Schwartz, Sajin et al. 2012; Straudi, Benedetti et al. 2013). Our studies

indicate that proper mechanical stimulation has therapeutic effect; therefore,

robot-driven orthosis could be developed based on this model as a postoperative

rehabilitation. The combination of surgical treatment and conservative treatment at the

early inflammatory phase might provide a better clinical outcome.

6.3.6 Tendon engineering

Tendon’s primary function is force transmission. It operates in a varying load

environment, both on short timescales (e.g. walking, running) and on longer timescales

(e.g. changes in body size with age). In fact, it has been shown that after 4 weeks in a

load-free culture environment tenocytes lost their native elongated morphology, became

increasingly rounded, and the collagen fiber became more crimped (Hannafin, Arnoczky

et al. 1995). Unsurprisingly, given its functional role, a suitable mechanical stimulus is

vital for engineered tendon/ligament formation. Cyclic stretching has been shown to

produce an up to 9 folds increase in the cell number of an engineered tendon compared

with a static culture over a 2 week period (Abousleiman, Reyes et al. 2009).

Furthermore, mechanical stimulation could help guide collagen fiber formation, i.e.

along the direction of loading, which is able to enhance or optimize the mechanical

properties including stiffness, elastic modulus, maximum tensile stress and maximum

force(Juncosa-Melvin, Shearn et al. 2006; Saber, Zhang et al. 2010; Woon, Kraus et al.

197

2011). However, most mechanical stimulation applied on the engineered tendon

cultured is lack of systemic optimization, the magnitude and frequency varied from

studies. Some studies used up to 10% strain for tendon stimulation, which might excess

the tolerance of native tendon (our study indicated 9% strain could cause partial tear of

rabbit Achilles tendon). In the present study, rabbit Achilles tendon was used as a model

to evaluation the effect of different mechanical stimulation on tendon homeostasis.

Knowing the fact that 6% strain was able to create anabolic effect and increased tendon

formation, the further application of this study is to combine the optimal mechanical

stimulation and the engineered tendon culture to create a better tendon construct for

implantation.

6.4 Conclusion

In the present PhD project, a functional bioreactor that can provide programmable

mechanical stimulation and sterilized culture environment was designed and

constructed. The application of this bioreactor allowed the evaluation on the effect

of different mechanical stimulation on rabbit Achilles tendon homeostasis. It was

observed that only a narrow range of mechanical stimulation can maintain the

tendon integrity, mechanical underloading and overloading were both detrimental

to tendon tissue. It was observed that successful restoration of the tendon’s

structural integrity and mechanical properties was available by the application of 6%

tensile cyclic loading from early degenerative load deprivation model. The current

results suggest that mechanical stimulation may provide therapeutic benefits.

198

REFERENCE

Abousleiman, R. I., Y. Reyes, et al. (2009). "Tendon tissue engineering using cell-seeded umbilical veins cultured in a mechanical stimulator." Tissue Eng Part A 15(4): 787-795.

Abrahamsson, S. O. (1991). "Matrix metabolism and healing in the flexor tendon. Experimental studies on rabbit tendon." Scand J Plast Reconstr Surg Hand Surg Suppl 23: 1-51.

Abrahamsson, S. O. and R. Gelberman (1994). "Maintenance of the gliding surface of tendon autografts in dogs." Acta Orthop Scand 65(5): 548-552.

Abrahamsson, S. O., G. Lundborg, et al. (1991). "Long-term explant culture of rabbit flexor tendon: effects of recombinant human insulin-like growth factor-I and serum on matrix metabolism." J Orthop Res 9(4): 503-515.

Alfredson, H. and R. Lorentzon (2000). "Chronic Achilles tendinosis: recommendations for treatment and prevention." Sports Med 29(2): 135-146.

Alfredson, H. and R. Lorentzon (2003). "Intratendinous glutamate levels and eccentric training in chronic Achilles tendinosis: a prospective study using microdialysis technique." Knee Surg Sports Traumatol Arthrosc 11(3): 196-199.

Alfredson, H., T. Pietila, et al. (1998). "Heavy-load eccentric calf muscle training for the treatment of chronic Achilles tendinosis." Am J Sports Med 26(3): 360-366.

Alter, D. and T. Tsao (1994). Dynamic stiffness enhancement of direct linear motor feed drives for machining. Proceedings of the American Control Conference.

Alter, D. and T. Tsao (1996). "Control of linear motors for machine tool feed drives: design and implementation of H1 optimal feedback control." ASME J Dyn Syst, Meas, Control 118: 649-656.

Altman, G. H., H. H. Lu, et al. (2002). "Advanced bioreactor with controlled application of multi-dimensional strain for tissue engineering." J Biomech Eng 124(6): 742-749.

Amiel, D., S. L. Woo, et al. (1982). "The effect of immobilization on collagen turnover in connective tissue: a biochemical-biomechanical correlation." Acta Orthop Scand 53(3): 325-332.

Androjna, C., R. K. Spragg, et al. (2007). "Mechanical conditioning of cell-seeded small intestine submucosa: a potential tissue-engineering strategy for tendon repair." Tissue Eng 13(2): 233-243.

Anitua, E., M. Sanchez, et al. (2006). "New insights into and novel applications for platelet-rich fibrin therapies." Trends Biotechnol 24(5): 227-234.

199

Arancia, G., P. Crateri Trovalusci, et al. (1989). "Ultrastructural changes induced by hyperthermia in Chinese hamster V79 fibroblasts." Int J Hyperthermia 5(3): 341-350.

Archambault, J., M. Tsuzaki, et al. (2002). "Stretch and interleukin-1beta induce matrix metalloproteinases in rabbit tendon cells in vitro." J Orthop Res 20(1): 36-39.

Arnoczky, S. P., M. Lavagnino, et al. (2007). "The mechanobiological aetiopathogenesis of tendinopathy: is it the over-stimulation or the under-stimulation of tendon cells?" Int J Exp Pathol 88(4): 217-226.

Arnoczky, S. P., M. Lavagnino, et al. (2008). "Loss of homeostatic strain alters mechanostat "set point" of tendon cells in vitro." Clin Orthop Relat Res 466(7): 1583-1591.

Arnoczky, S. P., T. Tian, et al. (2004). "Ex vivo static tensile loading inhibits MMP-1 expression in rat tail tendon cells through a cytoskeletally based mechanotransduction mechanism." J Orthop Res 22(2): 328-333.

Arnoczky, S. P., T. Tian, et al. (2002). "Activation of stress-activated protein kinases (SAPK) in tendon cells following cyclic strain: the effects of strain frequency, strain magnitude, and cytosolic calcium." J Orthop Res 20(5): 947-952.

Arnsdorf, E. J., L. M. Jones, et al. (2009). "The periosteum as a cellular source for functional tissue engineering." Tissue Eng Part A 15(9): 2637-2642.

Arora, N. S., T. Ramanayake, et al. (2009). "Platelet-rich plasma: a literature review." Implant Dent 18(4): 303-310.

Asundi, K. R., K. B. King, et al. (2008). "Evaluation of gene expression through qRT-PCR in cyclically loaded tendons: an in vivo model." Eur J Appl Physiol 102(3): 265-270.

Awad, H. A., D. L. Butler, et al. (1999). "Autologous mesenchymal stem cell-mediated repair of tendon." Tissue Eng 5(3): 267-277.

Awad, H. A., D. L. Butler, et al. (2000). "In vitro characterization of mesenchymal stem cell-seeded collagen scaffolds for tendon repair: effects of initial seeding density on contraction kinetics." J Biomed Mater Res 51(2): 233-240.

Baysal, A., S. Bilsel, et al. (2006). "Comparison of the usage of intravenous iloprost and nitroglycerin for pulmonary hypertension during valvular heart surgery." Heart Surg Forum 9(1): E536-542.

Benjamin, M., B. Moriggl, et al. (2004). "The "enthesis organ" concept: why enthesopathies may not present as focal insertional disorders." Arthritis Rheum 50(10): 3306-3313.

Benjamin, M. and J. R. Ralphs (1998). "Fibrocartilage in tendons and ligaments--an

200

adaptation to compressive load." J Anat 193 ( Pt 4): 481-494.

Bennett, N. T. and G. S. Schultz (1993). "Growth factors and wound healing: biochemical properties of growth factors and their receptors." Am J Surg 165(6): 728-737.

Berse, B., J. A. Hunt, et al. (1999). "Hypoxia augments cytokine (transforming growth factor-beta (TGF-beta) and IL-1)-induced vascular endothelial growth factor secretion by human synovial fibroblasts." Clin Exp Immunol 115(1): 176-182.

Beskin, J. L., R. A. Sanders, et al. (1987). "Surgical repair of Achilles tendon ruptures." Am J Sports Med 15(1): 1-8.

Bi, Y., D. Ehirchiou, et al. (2007). "Identification of tendon stem/progenitor cells and the role of the extracellular matrix in their niche." Nat Med 13(10): 1219-1227.

Birch, H. L. (2007). "Tendon matrix composition and turnover in relation to functional requirements." Int J Exp Pathol 88(4): 241-248.

Birch, H. L., G. A. Rutter, et al. (1997). "Oxidative energy metabolism in equine tendon cells." Research in Veterinary Science 62(2): 93-97.

Birch, H. L., A. M. Wilson, et al. (1997). "The effect of exercise-induced localised hyperthermia on tendon cell survival." J Exp Biol 200(Pt 11): 1703-1708.

Bjordal, J. M., R. A. Lopes-Martins, et al. (2006). "A randomised, placebo controlled trial of low level laser therapy for activated Achilles tendinitis with microdialysis measurement of peritendinous prostaglandin E2 concentrations." Br J Sports Med 40(1): 76-80; discussion 76-80.

Blair, B., A. S. Rokito, et al. (1996). "Efficacy of injections of corticosteroids for subacromial impingement syndrome." J Bone Joint Surg Am 78(11): 1685-1689.

Bomler, J. and J. Sturup (1989). "Achilles tendon rupture. An 8-year follow up." Acta Orthop Belg 55(3): 307-310.

Bonassar, L. J., A. J. Grodzinsky, et al. (2000). "Mechanical and physicochemical regulation of the action of insulin-like growth factor-I on articular cartilage." Arch Biochem Biophys 379(1): 57-63.

Bradley, J. P. and J. E. Tibone (1990). "Percutaneous and open surgical repairs of Achilles tendon ruptures. A comparative study." Am J Sports Med 18(2): 188-195.

Braembussche, P., J. Swevers, et al. (1996). "Accurate tracking control of linear synchronous motor machine tool axes." Mechatronics 6(5): 507-521.

Bramono, D. S., J. C. Richmond, et al. (2004). "Matrix metalloproteinases and their clinical applications in orthopaedics." Clin Orthop Relat Res(428): 272-285.

201

Bray, R. C., R. M. Rangayyan, et al. (1996). "Normal and healing ligament vascularity: a quantitative histological assessment in the adult rabbit medial collateral ligament." J Anat 188 ( Pt 1): 87-95.

Bressel, E. and P. J. McNair (2001). "Biomechanical behavior of the plantar flexor muscle-tendon unit after an Achilles tendon rupture." Am J Sports Med 29(3): 321-326.

Brown, C. H., Jr. and E. W. Carson (1999). Revision anterior cruciate ligament surgery.

Buckwalter, J. A. and E. B. Hunziker (1996). "Orthopaedics. Healing of bones, cartilages, tendons, and ligaments: a new era." Lancet 348 Suppl 2: sII18.

Bugg, E. I., Jr. and B. M. Boyd (1968). "Repair of neglected rupture or laceration of the Achilles tendon." Clin Orthop Relat Res 56: 73-75.

Burkhardt, A., J. Walther, et al. (2012). "Endoscopic optical coherence tomography device for forward imaging with broad field of view." J Biomed Opt 17(7): 071302.

Butler, D. L., C. Gooch, et al. (2010). "The use of mesenchymal stem cells in collagen-based scaffolds for tissue-engineered repair of tendons." Nat Protoc 5(5): 849-863.

Butler, D. L., S. A. Hunter, et al. (2009). "Using functional tissue engineering and bioreactors to mechanically stimulate tissue-engineered constructs." Tissue Eng Part A 15(4): 741-749.

Butler, D. L., N. Juncosa-Melvin, et al. (2008). "Functional tissue engineering for tendon repair: A multidisciplinary strategy using mesenchymal stem cells, bioscaffolds, and mechanical stimulation." J Orthop Res 26(1): 1-9.

Caliari, S. R., D. W. Weisgerber, et al. (2012). "The influence of collagen-glycosaminoglycan scaffold relative density and microstructural anisotropy on tenocyte bioactivity and transcriptomic stability." J Mech Behav Biomed Mater 11: 27-40.

Calve, S., I. F. Lytle, et al. (2010). "Implantation increases tensile strength and collagen content of self-assembled tendon constructs." J Appl Physiol 108(4): 875-881.

Carr, A. J. and S. H. Norris (1989). "The blood supply of the calcaneal tendon." J Bone Joint Surg Br 71(1): 100-101.

Cartmell, S. H., B. D. Porter, et al. (2003). "Effects of medium perfusion rate on cell-seeded three-dimensional bone constructs in vitro." Tissue Eng 9(6): 1197-1203.

Catapano, G., J. F. Patzer, 2nd, et al. (2010). "Transport advances in disposable bioreactors for liver tissue engineering." Adv Biochem Eng Biotechnol 115:

202

117-143.

Cerullo, G., G. Puddu, et al. (1995). "Anterior cruciate ligament patellar tendon reconstruction: it is probably better to leave the tendon defect open!" Knee Surg Sports Traumatol Arthrosc 3(1): 14-17.

Cetti, R. and S. E. Christensen (1983). "Surgical treatment under local anesthesia of Achilles tendon rupture." Clin Orthop Relat Res(173): 204-208.

Cetti, R., L. O. Henriksen, et al. (1994). "A new treatment of ruptured Achilles tendons. A prospective randomized study." Clin Orthop Relat Res(308): 155-165.

Chamberlain, C. S., E. Crowley, et al. (2009). "The spatio-temporal dynamics of ligament healing." Wound Repair Regen 17(2): 206-215.

Chan, B. P., K. M. Chan, et al. (1997). "Effect of basic fibroblast growth factor. An in vitro study of tendon healing." Clin Orthop Relat Res(342): 239-247.

Chan, B. P., S. Fu, et al. (2000). "Effects of basic fibroblast growth factor (bFGF) on early stages of tendon healing: a rat patellar tendon model." Acta Orthop Scand 71(5): 513-518.

Chan, B. P., S. C. Fu, et al. (2006). "Supplementation-time dependence of growth factors in promoting tendon healing." Clin Orthop Relat Res 448: 240-247.

Chan, K. M., S. C. Fu, et al. (2008). "Expression of transforming growth factor beta isoforms and their roles in tendon healing." Wound Repair Regen 16(3): 399-407.

Chang, J., D. Most, et al. (1997). "Gene expression of transforming growth factor beta-1 in rabbit zone II flexor tendon wound healing: evidence for dual mechanisms of repair." Plast Reconstr Surg 100(4): 937-944.

Chang, J., R. Thunder, et al. (2000). "Studies in flexor tendon wound healing: neutralizing antibody to TGF-beta1 increases postoperative range of motion." Plast Reconstr Surg 105(1): 148-155.

Chard, M. D., L. M. Sattelle, et al. (1988). "The long-term outcome of rotator cuff tendinitis--a review study." Br J Rheumatol 27(5): 385-389.

Chen, J., A. Wang, et al. (2010). "In chronic lateral epicondylitis, apoptosis and autophagic cell death occur in the extensor carpi radialis brevis tendon." J Shoulder Elbow Surg 19(3): 355-362.

Chen, J., J. Xu, et al. (2009). "Scaffolds for tendon and ligament repair: review of the efficacy of commercial products." Expert Rev Med Devices 6(1): 61-73.

Chen, J., Q. Yu, et al. (2011). "Autologous tenocyte therapy for experimental Achilles tendinopathy in a rabbit model." Tissue Eng Part A 17(15-16): 2037-2048.

203

Chen, J. L., Z. Yin, et al. (2010). "Efficacy of hESC-MSCs in knitted silk-collagen scaffold for tendon tissue engineering and their roles." Biomaterials 31(36): 9438-9451.

Chen, J. M., C. Willers, et al. (2007). "Autologous tenocyte therapy using porcine-derived bioscaffolds for massive rotator cuff defect in rabbits." Tissue Eng 13(7): 1479-1491.

Chen, X., Y. Y. Qi, et al. (2008). "Ligament regeneration using a knitted silk scaffold combined with collagen matrix." Biomaterials 29(27): 3683-3692.

Chen, X., X. H. Song, et al. (2009). "Stepwise differentiation of human embryonic stem cells promotes tendon regeneration by secreting fetal tendon matrix and differentiation factors." Stem Cells 27(6): 1276-1287.

Chen, Y. E. (2004). "MMP-12, an old enzyme plays a new role in the pathogenesis of rheumatoid arthritis?" Am J Pathol 165(4): 1069-1070.

Chhabra, A., D. Tsou, et al. (2003). "GDF-5 deficiency in mice delays Achilles tendon healing." J Orthop Res 21(5): 826-835.

Choi, H. R., S. Kondo, et al. (2002). "Expression and enzymatic activity of MMP-2 during healing process of the acute supraspinatus tendon tear in rabbits." J Orthop Res 20(5): 927-933.

Chong, A. K., A. D. Ang, et al. (2007). "Bone marrow-derived mesenchymal stem cells influence early tendon-healing in a rabbit achilles tendon model." J Bone Joint Surg Am 89(1): 74-81.

Christensen, I. (1953). "Rupture of the Achilles tendon; analysis of 57 cases." Acta Chir Scand 106(1): 50-60.

Clark, R. T., T. L. Johnson, et al. (2001). "GDF-5 deficiency in mice leads to disruption of tail tendon form and function." Connect Tissue Res 42(3): 175-186.

Connell, D., A. Datir, et al. (2009). "Treatment of lateral epicondylitis using skin-derived tenocyte-like cells." Br J Sports Med 43(4): 293-298.

Cook, J. L. and C. Purdam (2012). "Is compressive load a factor in the development of tendinopathy?" Br J Sports Med 46(3): 163-168.

Coombes, B. K., L. Bisset, et al. (2010). "Efficacy and safety of corticosteroid injections and other injections for management of tendinopathy: a systematic review of randomised controlled trials." Lancet 376(9754): 1751-1767.

Cooper, J. A., H. H. Lu, et al. (2005). "Fiber-based tissue-engineered scaffold for ligament replacement: design considerations and in vitro evaluation." Biomaterials 26(13): 1523-1532.

204

Coupens, S. D., C. K. Yates, et al. (1992). "Magnetic resonance imaging evaluation of the patellar tendon after use of its central one-third for anterior cruciate ligament reconstruction." Am J Sports Med 20(3): 332-335.

Cretnik, A., M. Kosanovic, et al. (2005). "Percutaneous versus open repair of the ruptured Achilles tendon: a comparative study." Am J Sports Med 33(9): 1369-1379.

Cribb, A. M. and J. E. Scott (1995). "Tendon response to tensile stress: an ultrastructural investigation of collagen:proteoglycan interactions in stressed tendon." J Anat 187 ( Pt 2): 423-428.

Dahlgren, L. A., H. O. Mohammed, et al. (2005). "Temporal expression of growth factors and matrix molecules in healing tendon lesions." J Orthop Res 23(1): 84-92.

Dahlgren, L. A., A. J. Nixon, et al. (2001). "Effects of beta-aminopropionitrile on equine tendon metabolism in vitro and on effects of insulin-like growth factor-I on matrix production by equine tenocytes." Am J Vet Res 62(10): 1557-1562.

Dahlgren, L. A., M. C. van der Meulen, et al. (2002). "Insulin-like growth factor-I improves cellular and molecular aspects of healing in a collagenase-induced model of flexor tendinitis." J Orthop Res 20(5): 910-919.

Darlington, L. G. and E. N. Coomes (1977). "The effects of local steroid injection for supraspinatus tears." Rheumatol Rehabil 16(3): 172-179.

Day, R. E., S. Megson, et al. (2005). "Iontophoresis as a means of delivering antibiotics into allograft bone." J Bone Joint Surg Br 87(11): 1568-1574.

De Bari, C., F. Dell'Accio, et al. (2001). "Multipotent mesenchymal stem cells from adult human synovial membrane." Arthritis Rheum 44(8): 1928-1942.

de Vos, R. J., A. Weir, et al. (2010). "Platelet-rich plasma injection for chronic Achilles tendinopathy: a randomized controlled trial." JAMA 303(2): 144-149.

Deng, D., W. Liu, et al. (2009). "Engineering human neo-tendon tissue in vitro with human dermal fibroblasts under static mechanical strain." Biomaterials 30(35): 6724-6730.

Dennis, R. G., B. Smith, et al. (2009). "Bioreactors for guiding muscle tissue growth and development." Adv Biochem Eng Biotechnol 112: 39-79.

Derwin, K. A., A. R. Baker, et al. (2006). "Commercial extracellular matrix scaffolds for rotator cuff tendon repair. Biomechanical, biochemical, and cellular properties." J Bone Joint Surg Am 88(12): 2665-2672.

Domm, C., M. Schunke, et al. (2002). "Redifferentiation of dedifferentiated bovine articular chondrocytes in alginate culture under low oxygen tension."

205

Osteoarthritis Cartilage 10(1): 13-22.

Donnelly, K., A. Khodabukus, et al. (2010). "A novel bioreactor for stimulating skeletal muscle in vitro." Tissue Eng Part C Methods 16(4): 711-718.

Doroski, D. M., M. E. Levenston, et al. (2010). "Cyclic tensile culture promotes fibroblastic differentiation of marrow stromal cells encapsulated in poly(ethylene glycol)-based hydrogels." Tissue Eng Part A 16(11): 3457-3466.

Drexler, W., U. Morgner, et al. (2001). "Ultrahigh-resolution ophthalmic optical coherence tomography." Nat Med 7(4): 502-507.

Drexler, W., U. Morgner, et al. (1999). "In vivo ultrahigh-resolution optical coherence tomography." Opt Lett 24(17): 1221-1223.

Ducy, P. and G. Karsenty (2000). "The family of bone morphogenetic proteins." Kidney Int 57(6): 2207-2214.

Dunphy, J. E. (1967). "The healing of wounds." Can J Surg 10(3): 281-287.

Egerbacher, M., S. P. Arnoczky, et al. (2008). "Loss of homeostatic tension induces apoptosis in tendon cells: an in vitro study." Clin Orthop Relat Res 466(7): 1562-1568.

Eliasson, P., T. Andersson, et al. (2011). "Achilles tendon healing in rats is improved by intermittent mechanical loading during the inflammatory phase." J Orthop Res 30(2): 274-279.

Enwemeka, C. S., N. I. Spielholz, et al. (1988). "The effect of early functional activities on experimentally tenotomized Achilles tendons in rats." Am J Phys Med Rehabil 67(6): 264-269.

Fahlstrom, M., P. Jonsson, et al. (2003). "Chronic Achilles tendon pain treated with eccentric calf-muscle training." Knee Surg Sports Traumatol Arthrosc 11(5): 327-333.

Fan, H., H. Liu, et al. (2009). "Anterior cruciate ligament regeneration using mesenchymal stem cells and silk scaffold in large animal model." Biomaterials 30(28): 4967-4977.

Fan, H., H. Liu, et al. (2008). "In vivo study of anterior cruciate ligament regeneration using mesenchymal stem cells and silk scaffold." Biomaterials 29(23): 3324-3337.

Fan, J., R. R. Varshney, et al. (2009). "Synovium-derived mesenchymal stem cells: a new cell source for musculoskeletal regeneration." Tissue Eng Part B Rev 15(1): 75-86.

Ferrero, G., E. Fabbro, et al. (2012). "Ultrasound-guided injection of platelet-rich

206

plasma in chronic Achilles and patellar tendinopathy." J Ultrasound 15(4): 260-266.

Filardo, G., M. L. Presti, et al. (2010). "Nonoperative biological treatment approach for partial Achilles tendon lesion." Orthopedics 33(2): 120-123.

Finnoff, J. T., S. P. Fowler, et al. (2011). "Treatment of chronic tendinopathy with ultrasound-guided needle tenotomy and platelet-rich plasma injection." PM R 3(10): 900-911.

Fleming, B. C., K. P. Spindler, et al. (2009). "Collagen-platelet composites improve the biomechanical properties of healing anterior cruciate ligament grafts in a porcine model." Am J Sports Med 37(8): 1554-1563.

Flynn, B. P., A. P. Bhole, et al. (2010). "Mechanical strain stabilizes reconstituted collagen fibrils against enzymatic degradation by mammalian collagenase matrix metalloproteinase 8 (MMP-8)." PLoS One 5(8): e12337.

Ford, L. T. and J. DeBender (1979). "Tendon rupture after local steroid injection." South Med J 72(7): 827-830.

Forslund, C., B. Bylander, et al. (2003). "Indomethacin and celecoxib improve tendon healing in rats." Acta Orthop Scand 74(4): 465-469.

Forslund, C., D. Rueger, et al. (2003). "A comparative dose-response study of cartilage-derived morphogenetic protein (CDMP)-1, -2 and -3 for tendon healing in rats." J Orthop Res 21(4): 617-621.

Franchi, M., A. Trire, et al. (2007). "Collagen structure of tendon relates to function." ScientificWorldJournal 7: 404-420.

Freed, L. E., J. C. Marquis, et al. (1993). "Neocartilage formation in vitro and in vivo using cells cultured on synthetic biodegradable polymers." J Biomed Mater Res 27(1): 11-23.

Freed, L. E., G. Vunjak-Novakovic, et al. (1993). "Cultivation of cell-polymer cartilage implants in bioreactors." J Cell Biochem 51(3): 257-264.

Freeman, J. W., M. D. Woods, et al. (2007). "Tissue engineering of the anterior cruciate ligament using a braid-twist scaffold design." J Biomech 40(9): 2029-2036.

Fu, S. C., Y. P. Wong, et al. (2003). "The roles of bone morphogenetic protein (BMP) 12 in stimulating the proliferation and matrix production of human patellar tendon fibroblasts." Life Sci 72(26): 2965-2974.

Galatz, L., S. Rothermich, et al. (2007). "Development of the supraspinatus tendon-to-bone insertion: localized expression of extracellular matrix and growth factor genes." J Orthop Res 25(12): 1621-1628.

207

Garner, W. L., J. A. McDonald, et al. (1989). "Identification of the collagen-producing cells in healing flexor tendons." Plast Reconstr Surg 83(5): 875-879.

Gaweda, K., M. Tarczynska, et al. (2010). "Treatment of Achilles tendinopathy with platelet-rich plasma." Int J Sports Med 31(8): 577-583.

Gebauer, M., F. T. Beil, et al. (2007). "Mechanical evaluation of different techniques for Achilles tendon repair." Arch Orthop Trauma Surg 127(9): 795-799.

Gilbert, T. W., A. M. Stewart-Akers, et al. (2007). "Degradation and remodeling of small intestinal submucosa in canine Achilles tendon repair." J Bone Joint Surg Am 89(3): 621-630.

Gillespie, H. S. and E. A. George (1969). "Results of surgical repair of spontaneous rupture of the Achilles tendon." J Trauma 9(3): 247-249.

Giombini, A., A. Di Cesare, et al. (2006). "Short-term effectiveness of hyperthermia for supraspinatus tendinopathy in athletes: a short-term randomized controlled study." Am J Sports Med 34(8): 1247-1253.

Goetz, M., S. Thomas, et al. (2008). "Dynamic in vivo imaging of microvasculature and perfusion by miniaturized confocal laser microscopy." Eur Surg Res 41(3): 290-297.

Goldberg, B. A., R. J. Nowinski, et al. (2001). "Outcome of nonoperative management of full-thickness rotator cuff tears." Clin Orthop Relat Res(382): 99-107.

Goldfarb, C. A., R. H. Gelberman, et al. (2007). "Extra-articular steroid injection: early patient response and the incidence of flare reaction." J Hand Surg Am 32(10): 1513-1520.

Goldman, S., R. L. Linscheid, et al. (1969). "Disruptions of the tendo Achillis: analysis of 33 cases." Mayo Clin Proc 44(1): 28-35.

Goodship, A. E., H. L. Birch, et al. (1994). "The pathobiology and repair of tendon and ligament injury." Vet Clin North Am Equine Pract 10(2): 323-349.

Goupille, P., M. I. Jayson, et al. (1998). "Matrix metalloproteinases: the clue to intervertebral disc degeneration?" Spine (Phila Pa 1976) 23(14): 1612-1626.

Grigg, N. L., S. C. Wearing, et al. (2009). "Eccentric calf muscle exercise produces a greater acute reduction in Achilles tendon thickness than concentric exercise." Br J Sports Med 43(4): 280-283.

Hamada, K., Y. Okawara, et al. (1994). "Localization of mRNA of procollagen alpha 1 type I in torn supraspinatus tendons. In situ hybridization using digoxigenin labeled oligonucleotide probe." Clin Orthop Relat Res(304): 18-21.

Hamilton, B. and C. Purdam (2004). "Patellar tendinosis as an adaptive process: a new

208

hypothesis." Br J Sports Med 38(6): 758-761.

Hankemeier, S., M. van Griensven, et al. (2007). "Tissue engineering of tendons and ligaments by human bone marrow stromal cells in a liquid fibrin matrix in immunodeficient rats: results of a histologic study." Arch Orthop Trauma Surg 127(9): 815-821.

Hannafin, J. A. and S. P. Arnoczky (1994). "Effect of cyclic and static tensile loading on water content and solute diffusion in canine flexor tendons: an in vitro study." J Orthop Res 12(3): 350-356.

Hannafin, J. A., S. P. Arnoczky, et al. (1995). "Effect of stress deprivation and cyclic tensile loading on the material and morphologic properties of canine flexor digitorum profundus tendon: an in vitro study." J Orthop Res 13(6): 907-914.

Hansen, U., M. Schunke, et al. (2001). "Combination of reduced oxygen tension and intermittent hydrostatic pressure: a useful tool in articular cartilage tissue engineering." J Biomech 34(7): 941-949.

Harner, C. D., E. Olson, et al. (1996). "Allograft versus autograft anterior cruciate ligament reconstruction: 3- to 5-year outcome." Clin Orthop Relat Res(324): 134-144.

Hart, D. A., J. M. Archambault, et al. (1998). "Gender and neurogenic variables in tendon biology and repetitive motion disorders." Clin Orthop Relat Res(351): 44-56.

Harwood, F. L., R. S. Goomer, et al. (1999). "Regulation of alpha(v)beta3 and alpha5beta1 integrin receptors by basic fibroblast growth factor and platelet-derived growth factor-BB in intrasynovial flexor tendon cells." Wound Repair Regen 7(5): 381-388.

Hattrup, S. J. and K. A. Johnson (1985). "A review of ruptures of the Achilles tendon." Foot Ankle 6(1): 34-38.

Hee, C. K., J. S. Dines, et al. (2011). "Augmentation of a rotator cuff suture repair using rhPDGF-BB and a type I bovine collagen matrix in an ovine model." Am J Sports Med 39(8): 1630-1639.

Hildebrand, K. A., C. B. Frank, et al. (2004). "Gene intervention in ligament and tendon: current status, challenges, future directions." Gene Ther 11(4): 368-378.

Hildebrand, K. A., S. L. Woo, et al. (1998). "The effects of platelet-derived growth factor-BB on healing of the rabbit medial collateral ligament. An in vivo study." Am J Sports Med 26(4): 549-554.

Hoffmeyer, P., C. Freuler, et al. (1990). "Pathological changes in the triceps surae muscle after rupture of the Achilles tendon." Int Orthop 14(2): 183-188.

209

Hogsaa, B., M. Nohr, et al. (1990). "Surgical treatment of Achilles tendon ruptures." Unfallchirurg 93(1): 40-43.

Holtorf, H. L., J. A. Jansen, et al. (2005). "Flow perfusion culture induces the osteoblastic differentiation of marrow stroma cell-scaffold constructs in the absence of dexamethasone." J Biomed Mater Res A 72(3): 326-334.

Hosaka, Y., H. Teraoka, et al. (2005). "Mechanism of cell death in inflamed superficial digital flexor tendon in the horse." J Comp Pathol 132(1): 51-58.

Hou, Y., Z. Mao, et al. (2009). "Effects of transforming growth factor-beta1 and vascular endothelial growth factor 165 gene transfer on Achilles tendon healing." Matrix Biol 28(6): 324-335.

Hoying, J. B. and S. K. Williams (1996). "Effects of basic fibroblast growth factor on human microvessel endothelial cell migration on collagen I correlates inversely with adhesion and is cell density dependent." J Cell Physiol 168(2): 294-304.

Huang, A. H., M. J. Farrell, et al. (2010). "Long-term dynamic loading improves the mechanical properties of chondrogenic mesenchymal stem cell-laden hydrogel." Eur Cell Mater 19: 72-85.

Huang, H. H., A. A. Qureshi, et al. (2000). "Sports and other soft tissue injuries, tendinitis, bursitis, and occupation-related syndromes." Curr Opin Rheumatol 12(2): 150-154.

Husemann, B., F. Muller, et al. (2007). "Effects of locomotion training with assistance of a robot-driven gait orthosis in hemiparetic patients after stroke: a randomized controlled pilot study." Stroke 38(2): 349-354.

Inglis, A. E. and T. P. Sculco (1981). "Surgical repair of ruptures of the tendo Achillis." Clin Orthop Relat Res(156): 160-169.

Ireland, D., R. Harrall, et al. (2001). "Multiple changes in gene expression in chronic human Achilles tendinopathy." Matrix Biol 20(3): 159-169.

Issa, R. I., B. Engebretson, et al. (2011). "The effect of cell seeding density on the cellular and mechanical properties of a mechanostimulated tissue-engineered tendon." Tissue Eng Part A 17(11-12): 1479-1487.

James, R., G. Kesturu, et al. (2008). "Tendon: biology, biomechanics, repair, growth factors, and evolving treatment options." J Hand Surg Am 33(1): 102-112.

Janssen, F. W., J. Oostra, et al. (2006). "A perfusion bioreactor system capable of producing clinically relevant volumes of tissue-engineered bone: in vivo bone formation showing proof of concept." Biomaterials 27(3): 315-323.

Jarvinen, M., L. Jozsa, et al. (1997). "Histopathological findings in chronic tendon disorders." Scand J Med Sci Sports 7(2): 86-95.

210

Jarvinen, T. A., T. L. Jarvinen, et al. (2004). "Collagen fibres of the spontaneously ruptured human tendons display decreased thickness and crimp angle." J Orthop Res 22(6): 1303-1309.

Jiang, Y., H. Liu, et al. (2011). "A proteomic analysis of engineered tendon formation under dynamic mechanical loading in vitro." Biomaterials 32(17): 4085-4095.

Jones, G. C., A. N. Corps, et al. (2006). "Expression profiling of metalloproteinases and tissue inhibitors of metalloproteinases in normal and degenerate human achilles tendon." Arthritis Rheum 54(3): 832-842.

Ju, Y. J., T. Muneta, et al. (2008). "Synovial mesenchymal stem cells accelerate early remodeling of tendon-bone healing." Cell Tissue Res 332(3): 469-478.

Juncosa-Melvin, N., J. T. Shearn, et al. (2006). "Effects of mechanical stimulation on the biomechanics and histology of stem cell-collagen sponge constructs for rabbit patellar tendon repair." Tissue Eng 12(8): 2291-2300.

Kabata, T., T. Kubo, et al. (2000). "Apoptotic cell death in steroid induced osteonecrosis: an experimental study in rabbits." J Rheumatol 27(9): 2166-2171.

Kannus, P. (2000). "Structure of the tendon connective tissue." Scand J Med Sci Sports 10(6): 312-320.

Kannus, P. and L. Jozsa (1991). "Histopathological changes preceding spontaneous rupture of a tendon. A controlled study of 891 patients." J Bone Joint Surg Am 73(10): 1507-1525.

Kannus, P., L. Jozsa, et al. (1997). "Effects of training, immobilization and remobilization on tendons." Scand J Med Sci Sports 7(2): 67-71.

Kannus, P. and A. Natri (1997). "Etiology and pathophysiology of tendon ruptures in sports." Scand J Med Sci Sports 7(2): 107-112.

Kashiwagi, K., Y. Mochizuki, et al. (2004). "Effects of transforming growth factor-beta 1 on the early stages of healing of the Achilles tendon in a rat model." Scand J Plast Reconstr Surg Hand Surg 38(4): 193-197.

Kastelic, J., A. Galeski, et al. (1978). "The multicomposite structure of tendon." Connect Tissue Res 6(1): 11-23.

Kastelic, J., I. Palley, et al. (1980). "A structural mechanical model for tendon crimping." J Biomech 13(10): 887-893.

Ker, R. F. (1981). "Dynamic tensile properties of the plantaris tendon of sheep (Ovis aries)." J Exp Biol 93: 283-302.

Khan, K. and J. Cook (2003). "The painful nonruptured tendon: clinical aspects." Clin Sports Med 22(4): 711-725.

211

Khan, K. M., J. L. Cook, et al. (1999). "Histopathology of common tendinopathies. Update and implications for clinical management." Sports Med 27(6): 393-408.

Khan, K. M. and N. Maffulli (1998). "Tendinopathy: an Achilles' heel for athletes and clinicians." Clin J Sport Med 8(3): 151-154.

Khan, K. M. and A. Scott (2009). "Mechanotherapy: how physical therapists' prescription of exercise promotes tissue repair." Br J Sports Med 43(4): 247-252.

Khayat, G., D. H. Rosenzweig, et al. (2012). "Low frequency mechanical stimulation inhibits adipogenic differentiation of C3H10T1/2 mesenchymal stem cells." Differentiation 83(4): 179-184.

Kim, H. J., S. W. Kang, et al. (2007). "The role of transforming growth factor-beta and bone morphogenetic protein with fibrin glue in healing of bone-tendon junction injury." Connect Tissue Res 48(6): 309-315.

Kim, H. M., L. M. Galatz, et al. (2011). "The role of transforming growth factor beta isoforms in tendon-to-bone healing." Connect Tissue Res 52(2): 87-98.

Kingma, J. J., R. de Knikker, et al. (2007). "Eccentric overload training in patients with chronic Achilles tendinopathy: a systematic review." Br J Sports Med 41(6): e3.

Kinneberg, K. R., V. S. Nirmalanandhan, et al. (2010). "Chondroitin-6-sulfate incorporation and mechanical stimulation increase MSC-collagen sponge construct stiffness." J Orthop Res 28(8): 1092-1099.

Kleinman, M. and A. E. Gross (1983). "Achilles tendon rupture following steroid injection. Report of three cases." J Bone Joint Surg Am 65(9): 1345-1347.

Knauper, V., C. Lopez-Otin, et al. (1996). "Biochemical characterization of human collagenase-3." J Biol Chem 271(3): 1544-1550.

Kobayashi, M., E. Itoi, et al. (2006). "Expression of growth factors in the early phase of supraspinatus tendon healing in rabbits." J Shoulder Elbow Surg 15(3): 371-377.

Komi, P. V., S. Fukashiro, et al. (1992). "Biomechanical loading of Achilles tendon during normal locomotion." Clin Sports Med 11(3): 521-531.

Kong, A., A. Van der Vliet, et al. (2007). "MRI and US of gluteal tendinopathy in greater trochanteric pain syndrome." Eur Radiol 17(7): 1772-1783.

Kovacevic, D., A. J. Fox, et al. (2011). "Calcium-phosphate matrix with or without TGF-beta3 improves tendon-bone healing after rotator cuff repair." Am J Sports Med 39(4): 811-819.

Krane, S. M., M. H. Byrne, et al. (1996). "Different collagenase gene products have different roles in degradation of type I collagen." J Biol Chem 271(45): 28509-28515.

212

Krueger-Franke, M., C. H. Siebert, et al. (1995). "Surgical treatment of ruptures of the Achilles tendon: a review of long-term results." Br J Sports Med 29(2): 121-125.

Kuhn, K. (1969). "The structure of collagen." Essays Biochem 5: 59-87.

Kulick, M. I., S. Smith, et al. (1986). "Oral ibuprofen: evaluation of its effect on peritendinous adhesions and the breaking strength of a tenorrhaphy." J Hand Surg Am 11(1): 110-120.

Kupcsik, L., M. J. Stoddart, et al. (2010). "Improving chondrogenesis: potential and limitations of SOX9 gene transfer and mechanical stimulation for cartilage tissue engineering." Tissue Eng Part A 16(6): 1845-1855.

Langberg, H., D. Skovgaard, et al. (1999). "Type I collagen synthesis and degradation in peritendinous tissue after exercise determined by microdialysis in humans." J Physiol 521 Pt 1: 299-306.

Larsen, N. P., M. R. Forwood, et al. (1987). "Immobilization and retraining of cruciate ligaments in the rat." Acta Orthop Scand 58(3): 260-264.

Lavagnino, M., S. P. Arnoczky, et al. (2006). "Isolated fibrillar damage in tendons stimulates local collagenase mRNA expression and protein synthesis." J Biomech 39(13): 2355-2362.

Lavagnino, M., S. P. Arnoczky, et al. (2003). "Effect of amplitude and frequency of cyclic tensile strain on the inhibition of MMP-1 mRNA expression in tendon cells: an in vitro study." Connect Tissue Res 44(3-4): 181-187.

Leach, R. E., A. A. Schepsis, et al. (1992). "Long-term results of surgical management of Achilles tendinitis in runners." Clin Orthop Relat Res(282): 208-212.

Leadbetter, W. B. (1992). "Cell-matrix response in tendon injury." Clin Sports Med 11(3): 533-578.

Lee, C. H., H. J. Shin, et al. (2005). "Nanofiber alignment and direction of mechanical strain affect the ECM production of human ACL fibroblast." Biomaterials 26(11): 1261-1270.

Lejard, V., G. Brideau, et al. (2007). "Scleraxis and NFATc regulate the expression of the pro-alpha1(I) collagen gene in tendon fibroblasts." J Biol Chem 282(24): 17665-17675.

Leppilahti, J. and S. Orava (1998). "Total Achilles tendon rupture. A review." Sports Med 25(2): 79-100.

Li, G., L. E. Defrate, et al. (2005). "In vivo kinematics of the ACL during weight-bearing knee flexion." J Orthop Res 23(2): 340-344.

Lian, O., A. Scott, et al. (2007). "Excessive apoptosis in patellar tendinopathy in

213

athletes." Am J Sports Med 35(4): 605-611.

Lin, T. W., L. Cardenas, et al. (2004). "Biomechanics of tendon injury and repair." J Biomech 37(6): 865-877.

Lindsay, W. K. and J. R. Birch (1964). "The Fibroblast in Flexor Tendon Healing." Plast Reconstr Surg 34: 223-232.

Liu, J., M. Hsu, et al. (2001). "On the design of rotating speed functions to improve the acceleration peak value of ball-screw transmission mechanism." Mechanism and Machine Theory 36(9): 1035-1049.

Liu, S. H., R. S. Yang, et al. (1995). "Collagen in tendon, ligament, and bone healing. A current review." Clin Orthop Relat Res(318): 265-278.

Liu, W., B. Chen, et al. (2006). "Repair of tendon defect with dermal fibroblast engineered tendon in a porcine model." Tissue Eng 12(4): 775-788.

Lo, A. C. and E. W. Triche (2008). "Improving gait in multiple sclerosis using robot-assisted, body weight supported treadmill training." Neurorehabil Neural Repair 22(6): 661-671.

Lo, I. K., L. L. Marchuk, et al. (2004). "Matrix metalloproteinase and tissue inhibitor of matrix metalloproteinase mRNA levels are specifically altered in torn rotator cuff tendons." Am J Sports Med 32(5): 1223-1229.

Lo, I. K., Y. Ou, et al. (2002). "The cellular networks of normal ovine medial collateral and anterior cruciate ligaments are not accurately recapitulated in scar tissue." J Anat 200(Pt 3): 283-296.

Loitz, B. J., R. F. Zernicke, et al. (1989). "Effects of short-term immobilization versus continuous passive motion on the biomechanical and biochemical properties of the rabbit tendon." Clin Orthop Relat Res(244): 265-271.

Lou, J., Y. Tu, et al. (2001). "BMP-12 gene transfer augmentation of lacerated tendon repair." J Orthop Res 19(6): 1199-1202.

Ma, G. W. and T. G. Griffith (1977). "Percutaneous repair of acute closed ruptured achilles tendon: a new technique." Clin Orthop Relat Res(128): 247-255.

Machner, A., A. Baier, et al. (2003). "Higher susceptibility to Fas ligand induced apoptosis and altered modulation of cell death by tumor necrosis factor-alpha in periarticular tenocytes from patients with knee joint osteoarthritis." Arthritis Res Ther 5(5): R253-261.

Maeda, T., T. Sakabe, et al. (2011). "Conversion of mechanical force into TGF-beta-mediated biochemical signals." Curr Biol 21(11): 933-941.

Maffulli, N. (1995). "Ultrasound of the Achilles tendon after surgical repair:

214

morphology and function." Br J Radiol 68(816): 1372-1373.

Maffulli, N. (1998). "Current concepts in the management of subcutaneous tears of the Achilles tendon." Bull Hosp Jt Dis 57(3): 152-158.

Maffulli, N., K. M. Khan, et al. (1998). "Overuse tendon conditions: time to change a confusing terminology." Arthroscopy 14(8): 840-843.

Maffulli, N., P. Sharma, et al. (2004). "Achilles tendinopathy: aetiology and management." J R Soc Med 97(10): 472-476.

Maffulli, N., V. Testa, et al. (1997). "Results of percutaneous longitudinal tenotomy for Achilles tendinopathy in middle- and long-distance runners." Am J Sports Med 25(6): 835-840.

Magnusson, S. P., P. Hansen, et al. (2003). "Tendon properties in relation to muscular activity and physical training." Scand J Med Sci Sports 13(4): 211-223.

Magra, M. and N. Maffulli (2005). "Matrix metalloproteases: a role in overuse tendinopathies." Br J Sports Med 39(11): 789-791.

Majewski, M., O. Betz, et al. (2008). "Ex vivo adenoviral transfer of bone morphogenetic protein 12 (BMP-12) cDNA improves Achilles tendon healing in a rat model." Gene Ther 15(16): 1139-1146.

Malcarney, H. L., F. Bonar, et al. (2005). "Early inflammatory reaction after rotator cuff repair with a porcine small intestine submucosal implant: a report of 4 cases." Am J Sports Med 33(6): 907-911.

Manning, C. N., H. M. Kim, et al. (2011). "Sustained delivery of transforming growth factor beta three enhances tendon-to-bone healing in a rat model." J Orthop Res 29(7): 1099-1105.

Manske, P. R. (1988). "Flexor tendon healing." J Hand Surg Br 13(3): 237-245.

Marui, T., C. Niyibizi, et al. (1997). "Effect of growth factors on matrix synthesis by ligament fibroblasts." J Orthop Res 15(1): 18-23.

Matsumoto, F., G. Trudel, et al. (2003). "Mechanical effects of immobilization on the Achilles' tendon." Arch Phys Med Rehabil 84(5): 662-667.

McCawley, L. J. and L. M. Matrisian (2001). "Matrix metalloproteinases: they're not just for matrix anymore!" Curr Opin Cell Biol 13(5): 534-540.

Meinel, L., V. Karageorgiou, et al. (2004). "Bone tissue engineering using human mesenchymal stem cells: effects of scaffold material and medium flow." Ann Biomed Eng 32(1): 112-122.

Mikic, B., B. J. Schalet, et al. (2001). "GDF-5 deficiency in mice alters the

215

ultrastructure, mechanical properties and composition of the Achilles tendon." J Orthop Res 19(3): 365-371.

Millar, N. L., J. H. Reilly, et al. (2012). "Hypoxia: a critical regulator of early human tendinopathy." Ann Rheum Dis 71(2): 302-310.

Millar, N. L., A. Q. Wei, et al. (2009). "Cytokines and apoptosis in supraspinatus tendinopathy." J Bone Joint Surg Br 91(3): 417-424.

Mishra, A., P. Randelli, et al. (2012). "Platelet-rich plasma and the upper extremity." Hand Clin 28(4): 481-491.

Moffat, K. L., A. S. Kwei, et al. (2009). "Novel nanofiber-based scaffold for rotator cuff repair and augmentation." Tissue Eng Part A 15(1): 115-126.

Molloy, T., Y. Wang, et al. (2003). "The roles of growth factors in tendon and ligament healing." Sports Med 33(5): 381-394.

Monto, R. R. (2012). "Platelet rich plasma treatment for chronic Achilles tendinosis." Foot Ankle Int 33(5): 379-385.

Morand, E. F., M. Leech, et al. (2006). "MIF: a new cytokine link between rheumatoid arthritis and atherosclerosis." Nat Rev Drug Discov 5(5): 399-410.

Morrison, D. S., A. D. Frogameni, et al. (1997). "Non-operative treatment of subacromial impingement syndrome." J Bone Joint Surg Am 79(5): 732-737.

Murata, H., H. Nishizono, et al. (2000). "Postnatal development of structure and arrangement of tendon cells. A scanning and transmission electron microscope study in the rat calcaneal tendon." Okajimas Folia Anat Jpn 77(2-3): 77-86.

Myers, B. and M. Wolf (1974). "Vascularization of the healing wound." Am Surg 40(12): 716-722.

Nagase, H., R. Visse, et al. (2006). "Structure and function of matrix metalloproteinases and TIMPs." Cardiovasc Res 69(3): 562-573.

Nakagawa, Y., M. Totsuka, et al. (1989). "Effect of disuse on the ultrastructure of the achilles tendon in rats." Eur J Appl Physiol Occup Physiol 59(3): 239-242.

Nakama, L. H., K. B. King, et al. (2005). "Evidence of tendon microtears due to cyclical loading in an in vivo tendinopathy model." J Orthop Res 23(5): 1199-1205.

Nakamura, N., K. Shino, et al. (1998). "Early biological effect of in vivo gene transfer of platelet-derived growth factor (PDGF)-B into healing patellar ligament." Gene Ther 5(9): 1165-1170.

Nakamura, N., S. A. Timmermann, et al. (1998). "A comparison of in vivo gene delivery methods for antisense therapy in ligament healing." Gene Ther 5(11):

216

1455-1461.

Nathan, C. (2002). "Points of control in inflammation." Nature 420(6917): 846-852.

Nguyen, M. T., H. Lue, et al. (2003). "The cytokine macrophage migration inhibitory factor reduces pro-oxidative stress-induced apoptosis." J Immunol 170(6): 3337-3347.

Nguyen, T. D., R. Liang, et al. (2009). "Effects of cell seeding and cyclic stretch on the fiber remodeling in an extracellular matrix-derived bioscaffold." Tissue Eng Part A 15(4): 957-963.

Ni, M., P. P. Lui, et al. (2012). "Tendon-derived stem cells (TDSCs) promote tendon repair in a rat patellar tendon window defect model." J Orthop Res 30(4): 613-619.

Ni, M., Y. F. Rui, et al. (2013). "Engineered scaffold-free tendon tissue produced by tendon-derived stem cells." Biomaterials 34(8): 2024-2037.

Nirmalanandhan, V. S., M. Rao, et al. (2008). "Effect of scaffold material, construct length and mechanical stimulation on the in vitro stiffness of the engineered tendon construct." J Biomech 41(4): 822-828.

Nirmalanandhan, V. S., J. T. Shearn, et al. (2008). "Improving linear stiffness of the cell-seeded collagen sponge constructs by varying the components of the mechanical stimulus." Tissue Eng Part A 14(11): 1883-1891.

Nistor, L. (1981). "Surgical and non-surgical treatment of Achilles Tendon rupture. A prospective randomized study." J Bone Joint Surg Am 63(3): 394-399.

Nyyssonen, T. and P. Luthje (2000). "Achilles tendon ruptures in South-East Finland between 1986-1996, with special reference to epidemiology, complications of surgery and hospital costs." Ann Chir Gynaecol 89(1): 53-57.

Obaid, H. and D. Connell (2010). "Cell therapy in tendon disorders: what is the current evidence?" Am J Sports Med 38(10): 2123-2132.

Obst, S. J., R. Newsham-West, et al. "In Vivo Measurement of Human Achilles Tendon Morphology Using Freehand 3-D Ultrasound." Ultrasound Med Biol.

Ohberg, L. and H. Alfredson (2003). "Sclerosing therapy in chronic Achilles tendon insertional pain-results of a pilot study." Knee Surg Sports Traumatol Arthrosc 11(5): 339-343.

Olivier, V., P. Hivart, et al. (2007). "In vitro culture of large bone substitutes in a new bioreactor: importance of the flow direction." Biomed Mater 2(3): 174-180.

Ouyang, H. W., J. C. Goh, et al. (2004). "Viability of allogeneic bone marrow stromal cells following local delivery into patella tendon in rabbit model." Cell

217

Transplant 13(6): 649-657.

Owens, R. F., Jr., J. Ginnetti, et al. (2011). "Clinical and magnetic resonance imaging outcomes following platelet rich plasma injection for chronic midsubstance Achilles tendinopathy." Foot Ankle Int 32(11): 1032-1039.

Paavola, M., S. Orava, et al. (2000). "Chronic Achilles tendon overuse injury: complications after surgical treatment. An analysis of 432 consecutive patients." Am J Sports Med 28(1): 77-82.

Parent, G., N. Huppe, et al. (2011). "Low stress tendon fatigue is a relatively rapid process in the context of overuse injuries." Ann Biomed Eng 39(5): 1535-1545.

Parkinson, J., T. Samiric, et al. (2010). "Change in proteoglycan metabolism is a characteristic of human patellar tendinopathy." Arthritis Rheum 62(10): 3028-3035.

Parkinson, J., T. Samiric, et al. (2011). "Involvement of proteoglycans in tendinopathy." J Musculoskelet Neuronal Interact 11(2): 86-93.

Peach, M. S., R. James, et al. (2012). "Polyphosphazene functionalized polyester fiber matrices for tendon tissue engineering: in vitro evaluation with human mesenchymal stem cells." Biomed Mater 7(4): 045016.

Petersen, W., T. Pufe, et al. (2003). "Hypoxia and PDGF have a synergistic effect that increases the expression of the angiogenetic peptide vascular endothelial growth factor in Achilles tendon fibroblasts." Arch Orthop Trauma Surg 123(9): 485-488.

Pham, Q. P., U. Sharma, et al. (2006). "Electrospinning of polymeric nanofibers for tissue engineering applications: a review." Tissue Eng 12(5): 1197-1211.

Pillai, R. S., D. Lorenser, et al. "Deep-tissue access with confocal fluorescence microendoscopy through hypodermic needles." Opt Express 19(8): 7213-7221.

Pryce, B. A., S. S. Watson, et al. (2009). "Recruitment and maintenance of tendon progenitors by TGFbeta signaling are essential for tendon formation." Development 136(8): 1351-1361.

Puddu, G., E. Ippolito, et al. (1976). "A classification of Achilles tendon disease." Am J Sports Med 4(4): 145-150.

Quirk, B. C., R. A. McLaughlin, et al. "In situ imaging of lung alveoli with an optical coherence tomography needle probe." J Biomed Opt 16(3): 036009.

Rauh, J., F. Milan, et al. (2011). "Bioreactor systems for bone tissue engineering." Tissue Eng Part B Rev 17(4): 263-280.

Rauh, J., F. Milan, et al. (2012). "Bioreactor systems for bone tissue engineering."

218

Tissue Eng Part B Rev 17(4): 263-280.

Rees, J. D., A. M. Wilson, et al. (2006). "Current concepts in the management of tendon disorders." Rheumatology (Oxford) 45(5): 508-521.

Reeves, N. D., C. N. Maganaris, et al. (2005). "Influence of 90-day simulated microgravity on human tendon mechanical properties and the effect of resistive countermeasures." J Appl Physiol 98(6): 2278-2286.

Riemersma, D. J. and H. C. Schamhardt (1985). "In vitro mechanical properties of equine tendons in relation to cross-sectional area and collagen content." Research in Veterinary Science 39(3): 263-270.

Riley, G. P., V. Curry, et al. (2002). "Matrix metalloproteinase activities and their relationship with collagen remodelling in tendon pathology." Matrix Biol 21(2): 185-195.

Riley, G. P., R. L. Harrall, et al. (1994). "Tendon degeneration and chronic shoulder pain: changes in the collagen composition of the human rotator cuff tendons in rotator cuff tendinitis." Ann Rheum Dis 53(6): 359-366.

Rolf, C. and T. Movin (1997). "Etiology, histopathology, and outcome of surgery in achillodynia." Foot Ankle Int 18(9): 565-569.

Rompe, J. D., J. Furia, et al. (2008). "Eccentric loading compared with shock wave treatment for chronic insertional achilles tendinopathy. A randomized, controlled trial." J Bone Joint Surg Am 90(1): 52-61.

Roos, E. M., M. Engstrom, et al. (2004). "Clinical improvement after 6 weeks of eccentric exercise in patients with mid-portion Achilles tendinopathy -- a randomized trial with 1-year follow-up." Scand J Med Sci Sports 14(5): 286-295.

Ross, J. J., M. J. Duxson, et al. (1987). "Neural determination of muscle fibre numbers in embryonic rat lumbrical muscles." Development 100(3): 395-409.

Ruedi, T. P. and M. Allgower (1979). "The operative treatment of intra-articular fractures of the lower end of the tibia." Clin Orthop Relat Res(138): 105-110.

Rui, Y. F., P. P. Lui, et al. (2011). "Mechanical loading increased BMP-2 expression which promoted osteogenic differentiation of tendon-derived stem cells." J Orthop Res 29(3): 390-396.

Saber, S., A. Y. Zhang, et al. (2010). "Flexor tendon tissue engineering: bioreactor cyclic strain increases construct strength." Tissue Eng Part A 16(6): 2085-2090.

Sahoo, S., J. G. Cho-Hong, et al. (2007). "Development of hybrid polymer scaffolds for potential applications in ligament and tendon tissue engineering." Biomed Mater 2(3): 169-173.

219

Sahoo, S., H. Ouyang, et al. (2006). "Characterization of a novel polymeric scaffold for potential application in tendon/ligament tissue engineering." Tissue Eng 12(1): 91-99.

Sahoo, S., S. L. Toh, et al. (2010). "A bFGF-releasing silk/PLGA-based biohybrid scaffold for ligament/tendon tissue engineering using mesenchymal progenitor cells." Biomaterials 31(11): 2990-2998.

Saini, S. and T. M. Wick (2004). "Effect of low oxygen tension on tissue-engineered cartilage construct development in the concentric cylinder bioreactor." Tissue Eng 10(5-6): 825-832.

Salter, E., B. Goh, et al. (2012). "Bone tissue engineering bioreactors: a role in the clinic?" Tissue Eng Part B Rev 18(1): 62-75.

Samilson, R. L. and W. F. Binder (1975). "Symptomatic full thickness tears of rotator cuff. An analysis of 292 shoulders in 276 patients." Orthop Clin North Am 6(2): 449-466.

Sanchez, M., E. Anitua, et al. (2007). "Comparison of surgically repaired Achilles tendon tears using platelet-rich fibrin matrices." Am J Sports Med 35(2): 245-251.

Schechtman, H. and D. L. Bader (1997). "In vitro fatigue of human tendons." J Biomech 30(8): 829-835.

Schepsis, A. A., H. Jones, et al. (2002). "Achilles tendon disorders in athletes." Am J Sports Med 30(2): 287-305.

Schepull, T., J. Kvist, et al. (2011). "Autologous platelets have no effect on the healing of human achilles tendon ruptures: a randomized single-blind study." Am J Sports Med 39(1): 38-47.

Schneewind, J. H., I. K. Kline, et al. (1964). "The Role of Paratenon in Healing of Experimental Tendon Transplants." J Occup Med 6: 429-436.

Schwartz, I., A. Sajin, et al. (2012). "Robot-assisted gait training in multiple sclerosis patients: a randomized trial." Mult Scler 18(6): 881-890.

Scott, J. E. (1984). "The periphery of the developing collagen fibril. Quantitative relationships with dermatan sulphate and other surface-associated species." Biochem J 218(1): 229-233.

Screen, H. R., V. H. Chhaya, et al. (2006). "The influence of swelling and matrix degradation on the microstructural integrity of tendon." Acta Biomater 2(5): 505-513.

Screen, H. R., D. A. Lee, et al. (2004). "An investigation into the effects of the hierarchical structure of tendon fascicles on micromechanical properties." Proc

220

Inst Mech Eng H 218(2): 109-119.

Sen, B., Z. Xie, et al. (2008). "Mechanical strain inhibits adipogenesis in mesenchymal stem cells by stimulating a durable beta-catenin signal." Endocrinology 149(12): 6065-6075.

Shalabi, A., M. Kristoffersen-Wilberg, et al. (2004). "Eccentric training of the gastrocnemius-soleus complex in chronic Achilles tendinopathy results in decreased tendon volume and intratendinous signal as evaluated by MRI." Am J Sports Med 32(5): 1286-1296.

Sharifpoor, S., C. A. Simmons, et al. (2011). "Functional characterization of human coronary artery smooth muscle cells under cyclic mechanical strain in a degradable polyurethane scaffold." Biomaterials 32(21): 4816-4829.

Shi, Y., H. Li, et al. (2011). "Continuous cyclic mechanical tension inhibited Runx2 expression in mesenchymal stem cells through RhoA-ERK1/2 pathway." J Cell Physiol 226(8): 2159-2169.

Silbernagel, K. G., R. Thomee, et al. (2001). "Eccentric overload training for patients with chronic Achilles tendon pain--a randomised controlled study with reliability testing of the evaluation methods." Scand J Med Sci Sports 11(4): 197-206.

Skutek, M., M. van Griensven, et al. (2001). "Cyclic mechanical stretching enhances secretion of Interleukin 6 in human tendon fibroblasts." Knee Surg Sports Traumatol Arthrosc 9(5): 322-326.

Skutek, M., M. van Griensven, et al. (2001). "Cyclic mechanical stretching modulates secretion pattern of growth factors in human tendon fibroblasts." Eur J Appl Physiol 86(1): 48-52.

Skutek, M., M. van Griensven, et al. (2003). "Cyclic mechanical stretching of human patellar tendon fibroblasts: activation of JNK and modulation of apoptosis." Knee Surg Sports Traumatol Arthrosc 11(2): 122-129.

Smith, K. D., A. Vaughan-Thomas, et al. (2011). "The organisation of elastin and fibrillins 1 and 2 in the cruciate ligament complex." J Anat 218(6): 600-607.

Smith, L. L. (2000). "Cytokine hypothesis of overtraining: a physiological adaptation to excessive stress?" Med Sci Sports Exerc 32(2): 317-331.

Smith, M. M., G. Sakurai, et al. (2008). "Modulation of aggrecan and ADAMTS expression in ovine tendinopathy induced by altered strain." Arthritis Rheum 58(4): 1055-1066.

Smithpeter, C. L., A. K. Dunn, et al. (1998). "Penetration depth limits of in vivo confocal reflectance imaging." Appl Opt 37(13): 2749-2754.

So, V. and H. Pollard (1997). "Management of Achilles tendon disorders. A case

221

review." Australas Chiropr Osteopathy 6(2): 58-62.

Sporn, M. B., A. B. Roberts, et al. (1986). "Transforming growth factor-beta: biological function and chemical structure." Science 233(4763): 532-534.

Stanish, W. D., R. M. Rubinovich, et al. (1986). "Eccentric exercise in chronic tendinitis." Clin Orthop Relat Res(208): 65-68.

Stoll, C., T. John, et al. (2010). "Extracellular matrix expression of human tenocytes in three-dimensional air-liquid and PLGA cultures compared with tendon tissue: implications for tendon tissue engineering." J Orthop Res 28(9): 1170-1177.

Straudi, S., M. G. Benedetti, et al. (2013). "Does robot-assisted gait training ameliorate gait abnormalities in multiple sclerosis? A pilot randomized-control trial." NeuroRehabilitation.

Suchak, A. A., G. Bostick, et al. (2005). "The incidence of Achilles tendon ruptures in Edmonton, Canada." Foot Ankle Int 26(11): 932-936.

Takahasih, S., M. Nakajima, et al. (2002). "Effect of recombinant basic fibroblast growth factor (bFGF) on fibroblast-like cells from human rotator cuff tendon." Tohoku J Exp Med 198(4): 207-214.

Tallon, C., B. D. Coleman, et al. (2001). "Outcome of surgery for chronic Achilles tendinopathy. A critical review." Am J Sports Med 29(3): 315-320.

Tang, J. B., Y. Cao, et al. (2008). "Adeno-associated virus-2-mediated bFGF gene transfer to digital flexor tendons significantly increases healing strength. an in vivo study." J Bone Joint Surg Am 90(5): 1078-1089.

Teitz, C. C., W. E. Garrett, Jr., et al. (1997). "Tendon problems in athletic individuals." Instr Course Lect 46: 569-582.

Thermann, H., O. Frerichs, et al. (2001). "Healing of the Achilles tendon: an experimental study." Foot Ankle Int 22(6): 478-483.

Thomopoulos, S., R. Das, et al. (2010). "bFGF and PDGF-BB for tendon repair: controlled release and biologic activity by tendon fibroblasts in vitro." Ann Biomed Eng 38(2): 225-234.

Thomopoulos, S., F. L. Harwood, et al. (2005). "Effect of several growth factors on canine flexor tendon fibroblast proliferation and collagen synthesis in vitro." J Hand Surg Am 30(3): 441-447.

Thompson, M. S. (2013). "Tendon mechanobiology: experimental models require mathematical underpinning." Bull Math Biol 75(8): 1238-1254.

Thornton, G. M., X. Shao, et al. (2008). "Changes in mechanical loading lead to tendonspecific alterations in MMP and TIMP expression: influence of stress

222

deprivation and intermittent cyclic hydrostatic compression on rat supraspinatus and Achilles tendons." Br J Sports Med 44(10): 698-703.

Tillander, B., L. E. Franzen, et al. (1999). "Effect of steroid injections on the rotator cuff: an experimental study in rats." J Shoulder Elbow Surg 8(3): 271-274.

Tkocz, C. and K. Kuhn (1969). "The formation of triple-helical collagen molecules from alpha-1 or alpha-2 polypeptide chains." Eur J Biochem 7(4): 454-462.

Tomita, F., K. Yasuda, et al. (2001). "Comparisons of intraosseous graft healing between the doubled flexor tendon graft and the bone-patellar tendon-bone graft in anterior cruciate ligament reconstruction." Arthroscopy 17(5): 461-476.

Trudel, G., Y. Koike, et al. (2007). "Mechanical alterations of rabbit Achilles' tendon after immobilization correlate with bone mineral density but not with magnetic resonance or ultrasound imaging." Arch Phys Med Rehabil 88(12): 1720-1726.

Tsuzaki, M., D. Bynum, et al. (2003). "ATP modulates load-inducible IL-1beta, COX 2, and MMP-3 gene expression in human tendon cells." J Cell Biochem 89(3): 556-562.

Uccelli, A., L. Moretta, et al. (2006). "Immunoregulatory function of mesenchymal stem cells." Eur J Immunol 36(10): 2566-2573.

Uemura, T., J. Dong, et al. (2003). "Transplantation of cultured bone cells using combinations of scaffolds and culture techniques." Biomaterials 24(13): 2277-2286.

Uysal, A. C. and H. Mizuno (2010). "Tendon regeneration and repair with adipose derived stem cells." Curr Stem Cell Res Ther 5(2): 161-167.

Uysal, A. C. and H. Mizuno (2011). "Differentiation of adipose-derived stem cells for tendon repair." Methods Mol Biol 702: 443-451.

Uysal, C. A., M. Tobita, et al. (2012). "Adipose-derived stem cells enhance primary tendon repair: biomechanical and immunohistochemical evaluation." J Plast Reconstr Aesthet Surg 65(12): 1712-1719.

Varseveld, R. B. v. and G. M. Bone (1997). "Accurate Position Control of a Pneumatic Actuator Using On/Off Solenoid Valves." TRANSACTIONS ON MECHATRONICS 2: 195-204.

Vogel, K. G. (2004). "What happens when tendons bend and twist? Proteoglycans." J Musculoskelet Neuronal Interact 4(2): 202-203.

Wang, A., W. Breidahl, et al. (2013). "Autologous Tenocyte Injection for the Treatment of Severe, Chronic Resistant Lateral Epicondylitis: A Pilot Study." Am J Sports Med.

223

Wang, J. H. (2006). "Mechanobiology of tendon." J Biomech 39(9): 1563-1582.

Wang, J. H., F. Jia, et al. (2003). "Cyclic mechanical stretching of human tendon fibroblasts increases the production of prostaglandin E2 and levels of cyclooxygenase expression: a novel in vitro model study." Connect Tissue Res 44(3-4): 128-133.

Wang, L., K. Seshareddy, et al. (2009). "Effect of initial seeding density on human umbilical cord mesenchymal stromal cells for fibrocartilage tissue engineering." Tissue Eng Part A 15(5): 1009-1017.

Wang, T., B. S. Gardiner, et al. (2013). "Bioreactor design for tendon/ligament engineering." Tissue Eng Part B Rev 19(2): 133-146.

Wang, T., B. Gardner, et al. (2012). "Bioreactor design for tendon/ligament engineering." Tissue Eng Part B Rev.

Wang, T., Z. Lin, et al. (2013). "Programmable mechanical stimulation influences tendon homeostasis in a bioreactor system." Biotechnol Bioeng 110(5): 1495-1507.

Webb, K., R. W. Hitchcock, et al. (2006). "Cyclic strain increases fibroblast proliferation, matrix accumulation, and elastic modulus of fibroblast-seeded polyurethane constructs." J Biomech 39(6): 1136-1144.

Weiler, A., F. N. Unterhauser, et al. (2002). "Alpha-smooth muscle actin is expressed by fibroblastic cells of the ovine anterior cruciate ligament and its free tendon graft during remodeling." J Orthop Res 20(2): 310-317.

West, J. R., N. Juncosa, et al. (2004). "Characterization of in vivo Achilles tendon forces in rabbits during treadmill locomotion at varying speeds and inclinations." J Biomech 37(11): 1647-1653.

Wiggins, M. E., P. D. Fadale, et al. (1995). "Effects of local injection of corticosteroids on the healing of ligaments. A follow-up report." J Bone Joint Surg Am 77(11): 1682-1691.

Williams, J. G. (1993). "Achilles tendon lesions in sport." Sports Med 16(3): 216-220.

Wilson, A. M. and A. E. Goodship (1994). "Exercise-induced hyperthermia as a possible mechanism for tendon degeneration." J Biomech 27(7): 899-905.

Wojciak, B. and J. F. Crossan (1994). "The effects of T cells and their products on in vitro healing of epitenon cell microwounds." Immunology 83(1): 93-98.

Woodley, B. L., R. J. Newsham-West, et al. (2007). "Chronic tendinopathy: effectiveness of eccentric exercise." Br J Sports Med 41(4): 188-198; discussion 199.

224

Woon, C. Y., S. Farnebo, et al. (2012). "Human Flexor Tendon Tissue Engineering: Revitalization of Biostatic Allograft Scaffolds." Tissue Eng Part A.

Woon, C. Y., A. Kraus, et al. (2011). "Three-dimensional-construct bioreactor conditioning in human tendon tissue engineering." Tissue Eng Part A 17(19-20): 2561-2572.

Woon, C. Y., A. Kraus, et al. (2011). "Three-Dimensional-Construct Bioreactor Conditioning in Human Tendon Tissue Engineering." Tissue Eng Part A.

Wren, T. A., D. P. Lindsey, et al. (2003). "Effects of creep and cyclic loading on the mechanical properties and failure of human Achilles tendons." Ann Biomed Eng 31(6): 710-717.

Wren, T. A., S. A. Yerby, et al. (2001). "Mechanical properties of the human achilles tendon." Clin Biomech (Bristol, Avon) 16(3): 245-251.

Yang, G., H. J. Im, et al. (2005). "Repetitive mechanical stretching modulates IL-1beta induced COX-2, MMP-1 expression, and PGE2 production in human patellar tendon fibroblasts." Gene 363: 166-172.

Ye, S., P. Eriksson, et al. (1996). "Progression of coronary atherosclerosis is associated with a common genetic variant of the human stromelysin-1 promoter which results in reduced gene expression." J Biol Chem 271(22): 13055-13060.

Yuan, J., M. X. Wang, et al. (2003). "Cell death and tendinopathy." Clin Sports Med 22(4): 693-701.

Zamara, E., S. Galastri, et al. (2007). "Prevention of severe toxic liver injury and oxidative stress in MCP-1-deficient mice." J Hepatol 46(2): 230-238.

Zell, R. A. and V. M. Santoro (2000). "Augmented repair of acute Achilles tendon ruptures." Foot Ankle Int 21(6): 469-474.

Zhang, F., H. Liu, et al. (2003). "Effect of vascular endothelial growth factor on rat Achilles tendon healing." Plast Reconstr Surg 112(6): 1613-1619.

Zhang, L., B. S. Gardiner, et al. (2007). "The effect of cyclic deformation and solute binding on solute transport in cartilage." Arch Biochem Biophys 457(1): 47-56.

Zhang, L., B. S. Gardiner, et al. (2008). "A fully coupled poroelastic reactive-transport model of cartilage." Mol Cell Biomech 5(2): 133-153.

Zheng, M. H., J. Chen, et al. (2005). "Porcine small intestine submucosa (SIS) is not an acellular collagenous matrix and contains porcine DNA: possible implications in human implantation." J Biomed Mater Res B Appl Biomater 73(1): 61-67.

225

APPENDIX

Bioreactor Design for Tendon/Ligament Engineering

Tao Wang, BS,1 Bruce S. Gardiner, PhD,2 Zhen Lin, PhD,1,3 Jonas Rubenson, PhD,4

Thomas B. Kirk, PhD,5 Allan Wang, PhD,6 Jiake Xu, PhD,7 David W. Smith, PhD,2

David G. Lloyd, PhD,4,8 and Ming H. Zheng, MD, PhD, FRCPath, FRCPA1

Tendon and ligament injury is a worldwide health problem, but the treatment options remain limited. Tendonand ligament engineering might provide an alternative tissue source for the surgical replacement of injuredtendon. A bioreactor provides a controllable environment enabling the systematic study of specific biological,biochemical, and biomechanical requirements to design and manufacture engineered tendon/ligament tissue.Furthermore, the tendon/ligament bioreactor system can provide a suitable culture environment, which mimicsthe dynamics of the in vivo environment for tendon/ligament maturation. For clinical settings, bioreactors alsohave the advantages of less-contamination risk, high reproducibility of cell propagation by minimizing manualoperation, and a consistent end product. In this review, we identify the key components, design preferences, andcriteria that are required for the development of an ideal bioreactor for engineering tendons and ligaments.

Introduction

Tendons and ligaments have an important function intransferring force from muscle to bone or bone to bone.

Tendons also help store elastic energy while walking, in-creasing locomotion efficiency. However, sudden, excessivestrain of ligaments and tendons from athletic or recreationalactivities can cause acute traumatic injury to these tissues,which can range from small, partial tears to complete rup-tures. Furthermore, repetitive loadings over long periods oftime can lead to similar tissue injuries, when the fatiguedamage within the tendon exceeds the capacity of the tissueto repair itself. In the United States, injuries to tendons andligaments represent about half of the 33 million musculo-skeletal injuries.1 Each year more than 33,000 tendon recon-structions occur in the United States, costing $30 billionUSD2,3 and in Australia, $250 million is spent annually juston rotator cuff repair.4 However, despite the high prevalenceof tendon injury and associated tendinopathy worldwide,treatment options remain poorly defined.

Autograft and allograft transplantations are the commonsurgical treatments for tendon and ligaments that are injured ordegenerated. However, the risks of damage to the donor sitefrom which the autografts are taken, and the potential immune

reactions for allografts are the major concerns.5–7 A promisingtranslational approach to the treatment of tendon/ligamentinjury or degeneration is through the use of engineered autol-ogous grafts made available through the development of bio-reactors that generate tendon/ligament tissue in vitro. Onecommon view is that the key to a successful bioreactor is beingable to recreate, in vitro, the cell microenvironments that areexperienced by cells in vivo. The cell microenvironments can bedefined using cellular morphological information with datafrom molecular biology, biochemistry, and biomechanics. Thisreview aims to clarify the requirements for a successful biore-actor that may be used for tendon/ligament engineering, andto provide an overview of the range of components found intendon/ligament bioreactors, including custom-made andcommercial products. We will also discuss the studies that haveinvolved the application of tendon/ligament bioreactors.

Key Elements for Tendon/Ligament Formationand Regeneration

Tendons and ligaments are force-transferring tissues frommuscle to bone and bone to bone, respectively. They consistof collagens, cells, proteoglycans, elastin, glycolipids, andwater. Although roughly 65%–70% of the total weight is

1Centre for Orthopaedic Translational Research, School of Surgery, University of Western Australia, Crawley, Australia.2School of Computer Science and Software Engineering, University of Western Australia, Crawley, Australia.3Division of Orthopaedic Surgery, Department of Surgery, Guangdong General Hospital, Guangdong Academy of Medical Sciences,

Guangzhou, Guangdong, China.4School of Sport Science, Exercise and Health, University of Western Australia, Crawley, Australia.5Department of Mechanical Engineering, Curtin University, Bentley, Australia.6Sir Charles Gairdner Hospital, Perth, Australia.7School of Pathology and Laboratory Medicine, University of Western Australia, Crawley, Australia.8Centre for Musculoskeletal Research, Griffith Health Institute, Griffith University, Gold Coast, Australia.

TISSUE ENGINEERING: Part BVolume 19, Number 2, 2013ª Mary Ann Liebert, Inc.DOI: 10.1089/ten.teb.2012.0295

133

226

water,8 tendon is a highly organized structure. Collagen typeI is the main structural/functional component and comprisesaround 70%–80% of the dry weight. Type III collagen ismainly present in the endotenon and epitenon, but is alsopresent in the early phase of tendon repair.9,10 Tendons andligaments are relatively hypovascular and hypocellular tis-sues, with their cells (tenocytes and fibroblasts) comprising< 5% of the total volume. The morphology of tenocytes andfibroblasts is sharp and usually elongated along collagen fi-bers in normal tissues, but some rounded tenoblasts are foundoccasionally.11–14 Based on the composition and function oftendon/ligament tissue, one must consider the four basic el-ements for their successful regeneration: the cell source, thecharacteristics of the scaffold matrix, and establishing an ap-propriate chemical and physical cellular microenvironment.

An ideal cell source should meet the following three re-quirements: availability, rapid proliferation, and the abilityto differentiate into in situ cells.15 Stem cells, dermal fibro-blasts, and tenocytes have all been tested as potential cellsources for tendon repair. In the past decade, promising re-sults have been achieved using mesenchymal stem cells(MSCs) in tendon/ligament engineering.16–23 However, thepotential for MSCs to also differentiate into osteoblastic cellsleading to ectopic bone formation raises concerns for theirapplication in tendon repair. Likewise, bone marrow-derivedMSCs that have been used in various studies24–28 also pose apotential risk of ectopic bone formation.29 A recent studydemonstrated that embryonic stem cells (ESCs) can be in-duced to exhibit a tenocyte-like morphology and expresstendon-related gene markers if exposed to an appropriatebiomechanical environment.2 Unfortunately, the clinical ap-plication of ESCs is restricted due to their limited availability(depending on the policy of different countries) and thecomplexity of cell manipulation required. Dermal fibroblastshave also been shown to form tendon tissue,30–32 althoughfurther research has suggested that the healing process usingskin-derived fibroblasts is suppressed with a lack of tenocytemarkers and histopathologic correlations.33,34 Being nativecell sources, tenocytes and in situ fibroblasts are perhaps themost ideal cell sources for engineered tendon and ligamenttissue, respectively. Preclinical and early clinical studies us-ing these native cell sources are promising.35–42 In the mostrecent clinical trial of autologous tenocyte therapy, a total of25 patients with recalcitrant lateral epicondylitis were treatedwith fibroblasts isolated from biopsied anterior cruciate lig-ament tissue using a patented technique developed by ourlaboratory. We found a 60% or greater improvement in boththe Quick DASH and pain assessment scores.43,44

Apart from the selection of the cell source, cell seeding alsoplays an essential role in the development of engineeredtendon and ligament. Several reports indicate that sufficientcell number and a uniform distribution throughout thescaffold are desirable for achieving a homogeneous extra-cellular matrix (ECM) deposition in vitro.45,46 Compared tolower initial cell-seeding density, high seeding density hasbeen shown to result in increasing ECM deposition rate, ahigher final cell number and better cellular morphology.47,48

However, overseeding also has potential for negative effectson nutrient delivery, and consequently cellular metabolismand cell viability. Nutrient depletion at high cell-seedingdensities could lead to spatially in homogenous ECM pro-duction.49 A study by Issa et al. showed that mechan-

ostimulated human umbilical veins seeded with 3 millioncells/mL had better cellular proliferation rates than othergroups.28 However, optimal seeding density will vary de-pending on the cell source and the bioscaffold physical andchemical properties, as these properties will affect nutrienttransport and rates of consumption, as well as the mechan-ical and chemical stimuli provided to the attached cells.49

The construct scaffold plays on important role in engi-neering the new tendon/ligament tissue. Ideally, the scaffoldshould be able to provide substantial initial mechanicalstrength for its immediate postimplantation functional role,while providing a suitable biological environment for cellmigration and proliferation. Furthermore, the degradationrate of the biomaterial needs to be comparable with the rateof tissue synthesis, to allow the eventual replacement of thestarting scaffold with neotissue.

Both synthetic and natural biomaterials are commonlyused in tendon/ligament engineering. Synthetic polymerscaffolds have the advantage of reproducible mechanical andchemical properties, and they are relatively easy to fabricateinto different sizes.24,25,40,41,50 However, their rapid degra-dation rate and potential risk of releasing acidic byproductsor toxic polyesters during degradation have limited theirapplication in clinical trials. Given these disadvantages, moreresearchers have turned their focus on exploring naturalbiomaterials.19–21,37,51–54 Being the main component of nativetendon, collagen type I is the most obvious choice of mate-rial. Although the biocompatibility is excellent, the poormechanical properties of reconstituted type I collagen scaf-folds has limited their further development as a load-bearingmaterial. Silk fibroin, on the other hand, has similar bio-compatibility as collagen scaffolds and comparable me-chanical properties as native tendon/ligament. In severalin vivo studies, silk fibroin-based engineered ligaments havebeen proven their ability to restore the function of injuredligament.16,17,26 Another option is the decellularised tendon/ligament construct. Although the mechanical and biologicalproperties are a better match to native tissue than any othercurrently available scaffolds, donor cells may remain in theallograft, even with strict sterilization and cleaning, and thusthey can potentially cause inflammatory responses.55,56

After choosing an appropriate cell source and scaffold type,the tissue needs to be encouraged to develop the properties ofnative tissue by providing an appropriate biochemical andbiomechanical environment to stimulate ECM synthesis. Re-garding the biochemical environment, several growth factorshave been found to play an important role in tendon/ligamentformation and healing. These include Insulin-like growth fac-tor-I, vascular endothelial growth factor, platelet-derivedgrowth factor, basic fibroblast growth factor, transforminggrowth factor (TGFb), and growth differentiation factor 5(GDF-5). The roles of TGFb and GDF-5 seem to be particularlyprominent. TGFb remains active throughout tendon/ligamenthealing,57 and is able to regulate cell migration, proteinaseexpression, fibronectin-binding interactions, cell proliferation,and collagen production.57–60 Recent studies demonstratedthat tendon and ligament formation was impaired in TGFb2and TGFb3 knockout mouse embryos, reinforce the importanceof TGFbs in tendon development and homeostasis.61 GDF-5regulates cell growth and differentiation, with a lack of GDF-5causing delayed tendon healing, irregular collagen type I fi-brils, and weakened fibril mechanical properties.62–64

134 WANG ET AL.

227

How these various biochemical factors should be intro-duced into the tissue bioreactor system to shape the chemicalenvironment is an extremely challenging and open question.Specifically, at what concentrations, in what combinations,and at what sequence or timing should they be made avail-able? Studies have shown that the expression of these factorsin tendon repair changes over periods of days.65–67 Pre-sumably, as the engineered tissue progresses, cells begin tocontrol their own biochemical environment, and the role of thebioreactor is to now provide the building block nutrients andexpected systemic signals, along with a mechanical stimulus.The early stage in the tissue-engineered tendon is likely to bethe most critical in establishing tendon development along apathway to resemble native tendon. Finding the correctcombination of factors is daunting due to the complexityarising from the multitude of possible combinations and in-teractions. Systematically, varying experimental conditions,coupled with computational modeling of transport processesand signaling molecule pathways leading to cell responses,provide the only conceivable means to both understandingand efficiently optimizing the tissue bioreactor system.

Tendon’s primary function is mechanical. It operates in avarying load environment, both on short timescales (e.g.,walking and running) and on longer timescales (e.g., changes inbody size with age). Tendon responds to its mechanical envi-ronment through changes in ECM biosynthesis and degrada-tion. Unsurprisingly, given its functional role, a suitablemechanical stimulus is vital for tendon/ligament homeostasis.In fact, it has been shown that after 4 weeks in a load-free cultureenvironment, tenocytes lose their native elongated morphology,become increasingly rounded, and the collagen fiber becomesmore crimped.68 In cell-free reconstructed collagen fibril net-work systems, a tensile load has been shown to be protective todegradation by MMP-8.69 Furthermore, cyclic stretching hasbeen shown to produce an up to nine-fold increase in the cellnumber of an engineered tendon compared with a static cultureover a 2-week period.23 Finally, an appropriate mechanical en-vironment could help guide collagen fiber formation, that is,along the direction of loading, which is able to enhance or op-timize the mechanical properties, including stiffness, elasticmodulus, maximum tensile stress, and maximum force.19,31,36

Rather than being two separate signals, there is a crosstalkbetween mechanical and chemical signals. Recent studieshave shown that gradual and temporary loss of tensileloading leads to reversible loss of Scleraxis (Scx) expression,which is a transcription factor specific for tenocytes and theirprogenitors. In addition, it has been shown that TGFb di-rectly induced the expression of Scx in cultured tenocytesisolated from mice.70 In Scx - / - mice, a disordered limbtendon phenotype was observed,71 and a similar phenome-non happened in TGFb type II receptor gene knockout(Tgfbr2 - / - ) mice with dramatic loss of Scx expression.61

Providing a suitable biomechanical signal is clearly animportant component for the success of tendon/ligamentengineering. Generating a suitable mechanical signal withinthe bioreactor system is critically important for tendon/lig-ament tissue engineering.

Bioreactor Design for Tendon/Ligament Engineering

Despite the increasing appreciation of tendon/ligamentbiology and function, conventional culture methods do not

seem to meet the biochemical and biomechanical require-ments to generate bioengineered tendon/ligament in vitro. Abioreactor system that subjects the cell culture to dynamicloading, mimicking the physiological conditions of tendon/ligament in vivo, while allowing cellular proliferation, dif-ferentiation, and matrix production in a mechanical envi-ronment, may provide a solution.

Bioreactors for tendon/ligament engineering are differentto the systems that have been used in various other tissue-engineering fields in the past decades, that is, systems for themuscle,72 liver,73 and bone.74,75 Compared to other bioreac-tors, the main task of the bioreactor for tendon/ligamentengineering is to provide the proper biomechanical andbiochemical environment specific to tendon/ligament for-mation. To achieve this, certain basic components are re-quired, that is, the actuating system and the culture chamber,which can provide the construct’s mechanical stimulationand controlled culture environment, respectively. Further-more, the bioreactor may also include a medium circulationsystem, monitoring system, feedback system, and a mediumanalysis system, depending on the operational requirements(Fig. 1). With these facts in mind, several custom-made bio-reactors have been developed for tendon/ligament engi-neering (see Table 1), and the aforementioned components ofthese are now are discussed in detail.

Actuator and Culture Chamber Design

The actuating system is the main component for providingdifferent mechanical stimulation to engineered tissue.Pneumatic actuators,19–21 linear motors,22,27 and step motor-ball screws (SMBSs)2,42,76 are the most common actuatorsused in tendon/ligament bioreactors. Pneumatic actuatorshave several advantages, including the ease of maintenance,cleanliness, low cost, and high power-to-weight ratio.77 Un-fortunately, by using air as a medium, pneumatic actuatorsare subject to high friction. The sensitivity and response to aninput signal are relatively slow because of the dead band anddead time caused by stiction and air compressibility. Due tothese nonlinearities, it is difficult to achieve accurate positioncontrol with pneumatic actuators.77 Typically, the accuracyof pneumatic actuators is *– 0.1 mm, which is not insignif-icant when typical tendon bioreactors require < 5% strain on1–5-cm tissues. Compared with pneumatic actuators, elec-trical actuators are more expensive, but the level of accuracyin their positional control is much higher. A direct-drivelinear motor is able to provide high-speed/high-accuracylinear motion by eliminating mechanical transmission,78–80

and the accuracy is the highest of all three actuators, that is,*– 1mm. SMBS transmission systems are based on a ball-screw linkage and a crank-slider mechanism. In general,SMBSs transfer the rotation of the crank to a reciprocatingmotion of the screw.81 By choosing different ball screws, theoptimal speed range and output force can be selected usingEquation (1), that is,

T¼ Fl

2p�(1)

where T is torque applied to screw; F is linear force; l is ballscrew lead; and n is ball screw efficiency. Although SMBS isnot as accurate as a linear motor, a positional accuracy ofaround – 5mm can still be achieved. In a multichamber-shared

BIOREACTOR DESIGN FOR TENDON/LIGAMENT ENGINEERING 135

228

loading system, SMBS is widely used given its high accuracyand relative high loading capacity.35,42,76 In independentloading multichamber systems, linear motors are morepopular due to their small size and relatively simple me-chanical arrangement.22,82

In the addition to the actuator, the connection between themechanical input and the tissue is vital and often presents amajor challenge. During dynamic loading of the tissue inculture, an even distribution of force throughout the entiresample is critical, otherwise tissue integrity is compromisedby overloading the mechanical connection regions and/or byinhomogeneous mechanical stimulation. Different strategiesof applying mechanical loads have been adopted based ondifferent construct dimensions. In a study by Chen et al., cell-seeded knitted silk–collagen sponge scaffolds were fixed onstainless rings and connected to sample hooks.2 The biore-actor system by Juncosa-Melvin et al. applied two posts to fixthe construct, punching through the scaffold as shownin Figure 2.19 However, nonuniform construct deformationis clearly apparent. Tensile force is focused on the side ofconstructs, which causes uneven distribution of mechanical

stimulation. Although tissue clamps are the most popularand relatively effective method for holding tendon con-tructs,22,42,82 the potential for damage at the clamping regionneeds to be considered. The method of reproducibly ap-plying uniform loads to soft tissue without tissue damageor slippage is a critical problem in need of a satisfactorysolution. It is our opinion that a robust clamping regionshould be designed along with the artificial bioscaffold toensure the proper connection between the sample and themechanical load.

The culture chamber is an essential part of whole biore-actor system. The high humidity of culturing conditions(99% humidity, 37�C, and 5% CO2) and chemistry of theculture medium are corrosive to many materials. Corrosionproducts may in turn be toxic to the tissue. Therefore,noncorrosive and autoclavable materials such as stainlesssteel, polymethylemethacrylate, polyoxymethylene, poly-carbonate, glass, and silicon are preferred and widely used inculture chamber design.2,23,27,76,83

The chamber structure is an important consideration in thebioreactor design, with most currently available culture

FIG. 1. Schematicdemonstration of connectionbetween differentcomponents of tendon/ligament bioreactor system.Color images available onlineat www.liebertpub.com/teb

136 WANG ET AL.

229

chambers divided into two groups: integrated42,83 and sep-arated chambers.2,3,19–27,35,76,82 In integrated chambers, mul-tiple samples are cultured while sharing the same culturemedium. Conversely, separated chambers can provide sep-arate culture environments for each sample. Although thecomplexity of design and manufacturing costs may be higherin a separated chamber system, reduced cross-contaminationand the option of independent environmental control aredistinct advantages.

During tissue culture, sufficient air exchange within theculture chamber is critical. Air exchange in conventional cellculture incubators is through the integrated hydrophobicfilter of the culture flask and the gap between the leak andthe culture dish/well plate. Therefore, the ideal design forthe culture chamber should be similar. Like conventional cellculture, an unsealed chamber bioreactor connects to theoutside environment through various ways,19,22,27,35 such asthe hydrophobic filter leak42 and the labyrinth channel (Fig.3).82 In a study by Webb et al.,42 a modified tissue cultureflask was used as an integrated culture chamber, and could

culture up to four samples simultaneously. Although thereare potential risks of cross-contamination from differentsamples and toxicity from autoclaving the culture flask, theintegrated hydrophobic filter leak can ensure adequate gasexchange without inducing contamination. In the bioreactorsystem used in Parent et al.’s research,82 a labyrinth channelwas added to improve the air exchange and eliminate con-tamination as shown as Figure 3. However, in closedchambers, air exchange mostly depends on medium circu-lation, which is now discussed in the following section.

Environmental Control and MediumCirculation Systems

Although to the best of our knowledge, no contaminationhas been reported in any bioreactor study, transportation ofthe bioreactor and opening of chamber for medium exchangeevery 3 days, especially for multichamber systems, are still apotential contamination risk. Therefore, a medium circula-tion system can be introduced to improve the efficiency andminimize these risks. In addition to the advantages of re-duced contamination, a circulating medium may be betterable to infiltrate into cultured tissue. For example, perfusionbioreactors used in bone engineering circulate the culturemedium for better nutrient delivery and subsequent im-proved cell numbers.84–89 Similarly, human umbilical veincultured under a circulating medium had approximatelythree times the cellular number as those cultured using aquiescent medium.23

FIG. 2. Stem cell-seeded Collagen sponge deformationduring mechanical stimulation. Modified from reference 19.Color images available online at www.liebertpub.com/teb

Table 1. Components of Custom-Made Bioreactor Systems

ReferenceActuating

system Culture chamber Monitor systemFeedbacksystem

Mediumcirculation system

Mediumanalysis system

76 Biaxial Multiple · · X ·42 Uniaxial Single chamber, multiple

samples· · · ·

19,20,21,3 Uniaxial Multiple Displacement · · ·35 Uniaxial Multiple Force · · ·22 Uniaxial Single Force · · ·23 Uniaxial Multiple · · X ·83 Uniaxial Single chamber, multiple

samplesDisplacement Force · · X

2 Uniaxial Multiple · · · ·27 Uniaxial Multiple Displacement · · ·82 Uniaxial Multiple Displacement Force X · ·

FIG. 3. Schematic drawings of air exchange through thelabyrinth channel in the culture chamber. Modified fromreference 82. Color images available online at www.liebertpub.com/teb

BIOREACTOR DESIGN FOR TENDON/LIGAMENT ENGINEERING 137

230

A traditional incubator-based bioreactor and/or inde-pendent bioreactor can be used for tendon/ligament engi-neering. In an incubator bioreactor system, air exchange isthrough the hydrophobic filter leak of the medium reservoir,and proper CO2 and temperature level is controlled by theincubator. Then, by circulating the culture medium, suitableconditions can be applied to engineered tendon/ligament, asshown in Figure 4.23 However, an independent bioreactorsystem, as shown Figure 5, does not rely on an incubator tocontrol the culture environment. The percentage of differentgases (PO2, CO2, and N2) are controlled by air valves,76 andthe mixed gas is humidified first, and then passed into amedium heater. Waste gas emission is transported through afilter in the case of contamination. The prepared, warm cul-ture medium is circulated through the bioreactor culturechamber. For certain tissues such as cartilage, some specificculture conditions are required. For example, under a low-oxygen environment, engineered cartilage displays fastermatrix glycosaminoglycan deposition rate and better cellularmorphology, but with less dedifferentiation.90–92 The envi-ronmental chamber allows researchers to manipulate differ-ent culture conditions, thereby enabling a systematic studyof cell growth and differentiation into functional tissue.

Monitoring and Feedback Systems

The biomechanical properties of the final engineered ten-don/ligament should be a central concern of bioreactor de-sign, as the tendon/ligament’s functional role in the body isprimarily mechanical. The maximum load and elastic mod-

ulus are essential mechanical properties to evaluate thesuitability of engineered tendon/ligament. However, in moststudies, mechanical tests are performed only at the end oftissue culture. For example, the stiffness of the constructs atdifferent time points during culture is rarely recorded orassessed. The correlation between stiffness and tissue mat-uration may provide a better understanding about how thecell differentiates into functional tissue, and for this reason,online stiffness monitoring is likely to be invaluable.

In tendon/ligament bioreactors, force measurement isenabled using sensors called load cells,22,82,93 which are lo-cated differently in various bioreactor designs. Load cellslocated between the actuator and sample clamps,83 shown inFigure 6A, requires high manufacturing accuracy, as thefriction between the shaft and culture chamber can induceerror. For fragile materials such as biological tissues, thisfriction might be higher than the applied load. Conversely, aload cell placed at the end of the culture chamber, shown inFigure 6B, can minimize the fiction between the actuator andculture chamber and also the fluid resistance during stimu-lation. To acquire accurate data, load cell selection should bebased on the initial mechanical properties of the constructs.Working with the sensitivity of the load cell, extra attentionis needed when manipulating the construct so as to avoidcausing damage to the load cell through overloading.

In addition to force monitoring, tissue displacement isanother important variable that requires monitoring. Linearvariable differential transformers and optical decoders arecommonly used position sensors.19,82 Although the motionof actuating systems is preprogrammed, overload of theactuator and manual misoperation can cause desynchroni-zation between the program and actual stimulation. The real-time displacement monitor can produce a full record ofstimulation position, which then allows researchers to trackif there are any unusual features in the results. Anotherfunction of the position sensor is to provide feedback of thedisplacement information to the actuator control system tocorrect for any desynchronization.82 Critically, by monitor-ing load and displacement, real-time stiffness during en-gineered tendon/ligament culture can be estimated.

Lastly, imaging the construct might be a potentially im-portant monitoring method that can be adopted in bioreactorsystems. Two imaging modalities, the confocal microscopeand optical coherence tomography, are potential candi-dates.94,95 However, sample deformation induced by me-chanical stimulation can cause image shift resulting in poorfocus or loss of image. Imaging dynamic sample is extremelydifficult in conventional bioreactors, as features of interest(e.g., cells) soon leave the field of view. Techniques need tobe developed to move the sample and microscope together.This can become difficult, especially, when doing in situimaging of a tendon tissue culture.

Commercial Bioreactor Systems for Tendon/Ligament Engineering

Recently, commercial tendon-/ligament-engineering bio-reactor systems have become available. As discussed above,these basic principles are applied to customized bioreactors;however, with greater design input and advancedmanufacturing techniques, commercial products are ableto provide more accurate and complex environments for

FIG. 4. Schematic illustration of an incubator bioreactorsystem. Suitable temperature, humidity, and CO2 level of theculture medium are maintained by the incubator in the me-dium reservoir. The medium is circulated by a pump. Thewaste valve is closed normally, and it will open duringmedium exchange. Color images available online at www.liebertpub.com/teb

138 WANG ET AL.

231

FIG. 5. Schematic diagram of an independent bioreactor system. Suitable temperature, humidity, and CO2 level of theculture medium are controlled by the environmental chamber. Cold culture medium is pumped to the environmentalchamber for heating and then circulated through the culture chamber. The waste valve is closed normally, and it will openduring medium exchange. Color images available online at www.liebertpub.com/teb

FIG. 6. Load cell position indifferent bioreactor systems. (A)Load cell is located between theactuator and sample clamps. (B) Aseparate load cell placed at the endof culture chamber system. Colorimages available online at www.liebertpub.com/teb

BIOREACTOR DESIGN FOR TENDON/LIGAMENT ENGINEERING 139

232

Ta

ble

2.S

tu

dies

Usin

gB

io

rea

cto

rS

ystem

sfo

rT

en

do

n/

Lig

am

en

tE

ng

in

eer

in

g

Fir

stau

thor

Bio

reac

tor

type

(Com

pan

y)

Par

amet

ers

ofm

echan

ical

stim

ula

tion

Sca

ffol

dm

ater

ial

(dim

ensi

ons)

Cel

lso

urc

eE

ffec

tR

efer

ence

Alt

man

(200

2)C

ust

om

-mad

est

epm

oto

rb

iore

acto

rw

ith

env

iro

nm

enta

lch

amb

erC

ycl

icst

retc

hin

g2-

mm

,90

o

rota

tio

n,

0.01

67H

z,21

day

sC

oll

agen

typ

eI

gel

s,B

omby

xm

ori

silk

wo

rmsi

lkfi

ber

mat

rice

s(l

eng

th:3

0m

m)

Hu

man

bo

ne

mar

row

stro

mal

cell

s(h

BM

SC

)

Elo

ng

atio

no

fh

BM

SC

s,cr

oss

-sec

tio

nce

lld

ensi

ty[

76

Jun

cosa

-Mel

vin

(200

6)C

ust

om

-mad

ep

neu

mat

iccy

lin

der

bio

reac

tor

wit

hL

VD

Tfo

rd

isp

lace

men

tm

on

ito

rin

g

Cy

clic

stre

tch

ing

2.4%

stra

in,

0.00

33H

z,8

h/

day

for

2w

eek

sT

yp

eI

coll

agen

spo

ng

e(2

3–

0.8

mm

·9

–0.

8m

3–

0.1

mm

)

Rab

bit

MS

Cs

Max

imu

mfo

rce[

,li

nea

rst

iffn

ess[

,m

axim

um

stre

ss[

,li

nea

rm

od

ulu

s[

19

Web

b(2

006)

Cu

sto

m-m

ade

step

mo

tor

bio

reac

tor

Cy

clic

stre

tch

ing

10%

stra

in,

0.25

Hz,

8h

/d

ayfo

r7

day

sP

oly

ure

than

eco

nst

ruct

(20

·10

·2

mm

)H

um

antr

ach

eal

fib

rob

last

sT

yp

eI

coll

agen

[,

TG

Fb-

1[,

CT

GF

,E

last

in[

,al

ph

aI[

,P

roco

llag

en[

,F

ibro

nec

tin

[,

MM

P-1

[,

elas

tic

mo

du

lus[

42

An

dro

jna

(200

7)C

ust

om

-mad

eb

iore

acto

rw

ith

load

-d

isp

lace

men

tm

easu

resy

stem

Cy

clic

stre

tch

ing

9%st

rain

,tw

ice

dai

lyfo

r30

min

each

per

iod

sep

arat

edb

y8-

hre

stfo

r2

wee

ks

Sm

all-

inte

stin

esu

bm

uco

sa(3

·5

·95

mm)

Do

gte

no

cyte

sC

ell

den

sity

[,

stif

fnes

s[35

Nir

mal

anan

dh

an(2

008)

Cu

sto

m-m

ade

pn

eum

atic

cyli

nd

erb

iore

acto

rC

ycl

icst

retc

hin

g2.

4%an

d1.

2%st

rain

,1

Hz,

stim

ula

tio

np

erio

d:

8h

/d

ay,

100

and

3000

cycl

es/

day

for

12d

ays

Ty

pe

Ico

llag

ensp

on

ge

Rab

bit

ilia

c-cr

est

MS

Cs

Th

est

imu

lati

on

pat

tern

of

2.4%

stra

in,

3000

cycl

es/

day

,1

Hz

has

the

bes

tef

fect

on

incr

easi

ng

stif

fnes

s.

21

Nir

mal

anan

dh

an(2

008)

Cu

sto

m-m

ade

pn

eum

atic

cyli

nd

erb

iore

acto

rC

ycl

icst

retc

hin

g2.

4%st

rain

,0.

0033

Hz,

8h

/d

ayfo

r12

day

sT

yp

eI

pu

rifi

edb

ov

ine

coll

agen

gel

,ty

pe

Ico

llag

ensp

on

ges

(len

gth

:11

and

51m

m)

Rab

bit

MS

Cs

Sti

ffn

ess[

20

Ng

uy

en(2

009)

Cu

sto

m-m

ade

lin

ear

actu

ato

rb

iore

acto

rw

ith

load

cell

for

forc

em

easu

rem

ent

Pre

load

edw

ith

0.05

N,

cycl

icst

retc

hin

g10

%,

0.5

Hz,

2h

stim

ula

tio

n—

2h

rest

—2

hst

imu

lati

on

—18

hre

stfo

r5

day

s

Po

rcin

esm

all

inte

stin

esu

bm

uco

sa(l

eng

th:

2cm

and

wid

th:

1cm

)

Rab

bit

MC

Lfi

bro

bla

sts

Imp

rov

edfi

ber

ori

enta

tio

n,

fib

eran

gu

lar

dis

per

sio

nY

,b

ette

ro

rgan

ized

coll

agen

fib

er,

elo

ng

ated

cell

mo

rph

olo

gy

,

22

Ab

ou

slei

man

(200

9)C

ust

om

-mad

eli

nea

rac

tuat

or

bio

reac

tor

wit

hm

ediu

mci

rcu

lati

on

syst

em

Cy

clic

stre

tch

ing

2%st

rain

,1

h/

day

,0.

0167

Hz

for

1an

d2

wee

ks.

Hu

man

um

bil

ical

vei

ns

(wal

lth

ick

nes

s:0.

75m

m,

ou

ter

dia

met

er:

6.75

–0.

25m

m,

len

gth

:8.

5cm

)

Wis

tar

Rat

bo

ne

mar

row

MS

Cs

bet

ter

cell

dis

trib

uti

on

,m

ore

elo

ng

ated

cell

mo

rph

olo

gy

,ce

llp

roli

fera

tio

n[

,u

ltim

ate

stre

ss[

,el

asti

cm

od

ulu

s[

23

Bu

tler

2009

)C

ust

om

-mad

ep

neu

mat

iccy

lin

der

bio

reac

tor

wit

hL

VD

Tfo

rd

isp

lace

men

tm

on

ito

rin

g

Cy

clic

stre

tch

ing

2.4%

stra

in,

0.00

33H

z,8

h/

day

for

2w

eek

sT

yp

eI

coll

agen

spo

ng

e(9

4%p

ore

vo

lum

e;62

-mm

mea

np

ore

dia

met

er)

Mo

use

mes

ench

ym

alst

emce

llT

yp

eI

coll

agen

[,

lin

ear

stif

fnes

s[83

(con

tin

ued

)

140

233

Ta

bl

e2.

(Co

nt

in

ue

d)

Fir

stau

thor

Bio

reac

tor

type

(Com

pan

y)

Par

amet

ers

ofm

echan

ical

stim

ula

tion

Sca

ffol

dm

ater

ial

(dim

ensi

ons)

Cel

lso

urc

eE

ffec

tR

efer

ence

Ch

en(2

010)

Cu

sto

m-m

ade

step

mo

tor

bio

reac

tor

Cy

clic

stre

tch

ing

10%

stra

in,

2h

/d

ay,

1H

zfo

r14

day

sK

nit

ted

silk

-co

llag

ensp

on

ge

scaf

fold

(5·

0.5

·0.

2cm

)H

um

anem

bry

on

icst

emce

llC

oll

agen

I[,

Co

llag

enII

I[,

Ep

ha4

[,

Scx

[,

So

x9Y

,M

yo

sin

[,

Inte

gri

na1

[,

Inte

gri

na2

[,

Inte

gri

nb1

[,

Co

llag

enco

nte

nt[

,C

oll

agen

dia

met

er[

,b

ette

rce

llal

ign

men

t

2

Do

rosk

i(2

010)

Cu

sto

m-m

ade

lin

ear

mo

tor

bio

reac

tor

Cy

clic

stre

tch

ing

10%

stra

in(5

%o

ffse

t,5%

amp

litu

de)

,1

Hz,

3h

/d

ay,

1.7,

14,

and

21d

ays

Po

ly(e

thy

len

eg

lyco

l)-b

ased

hy

dro

gel

mat

eria

lo

lig

o(p

oly

(eth

yle

ne

gly

col)

fum

arat

e)(1

2.5

·9.

1.6

mm

)

MS

Cs

(PT

-251

0;L

on

za)

Co

llag

enI[

,C

oll

agen

III[

,T

NC

[,

Ten

asci

n-C

[27

Sab

er(2

010)

Lig

agen

L30

-4C

(Tis

sue

Gro

wth

Tec

hn

olo

gie

s)C

ycl

icst

retc

hin

g1.

25N

,1

cycl

e/m

in,

1h

/d

ayfo

r5

day

sA

cell

ula

rra

bb

ith

ind

paw

ten

do

n(l

eng

th:

5cm

)R

abb

itte

no

cyte

sU

ltim

ate

ten

sile

stre

ss[

(clo

sed

tofl

esh

lyh

arv

este

dte

nd

on

),el

asti

cm

od

ulu

s[

36

Issa

(201

1)C

ust

om

-mad

eb

iore

acto

rC

ycl

icst

retc

hin

g2%

stra

in,

1h

/d

ay,

0.01

67H

zfo

r1

and

2w

eek

s.H

um

anu

mb

ilic

alv

ein

s(w

all

thic

kn

ess:

0.41

mm

)w

ith

Wh

arto

n’s

jell

ym

atri

xas

cen

tral

po

rtio

n(t

ota

lth

ick

nes

s:0.

75m

m)

Wis

tar

rat

bo

ne

mar

row

MS

Cs

Cel

lp

roli

fera

tio

n[

,te

nsi

lest

ren

gth

[28

Wo

on

(201

1)L

igag

enL

30-1

C&

Lig

agen

L30

-4C

(Tis

sue

Gro

wth

Tec

hn

olo

gie

s)L

30-1

C:

dy

nam

iclo

adin

g,

10N

,1

h/

day

,0.

0167

Hz

for

5d

ays

L30

-4C

:d

yn

amic

load

ing

,0.

625

N1.

25N

2.5

N,

1h

/d

ay,

0.01

67H

zfo

r3,

5,an

d8

day

s

Ace

llu

lar

hu

man

flex

or

ten

do

nsc

affo

lds

Hu

man

der

mal

fib

rob

last

sU

ltim

ate

ten

sile

stre

ss[

,el

asti

cm

od

ulu

s[31

LV

DT

,li

nea

rv

aria

ble

dif

fere

nti

altr

ansf

orm

ers;

[,

up

reg

ula

tio

n;

Y,

do

wn

reg

ula

tio

n.

141

234

tendon/ligament culture. To our knowledge, two relativelycomplete commercial bioreactor systems have been devel-oped recently: The Bose� ElectroForce� BioDynamic� sys-tem and the LigaGen system.

The Bose ElectroForce BioDynamic test instrument provi-des an accurate programmable uniaxial stretching stimula-tion and a controllable medium circulation environment toengineered tendon/ligament, which theoretically can beadjusted to mimic the in vivo biomechanical environment.With a load cell and optional laser micrometer, this biore-actor system is able to monitor the force/strain curve ofengineered tendon/ligament during the culture period. Twodifferent force and displacement ranges are available. Single-chamber and multiple-chamber systems with shared orindependent loading are optional. As there are culturechambers compatible to the Bose testing devices such asElectroForce 3200, biomechanical tests can be done at dif-ferent time points without disruption of the tissue culture(www.bose-electroforce.com).

The LigaGen system is a lightweight (< 3 kg) incubator-compatible bioreactor. It is capable of applying a maximumforce of 40 N to the tissue sample and simulating complex,and presumably more physiologically realistic, loading pat-terns. Two systems are available from LigaGen. L30-1 · is asingle-culture model, which has a 23-mL-internal-volumechamber for single-tissue culture, and the L30-4C is a multi-ple-culture model with an 80-mL chamber for shared dynamicculture on two or four samples. The standard medium circu-lation system can reduce contamination risk during mediumexchange. Rather than being a comprehensive system, thebioreactor is extensible to suit individual needs, by addingvarious components to the (universal) basic model as re-quired. With an accessory tissue-monitoring sensor, this sys-tem is able to achieve real-time measurements of the samplestiffness during culture. If flow control is necessary, extracontrol systems can be installed (www.tissuegrowth.com/).

However, there are some disadvantages in the commercialbioreactors. First, the fixation mechanism to stabilize thetendon tissue in the bioreactor cannot be adjusted. Tissueclamps provided in the commercial bioreactor systems canbecome less effective when it comes to the use of a cylindricalscaffold. Secondly, capacities of the chamber and mechanicalinput are limited, and they can only host small-animal tis-sues. This may restrict clinical development. Moreover, someligaments, such as the anterior cruciate ligament, are not onlysubjected to tensile force, but also to rotational loading, andnone of the commercial bioreactor systems provide the ad-dition of torsional loading. Lastly, full-scale commercialbioreactor systems are expensive, and for most of the time,not all of the functions they provide are commonly used inevery study.

Ideal Bioreactors for Tendon/Ligament Engineering

The ideal bioreactor should be able culture tendon- andligament-like constructs, which are well-organized, cell-seeded assemblies of collagen bundles with mechanical prop-erties functionally similar to the native tissue. Autologouscell-seeded constructs are biologically compatible, and ableto provide mechanical support similar to native tissue, andconsequently, they are widely studied in tendon/ligamentengineering. However, induction of cell-directed collagen

fiber reorganization and assembly of collagen bundles aretwo important impediments to this approach. Mechanicaltensile loads can help provide the necessary signals to cells toincrease collagen synthesis, spatially organize of the collagenalong the primary stress direction, and stabilize collagenfrom collagenase degradation.22,96,97 Moreover, proper me-chanical stimulation can upregulate different proteoglycans,such as decorin, biglycan, fibromodulin, and fibronectin,which help the cells organize the parallel collagen fibrilsforming bundles.42,98–101

The host body is in many ways the ultimate bioreactor forall engineered tissues. The study of Juncosa-Melvin et al.indicated that the maximum force of engineered patellartendons increased more than 3000 times after 2 weeks ofimplantation in rabbits,19 and to date, none of the bioreactorshave been able to accomplish this outcome. Therefore, anideal bioreactor should aim to mimic the dynamic bio-chemical and biophysical environments in vivo. In musculo-skeletal tissue engineering, various bioreactors have beendeveloped. For instance, muscle tissues are not only sub-jected to mechanical stretching, but also able to receive theelectrical impulses to simulate inputs from the central ner-vous system,29 and mechanical and electrical stimulationbioreactors have been developed based on mimicking thein vivo environment.102,103 Compared to muscle, the in vivoenvironment is comparatively less complex in tendon/liga-ment and appears to require only passive mechanical input.

Summarizing, the ideal tendon-/ligament-engineeringbioreactor that enables systemic research should integrate allaforementioned components (Fig. 1). The bioreactor shouldhave culture chambers and an actuating system, but also befitted with a medium circulation system, an environmentalsystem, a monitoring system, a feedback system, and a wastemedium analysis system. This bioreactor should first be ableto provide not only a multiple, suitably sized. and sterilizedchambers for tissue culture, but also accurate and program-mable mechanical stimulation. Tensile strain and rotation areneeded to mimic different in vivo tendon/ligament loads.Second, the circulated medium should infiltrate into tissuebetter than a static medium configuration, and the circulationsystem needs to reduce the risks of contamination duringmedium exchange and drug delivery. Moreover, environ-mental control, such as PO2, CO2, and pH level, allows re-searchers to explore the impact of different cultureconditions on tissue maturation. Third, a monitoring systemshould provide the real-time status of cultured tendon/lig-ament, such as force and displacement, and based on thesedata, adjustment of mechanical stimulation by use of afeedback system. Fourth, through analysis of the waste me-dium, nutrient consumption needs to be evaluated, whichmay enable the changing of the medium base at differentstages as required, rather than fixed, regular medium ex-change every 3 days. Finally and importantly, the best pat-terns of mechanical stimulation for culturing engineeredtendon/ligaments need to be defined.

Investigation of Evidence

Since the first 3D engineered tendon/ligament bioreactorsystem published by Altman et al. in 2002, the effect of me-chanical stimulation on the engineered tendon/ligament hasdrawn a lot of research attention. In the past decade,

142 WANG ET AL.

235

dynamic loading of culture in bioreactor systems has beenproven to have been a significant development in producingan engineered tendon/ligament. Indeed, various studieshave been performed using bioreactor systems and have metwith considerably success, and these are summarized inTable 2. Compared with static culture, the tissue producedusing dynamic mechanical stimulation has a better cellmorphology, including elongated cellular morphology andincreased cell density. The mechanical properties of en-gineered tendon/ligament, such as tensile strength andelastic modulus, are also greatly improved by cyclic loadingof the tissue culture, as is the microstructure of the extra-cellular matrix, such as collagen fiber alignment. Gene ex-pression is also positively influenced by cyclic mechanicalstimulation. For instance, collagen I expression under dy-namic loading is three times higher than static culture in 2weeks.83 Although cyclic stretching has been proven to be aneffective way to stimulate the engineered tendon/ligamentculture, the optimal stimulation pattern is still unknown.Nirmalanandhan et al. revealed that a 2.4% strain cycleconsisting of 3000 cycles per day produced the best linearstiffness in rabbit MSCs seeded in type I collagen sponge.21

However, from the perspective of Butler et al., the stimula-tion pattern should be adjusted based on maturation of theengineered tendon/ligament, with higher dose loading ap-plied at the later stage of tissue culture.83

Conclusion

The goal should be to create a construct with similar mi-crostructure and mechanical properties, as native tissue, us-ing bioscaffolds and autologous tenocytes. However, atendon-like uniorientated collagen structure cannot beachieved without a mechanical stimulus within the cultureenvironment. Traditional culture techniques do not providethis mechanical stimulation. The use of a bioreactor system isable to bridge the gap between in vitro and in vivo systems bycreating suitable biochemical and biomechanical environ-ments. Although it is clearly very difficult to reproduce thein vivo microenvironmental conditions exactly within thebioreactor, the goal is to mimic the in vivo biomechanicalcondition as closely as possible. The essential components ofa tendon-/ligament-engineering bioreactor are the actuatingsystem and culture chamber, which are responsible for themechanical stimulation and providing sterilized environ-ment for tissue culture. For better manipulation of the cultureenvironment, accurate stimulation and more precise report-ing of mechanical maturation of engineered tendon/liga-ment, an environmental control system, medium circulationsystem, monitor, and feedback system need to be included.As bioreactors nowadays are becoming more focused onpreclinical research, using these clinically in the future stillpresents a substantial challenge. Tissue engineering requiressubstantial system optimization to achieve a reproduciblefunctional engineered tendon/ligament consistently. Thisprocess is difficult and expensive in animal models, let alonein humans. However, there are additional problems inmoving from small-animal models to humans related to thephysical size of human tissue samples. Issues of nutrienttransport, cell source, and spatial heterogeneity of scaffoldproperties and cell stimulation become more prominent inthese larger tissues.

However, all the evidence suggests that by using biore-actors as described above, it will be possible to successfullyproduce engineered tendons and ligaments. When this oc-curs, we will be able to manufacture basic multichamberbioreactors with accurate and appropriate patterns of me-chanical stimulation, which are affordable, and have lowmaintenance costs that can be used in commercial settings.

Acknowledgments

This work is supported by the Australia Research CouncilLinkage Grant (LP110100581). We would like to thank ourfellow group members, in particular Ms. Euphemie Landaoand Mr. Robert Day for helpful discussion.

Disclosure Statement

No competing financial interests exist.

References

1. Huang, H.H., Qureshi, A.A., and Biundo, J.J., Jr. Sports andother soft tissue injuries, tendinitis, bursitis, and occupation-related syndromes. Curr Opin Rheumatol 12, 150, 2000.

2. Chen, J.L., Yin, Z., Shen, W.L., Chen, X., Heng, B.C., Zou,X.H., and Ouyang, H.W. Efficacy of hESC-MSCs in knittedsilk-collagen scaffold for tendon tissue engineering andtheir roles. Biomaterials 31, 9438, 2010.

3. Butler, D.L., Gooch, C., Kinneberg, K.R., Boivin, G.P.,Galloway, M.T., Nirmalanandhan, V.S., Shearn, J.T.,Dyment, N.A., and Juncosa-Melvin, N. The use of mesen-chymal stem cells in collagen-based scaffolds for tissue-engineered repair of tendons. Nat Protoc 5, 849, 2010.

4. Chen, J., Xu, J., Wang, A., and Zheng, M. Scaffolds fortendon and ligament repair: review of the efficacy ofcommercial products. Expert Rev Med Devices 6, 61,2009.

5. Cerullo, G., Puddu, G., Gianni, E., Damiani, A., and Pi-gozzi, F. Anterior cruciate ligament patellar tendon recon-struction: it is probably better to leave the tendon defectopen! Knee Surg Sports Traumatol Arthrosc 3, 14, 1995.

6. Coupens, S.D., Yates, C.K., Sheldon, C., and Ward, C.Magnetic resonance imaging evaluation of the patellartendon after use of its central one-third for anterior cruciateligament reconstruction. Am J Sports Med 20, 332, 1992.

7. Harner, C.D., Olson, E., Irrgang, J.J., Silverstein, S., Fu, F.H.,and Silbey, M. Allograft versus autograft anterior cruciateligament reconstruction: 3- to 5-year outcome. Clin OrthopRelat Res 324, 134, 1996.

8. Lin, T.W., Cardenas, L., and Soslowsky, L.J. Biomechanicsof tendon injury and repair. J Biomech 37, 865, 2004.

9. Jarvinen, T.A., Jarvinen, T.L., Kannus, P., Jozsa, L., andJarvinen, M. Collagen fibres of the spontaneously rupturedhuman tendons display decreased thickness and crimpangle. J Orthop Res 22, 1303, 2004.

10. Buckwalter, J.A. and Hunziker, E.B. Orthopaedics. Healingof bones, cartilages, tendons, and ligaments: a new era.Lancet 348 Suppl 2, sII18, 1996.

11. Hildebrand, K.A., Frank, C.B., and Hart, D.A. Gene inter-vention in ligament and tendon: current status, challenges,future directions. Gene Ther 11, 368, 2004.

12. Bray, R.C., Rangayyan, R.M., and Frank, C.B. Normal andhealing ligament vascularity: a quantitative histologicalassessment in the adult rabbit medial collateral ligament. JAnat 188 (Pt 1), 87, 1996.

BIOREACTOR DESIGN FOR TENDON/LIGAMENT ENGINEERING 143

236

13. Lo, I.K., Ou, Y., Rattner, J.P., Hart, D.A., Marchuk, L.L., Frank,C.B., and Rattner, J.B. The cellular networks of normal ovinemedial collateral and anterior cruciate ligaments are not ac-curately recapitulated in scar tissue. J Anat 200, 283, 2002.

14. Manske, P.R. Flexor tendon healing. J Hand Surg Br 13, 237,1988.

15. Arnsdorf, E.J., Jones, L.M., Carter, D.R., and Jacobs, C.R.The periosteum as a cellular source for functional tissueengineering. Tissue Eng Part A 15, 2637, 2009.

16. Fan, H., Liu, H., Wong, E.J., Toh, S.L., and Goh, J.C. In vivostudy of anterior cruciate ligament regeneration usingmesenchymal stem cells and silk scaffold. Biomaterials 29,

3324, 2008.17. Fan, H., Liu, H., Toh, S.L., and Goh, J.C. Anterior cruciate

ligament regeneration using mesenchymal stem cells and silkscaffold in large animal model. Biomaterials 30, 4967, 2009.

18. Garcia-Fuentes, M., Meinel, A.J., Hilbe, M., Meinel, L., andMerkle, H.P. Silk fibroin/hyaluronan scaffolds for humanmesenchymal stem cell culture in tissue engineering. Bio-materials 30, 5068, 2009.

19. Juncosa-Melvin, N., Shearn, J.T., Boivin, G.P., Gooch, C.,Galloway, M.T., West, J.R., Nirmalanandhan, V.S., Bradica,G., and Butler, D.L. Effects of mechanical stimulation on thebiomechanics and histology of stem cell-collagen spongeconstructs for rabbit patellar tendon repair. Tissue Eng 12,

2291, 2006.20. Nirmalanandhan, V.S., Rao, M., Shearn, J.T., Juncosa-

Melvin, N., Gooch, C., and Butler, D.L. Effect of scaffoldmaterial, construct length and mechanical stimulation onthe in vitro stiffness of the engineered tendon construct. JBiomech 41, 822, 2008.

21. Nirmalanandhan, V.S., Shearn, J.T., Juncosa-Melvin, N.,Rao, M., Gooch, C., Jain, A., Bradica, G., and Butler, D.L.Improving linear stiffness of the cell-seeded collagensponge constructs by varying the components of the me-chanical stimulus. Tissue Eng Part A 14, 1883, 2008.

22. Nguyen, T.D., Liang, R., Woo, S.L., Burton, S.D., Wu, C.,Almarza, A., Sacks, M.S., and Abramowitch, S. Effects ofcell seeding and cyclic stretch on the fiber remodeling in anextracellular matrix-derived bioscaffold. Tissue Eng Part A15, 957, 2009.

23. Abousleiman, R.I., Reyes, Y., McFetridge, P., and Sika-vitsas, V. Tendon tissue engineering using cell-seededumbilical veins cultured in a mechanical stimulator. TissueEng Part A 15, 787, 2009.

24. Sahoo, S., Ouyang, H., Goh, J.C., Tay, T.E., and Toh, S.L.Characterization of a novel polymeric scaffold for potentialapplication in tendon/ligament tissue engineering. TissueEng 12, 91, 2006.

25. Sahoo, S., Cho-Hong, J.G., and Siew-Lok, T. Development ofhybrid polymer scaffolds for potential applications in ligamentand tendon tissue engineering. Biomed Mater 2, 169, 2007.

26. Chen, X., Qi, Y.Y., Wang, L.L., Yin, Z., Yin, G.L., Zou, X.H.,and Ouyang, H.W. Ligament regeneration using a knittedsilk scaffold combined with collagen matrix. Biomaterials29, 3683, 2008.

27. Doroski, D.M., Levenston, M.E., and Temenoff, J.S. Cyclictensile culture promotes fibroblastic differentiation ofmarrow stromal cells encapsulated in poly(ethylene gly-col)-based hydrogels. Tissue Eng Part A 16, 3457, 2010.

28. Issa, R.I., Engebretson, B., Rustom, L., McFetridge, P.S., andSikavitsas, V.I. The effect of cell seeding density on thecellular and mechanical properties of a mechanostimulatedtissue-engineered tendon. Tissue Eng Part A 17, 1479, 2011.

29. Ross, J.J., Duxson, M.J., and Harris, A.J. Neural determi-nation of muscle fibre numbers in embryonic rat lumbricalmuscles. Development 100, 395, 1987.

30. Liu, W., Chen, B., Deng, D., Xu, F., Cui, L., and Cao, Y.Repair of tendon defect with dermal fibroblast engineeredtendon in a porcine model. Tissue Eng 12, 775, 2006.

31. Woon, C.Y., Kraus, A., Raghavan, S.S., Pridgen, B.C., Me-gerle, K., Pham, H., and Chang, J. Three-dimensional-construct bioreactor conditioning in human tendon tissueengineering. Tissue Eng Part A 17, 2561, 2011.

32. Deng, D., Liu, W., Xu, F., Yang, Y., Zhou, G., Zhang, W.J.,Cui, L., and Cao, Y. Engineering human neo-tendon tissuein vitro with human dermal fibroblasts under static me-chanical strain. Biomaterials 30, 6724, 2009.

33. Obaid, H. and Connell, D. Cell therapy in tendon disorders:what is the current evidence? Am J Sports Med 38, 2123, 2010.

34. Chen, J., Wang, A., Xu, J., and Zheng, M. In chronic lateralepicondylitis, apoptosis and autophagic cell death occur inthe extensor carpi radialis brevis tendon. J Shoulder ElbowSurg 19, 355, 2010.

35. Androjna, C., Spragg, R.K., and Derwin, K.A. Mechanicalconditioning of cell-seeded small intestine submucosa: apotential tissue-engineering strategy for tendon repair.Tissue Eng 13, 233, 2007.

36. Saber, S., Zhang, A.Y., Ki, S.H., Lindsey, D.P., Smith, R.L.,Riboh, J., Pham, H., and Chang, J. Flexor tendon tissueengineering: bioreactor cyclic strain increases constructstrength. Tissue Eng Part A 16, 2085, 2010.

37. Chen, J.M., Willers, C., Xu, J., Wang, A., and Zheng, M.H.Autologous tenocyte therapy using porcine-derived bios-caffolds for massive rotator cuff defect in rabbits. TissueEng 13, 1479, 2007.

38. Cooper, J.A., Lu, H.H., Ko, F.K., Freeman, J.W., and Laur-encin, C.T. Fiber-based tissue-engineered scaffold for liga-ment replacement: design considerations and in vitroevaluation. Biomaterials 26, 1523, 2005.

39. Freeman, J.W., Woods, M.D., and Laurencin, C.T. Tissueengineering of the anterior cruciate ligament using a braid-twist scaffold design. J Biomech 40, 2029, 2007.

40. Lee, C.H., Shin, H.J., Cho, I.H., Kang, Y.M., Kim, I.A., Park,K.D., and Shin, J.W. Nanofiber alignment and direction ofmechanical strain affect the ECM production of humanACL fibroblast. Biomaterials 26, 1261, 2005.

41. Moffat, K.L., Kwei, A.S., Spalazzi, J.P., Doty, S.B., Levine,W.N., and Lu, H.H. Novel nanofiber-based scaffold for rotatorcuff repair and augmentation. Tissue Eng Part A 15, 115, 2009.

42. Webb, K., Hitchcock, R.W., Smeal, R.M., Li, W., Gray, S.D.,and Tresco, P.A. Cyclic strain increases fibroblast prolifera-tion, matrix accumulation, and elastic modulus of fibroblast-seeded polyurethane constructs. J Biomech 39, 1136, 2006.

43. Zheng, M., and Wang, A., Autologous Cell Therapy for Ten-don Tissue Reconstruction, in Sino-American Symposium onClinical and Translational Medicine. Washington, DC, 2011.

44. Zheng, M., Tenocyte cell culturing method, 2009.US7985408 (Issued date July 26, 2011).

45. Freed, L.E., Vunjak-Novakovic, G., and Langer, R. Culti-vation of cell-polymer cartilage implants in bioreactors. JCell Biochem 51, 257, 1993.

46. Freed, L.E., Marquis, J.C., Nohria, A., Emmanual, J., Mikos,A.G., and Langer, R. Neocartilage formation in vitro and invivo using cells cultured on synthetic biodegradable poly-mers. J Biomed Mater Res 27, 11, 1993.

47. Awad, H.A., Butler, D.L., Harris, M.T., Ibrahim, R.E., Wu,Y., Young, R.G., Kadiyala, S., and Boivin, G.P. In vitro

144 WANG ET AL.

237

characterization of mesenchymal stem cell-seeded collagenscaffolds for tendon repair: effects of initial seeding densityon contraction kinetics. J Biomed Mater Res 51, 233, 2000.

48. Wang, L., Seshareddy, K., Weiss, M.L., and Detamore, M.S.Effect of initial seeding density on human umbilical cordmesenchymal stromal cells for fibrocartilage tissue engi-neering. Tissue Eng Part A 15, 1009, 2009.

49. Zhang, L., Gardiner, B.S., Smith, D.W., Pivonka, P., andGrodzinsky, A. A fully coupled poroelastic reactive-trans-port model of cartilage. Mol Cell Biomech 5, 133, 2008.

50. Pham, Q.P., Sharma, U., and Mikos, A.G. Electrospinningof polymeric nanofibers for tissue engineering applications:a review. Tissue Eng 12, 1197, 2006.

51. Derwin, K.A., Baker, A.R., Spragg, R.K., Leigh, D.R., andIannotti, J.P. Commercial extracellular matrix scaffolds forrotator cuff tendon repair. Biomechanical, biochemical, andcellular properties. J Bone Joint Surg Am 88, 2665, 2006.

52. Kinneberg, K.R., Nirmalanandhan, V.S., Juncosa-Melvin,N., Powell, H.M., Boyce, S.T., Shearn, J.T., and Butler, D.L.Chondroitin-6-sulfate incorporation and mechanical stim-ulation increase MSC-collagen sponge construct stiffness. JOrthop Res 28, 1092, 2010.

53. Gilbert, T.W., Stewart-Akers, A.M., Simmons-Byrd, A., andBadylak, S.F. Degradation and remodeling of small intes-tinal submucosa in canine Achilles tendon repair. J BoneJoint Surg Am 89, 621, 2007.

54. Fleming, B.C., Spindler, K.P., Palmer, M.P., Magarian, E.M.,and Murray, M.M. Collagen-platelet composites improvethe biomechanical properties of healing anterior cruciateligament grafts in a porcine model. Am J Sports Med 37,

1554, 2009.55. Malcarney, H.L., Bonar, F., and Murrell, G.A. Early in-

flammatory reaction after rotator cuff repair with a porcinesmall intestine submucosal implant: a report of 4 cases. AmJ Sports Med 33, 907, 2005.

56. Zheng, M.H., Chen, J., Kirilak, Y., Willers, C., Xu, J., andWood, D. Porcine small intestine submucosa (SIS) is not anacellular collagenous matrix and contains porcine DNA:possible implications in human implantation. J BiomedMater Res B Appl Biomater 73, 61, 2005.

57. Chang, J., Thunder, R., Most, D., Longaker, M.T., andLineaweaver, W.C. Studies in flexor tendon wound heal-ing: neutralizing antibody to TGF-beta1 increases postop-erative range of motion. Plast Reconstr Surg 105, 148, 2000.

58. Bennett, N.T., and Schultz, G.S. Growth factors and woundhealing: biochemical properties of growth factors and theirreceptors. Am J Surg 165, 728, 1993.

59. Wojciak, B., and Crossan, J.F. The effects of T cells and theirproducts on in vitro healing of epitenon cell microwounds.Immunology 83, 93, 1994.

60. Marui, T., Niyibizi, C., Georgescu, H.I., Cao, M., Kavalk-ovich, K.W., Levine, R.E., and Woo, S.L. Effect of growthfactors on matrix synthesis by ligament fibroblasts. J Or-thop Res 15, 18, 1997.

61. Pryce, B.A., Watson, S.S., Murchison, N.D., Staverosky, J.A.,Dunker, N., and Schweitzer, R. Recruitment and mainte-nance of tendon progenitors by TGFbeta signaling are es-sential for tendon formation. Development 136, 1351, 2009.

62. Clark, R.T., Johnson, T.L., Schalet, B.J., Davis, L., Gaschen,V., Hunziker, E.B., Oldberg, A., and Mikic, B. GDF-5 defi-ciency in mice leads to disruption of tail tendon form andfunction. Connect Tissue Res 42, 175, 2001.

63. Mikic, B., Schalet, B.J., Clark, R.T., Gaschen, V., and Hun-ziker, E.B. GDF-5 deficiency in mice alters the ultrastruc-

ture, mechanical properties and composition of the Achillestendon. J Orthop Res 19, 365, 2001.

64. Chhabra, A., Tsou, D., Clark, R.T., Gaschen, V., Hunziker,E.B., and Mikic, B. GDF-5 deficiency in mice delays Achillestendon healing. J Orthop Res 21, 826, 2003.

65. Dahlgren, L.A., Mohammed, H.O., and Nixon, A.J. Tem-poral expression of growth factors and matrix molecules inhealing tendon lesions. J Orthop Res 23, 84, 2005.

66. Kobayashi, M., Itoi, E., Minagawa, H., Miyakoshi, N., Taka-hashi, S., Tuoheti, Y., Okada, K., and Shimada, Y. Expressionof growth factors in the early phase of supraspinatus tendonhealing in rabbits. J Shoulder Elbow Surg 15, 371, 2006.

67. Chamberlain, C.S., Crowley, E., and Vanderby, R. Thespatio-temporal dynamics of ligament healing. WoundRepair Regen 17, 206, 2009.

68. Hannafin, J.A., Arnoczky, S.P., Hoonjan, A., and Torzilli,P.A. Effect of stress deprivation and cyclic tensile loadingon the material and morphologic properties of canine flexordigitorum profundus tendon: an in vitro study. J OrthopRes 13, 907, 1995.

69. Flynn, B.P., Bhole, A.P., Saeidi, N., Liles, M., Dimarzio,C.A., and Ruberti, J.W. Mechanical strain stabilizes recon-stituted collagen fibrils against enzymatic degradation bymammalian collagenase matrix metalloproteinase 8 (MMP-8). PLoS One 5, e12337, 2010.

70. Maeda, T., Sakabe, T., Sunaga, A., Sakai, K., Rivera, A.L.,Keene, D.R., Sasaki, T., Stavnezer, E., Iannotti, J., Schweit-zer, R., Ilic, D., Baskaran, H., and Sakai, T. Conversion ofmechanical force into TGF-beta-mediated biochemical sig-nals. Curr Biol 21, 933, 2011.

71. Lejard, V., Brideau, G., Blais, F., Salingcarnboriboon, R.,Wagner, G., Roehrl, M.H., Noda, M., Duprez, D., Houillier,P., and Rossert, J. Scleraxis and NFATc regulate the ex-pression of the pro-alpha1(I) collagen gene in tendon fi-broblasts. J Biol Chem 282, 17665, 2007.

72. Dennis, R.G., Smith, B., Philp, A., Donnelly, K., and Baar,K. Bioreactors for guiding muscle tissue growth and de-velopment. Adv Biochem Eng Biotechnol 112, 39, 2009.

73. Catapano, G., Patzer, J.F., 2nd, and Gerlach, J.C. Transportadvances in disposable bioreactors for liver tissue engi-neering. Adv Biochem Eng Biotechnol 115, 117, 2010.

74. Rauh, J., Milan, F., Gunther, K.P., and Stiehler, M. Bior-eactor systems for bone tissue engineering. Tissue Eng PartB Rev 17, 263, 2011.

75. Salter, E., Goh, B., Hung, B., Hutton, D., Ghone, N., andGrayson, W.L. Bone tissue engineering bioreactors: a role inthe clinic? Tissue Eng Part B Rev 18, 62, 2012.

76. Altman, G.H., Lu, H.H., Horan, R.L., Calabro, T., Ryder, D.,Kaplan, D.L., Stark, P., Martin, I., Richmond, J.C., andVunjak-Novakovic, G. Advanced bioreactor with con-trolled application of multi-dimensional strain for tissueengineering. J Biomech Eng 124, 742, 2002.

77. Varseveld, R.B.v. and Bone, G.M. Accurate Position Con-trol of a Pneumatic Actuator Using On/Off SolenoidValves. Trans Mechatronics 2, 195, 1997.

78. Alter, D. and Tsao, T. Control of linear motors for machine toolfeed drives: design and implementation of H1 optimal feed-back control. ASME J Dyn Syst, Meas, Control. 118, 649, 1996.

79. Alter, D., and Tsao, T., Dynamic stiffness enhancement ofdirect linear motor feed drives for machining, in Proceedingsof the American Control Conference, Baltimore, MD, 1994.

80. Braembussche, P., Swevers, J., Brussel, H., and Vanherck, P.Accurate tracking control of linear synchronous motormachine tool axes. Mechatronics 6, 507, 1996.

BIOREACTOR DESIGN FOR TENDON/LIGAMENT ENGINEERING 145

238

81. Liu, J., Hsu, M., and Chen, F. On the design of rotatingspeed functions to improve the acceleration peak value ofball-screw transmission mechanism. Mech Machine Theory36, 1035, 2001.

82. Parent, G., Huppe, N., and Langelier, E. Low stress tendonfatigue is a relatively rapid process in the context of over-use injuries. Ann Biomed Eng 39, 1535, 2011.

83. Butler, D.L., Hunter, S.A., Chokalingam, K., Cordray, M.J.,Shearn, J., Juncosa-Melvin, N., Nirmalanandhan, S., andJain, A. Using functional tissue engineering and bioreactorsto mechanically stimulate tissue-engineered constructs.Tissue Eng Part A 15, 741, 2009.

84. Janssen, F.W., Oostra, J., Oorschot, A., and van Blitterswijk,C.A. A perfusion bioreactor system capable of producingclinically relevant volumes of tissue-engineered bone: invivo bone formation showing proof of concept. Biomater-ials 27, 315, 2006.

85. Cartmell, S.H., Porter, B.D., Garcia, A.J., and Guldberg, R.E.Effects of medium perfusion rate on cell-seeded three-dimensional bone constructs in vitro. Tissue Eng 9, 1197, 2003.

86. Meinel, L., Karageorgiou, V., Fajardo, R., Snyder, B.,Shinde-Patil, V., Zichner, L., Kaplan, D., Langer, R., andVunjak-Novakovic, G. Bone tissue engineering using hu-man mesenchymal stem cells: effects of scaffold materialand medium flow. Ann Biomed Eng 32, 112, 2004.

87. Uemura, T., Dong, J., Wang, Y., Kojima, H., Saito, T., Iejima,D., Kikuchi, M., Tanaka, J., and Tateishi, T. Transplantationof cultured bone cells using combinations of scaffolds andculture techniques. Biomaterials 24, 2277, 2003.

88. Holtorf, H.L., Jansen, J.A., and Mikos, A.G. Flow perfusionculture induces the osteoblastic differentiation of marrowstroma cell-scaffold constructs in the absence of dexa-methasone. J Biomed Mater Res A 72, 326, 2005.

89. Olivier, V., Hivart, P., Descamps, M., and Hardouin, P. Invitro culture of large bone substitutes in a new bioreactor:importance of the flow direction. Biomed Mater 2, 174,2007.

90. Saini, S. and Wick, T.M. Effect of low oxygen tension ontissue-engineered cartilage construct development in theconcentric cylinder bioreactor. Tissue Eng 10, 825, 2004.

91. Domm, C., Schunke, M., Christesen, K., and Kurz, B. Re-differentiation of dedifferentiated bovine articular chon-drocytes in alginate culture under low oxygen tension.Osteoarthritis Cartilage 10, 13, 2002.

92. Hansen, U., Schunke, M., Domm, C., Ioannidis, N.,Hassenpflug, J., Gehrke, T., and Kurz, B. Combination ofreduced oxygen tension and intermittent hydrostatic pres-sure: a useful tool in articular cartilage tissue engineering.J Biomech 34, 941, 2001.

93. Butler, D.L., Juncosa-Melvin, N., Boivin, G.P., Galloway,M.T., Shearn, J.T., Gooch, C., and Awad, H. Functionaltissue engineering for tendon repair: a multidisciplinarystrategy using mesenchymal stem cells, bioscaffolds, andmechanical stimulation. J Orthop Res 26, 1, 2008.

94. Goetz, M., Thomas, S., Heimann, A., Delaney, P., Schnei-der, C., Relle, M., Schwarting, A., Galle, P.R., Kempski, O.,Neurath, M.F., and Kiesslich, R. Dynamic in vivo imagingof microvasculature and perfusion by miniaturized confo-cal laser microscopy. Eur Surg Res 41, 290, 2008.

95. Drexler, W., Morgner, U., Kartner, F.X., Pitris, C., Boppart,S.A., Li, X.D., Ippen, E.P., and Fujimoto, J.G. In vivo ul-trahigh-resolution optical coherence tomography. Opt Lett24, 1221, 1999.

96. Yang, G., Im, H.J., and Wang, J.H. Repetitive mechanicalstretching modulates IL-1beta induced COX-2, MMP-1 ex-pression, and PGE2 production in human patellar tendonfibroblasts. Gene 363, 166, 2005.

97. Lavagnino, M., Arnoczky, S.P., Tian, T., and Vaupel, Z.Effect of amplitude and frequency of cyclic tensile strain onthe inhibition of MMP-1 mRNA expression in tendon cells:an in vitro study. Connect Tissue Res 44, 181, 2003.

98. Cribb, A.M. and Scott, J.E. Tendon response to tensile stress:an ultrastructural investigation of collagen:proteoglycan in-teractions in stressed tendon. J Anat 187 (Pt 2), 423, 1995.

99. Scott, J.E. The periphery of the developing collagen fibril.Quantitative relationships with dermatan sulphate andother surface-associated species. Biochem J 218, 229, 1984.

100. Vogel, K.G. What happens when tendons bend andtwist? Proteoglycans. J Musculoskelet Neuronal Interact 4,

202, 2004.101. Franchi, M., Trire, A., Quaranta, M., Orsini, E., and Ottani,

V. Collagen structure of tendon relates to function. SciWorld J 7, 404, 2007.

102. Sharifpoor, S., Simmons, C.A., Labow, R.S., and Paul Santerre,J. Functional characterization of human coronary arterysmooth muscle cells under cyclic mechanical strain in a de-gradable polyurethane scaffold. Biomaterials 32, 4816, 2011.

103. Donnelly, K., Khodabukus, A., Philp, A., Deldicque, L.,Dennis, R.G., and Baar, K. A novel bioreactor for stimu-lating skeletal muscle in vitro. Tissue Eng Part C Methods16, 711, 2010.

Address correspondence to:Ming H. Zheng, MD, PhD, FRCPath, FRCPA

Centre for Orthopaedic ResearchSchool of Surgery

The University of Western AustraliaM Block, QE2 Medical Centre

Nedlands, WA 6009Australia

E-mail: [email protected]

David G. Lloyd, PhDCentre for Musculoskeletal Research

Griffith Health InstituteClinical Science 1 (G02) Room 2.40

Griffith UniversityGold Coast Campus, QLD 4222

Australia

E-mail: [email protected]

David W. Smith, PhDSchool of Computer Science and Software Engineering

The University of Western Australia (M002)35 Stirling Highway

Crawley, WA 6009Australia

E-mail: [email protected]

Received: May 17, 2012Accepted: September 25, 2012

Online Publication Date: November 19, 2012

146 WANG ET AL.

239

ARTICLE

Programmable Mechanical Stimulation InfluencesTendon Homeostasis in a Bioreactor System

Tao Wang,1 Zhen Lin,1,2 Robert E. Day,3 Bruce Gardiner,4 Euphemie Landao-Bassonga,1

Jonas Rubenson,5 Thomas B. Kirk,6 David W. Smith,4 David G. Lloyd,7 Gerard Hardisty,8

Allan Wang,9 Qiujian Zheng,2 Ming H. Zheng1

1Centre for Orthopaedic Translational Research, School of Surgery, University of Western

Australia, M Block, QE2 Medical Centre, Nedlands, Crawley, Western Australia 6009,

Australia; telephone: 61-8-93464050; fax: 61-8-93463210;

e-mail: [email protected] of Orthopaedic Surgery, Department of Surgery, Guangdong General Hospital,

Guangdong Academy of Medicine Science, Guangzhou, Guangdong, China3Department of Medical Engineering and Physics, Royal Perth Hospital, Perth, Western

Australia, Australia4School of Computer Science and Software Engineering, University of Western Australia

(M002), Crawley, Western Australia, Australia5School of Sport Science, Exercise and Health, Musculoskeletal Tissue Mechanics

and Muscle Energetics University of Western Australia (M408), Crawley,

Western Australia, Australia6Curtin University, Bentley, Chancellory, Western Australia, Australia7Centre for Musculoskeletal Research, Griffith Health Institute, Griffith University,

Gold Coast Campus, Queensland, Australia8St John of God Medical Clinic, Perth, Subiaco, Western Australia, Australia9Sir Charles Gairdner Hospital, Perth, Nedlands, Western Australia, Australia

ABSTRACT: Identification of functional programmable me-chanical stimulation (PMS) on tendon not only provides theinsight of the tendon homeostasis under physical/patholog-ical condition, but also guides a better engineering strategyfor tendon regeneration. The aims of the study are to designa bioreactor system with PMS to mimic the in vivo loadingconditions, and to define the impact of different cyclictensile strain on tendon. Rabbit Achilles tendons wereloaded in the bioreactor with/without cyclic tensile loading(0.25Hz for 8 h/day, 0–9% for 6 days). Tendons withoutloading lost its structure integrity as evidenced by disor-ientated collagen fiber, increased type III collagen expres-sion, and increased cell apoptosis. Tendons with 3% of cyclictensile loading had moderate matrix deterioration andelevated expression levels of MMP-1, 3, and 12, whilst

exceeded loading regime of 9% caused massive rupture ofcollagen bundle. However, 6% of cyclic tensile strain wasable to maintain the structural integrity and cellular func-tion. Our data indicated that an optimal PMS is required tomaintain the tendon homeostasis and there is only a narrowrange of tensile strain that can induce the anabolic action.The clinical impact of this study is that optimized eccentrictraining program is needed to achieve maximum beneficialeffects on chronic tendinopathy management.

Biotechnol. Bioeng. 2013;110: 1495–1507.

� 2012 Wiley Periodicals, Inc.

KEYWORDS: bioreactor; programmable mechanical stimu-lation; tendon; collagen

Introduction

Tendinopathy is common in both young athletes and theelderly and is thought to be due to overuse and/oroverloading. Tendinopathy causes significant pain andcan lead to chronic degenerative changes in the tendon

Correspondence to: M. H. Zheng

Contract grant sponsor: Australia Research Council Linkage

Contract grant number: LP110100581

Received 28 September 2012; Revision received 27 November 2012;

Accepted 7 December 2012

Accepted manuscript online 14 December 2012;

Article first published online 4 February 2013 in Wiley Online Library

(http://onlinelibrary.wiley.com/doi/10.1002/bit.24809/abstract)

DOI 10.1002/bit.24809

� 2012 Wiley Periodicals, Inc. Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013 1495

240

matrix (Smith, 2000), and potentially tendon rupture (Leaand Smith, 1972; Leppilahti and Orava, 1998). Achillestendinopathy is one of the most common tendinopathiesand has the greatest potential to restrict the activities in dailyliving. Treatment outcomes are often poor. These poorresults are largely due to the underlying physiologicalprocesses within tendons, which are not well understood.

Tissue engineering methods such as cell-seeded bioscaf-folds have great potential in the repair of damaged tendons.However, these methods are still very much in their infancy,due to the lack of an integrated understanding of the keyconditions required for successful tendinopathy interven-tion and repair. Recent tendon research has focused on thedevelopment of tendon tissue engineering, with the aim ofconstructing a bio-substitute material for tendon/ligamentrepair (Chen et al., 2007, 2008; Juncosa-Melvin et al., 2006;Nirmalanandhan et al., 2008a; Peach et al., 2012; Saber et al.,2010; Sahoo et al., 2006; Stoll et al., 2010; Wang et al., 2012;Webb et al., 2006; Woon et al., 2012); This strategy involvesan integration of technologies from cell biology, materialscience, and mechanical engineering.

Engineered bioreactor systems have been demonstrated tobe effective tools for constructing various biological tissuesand organs such as muscle (Dennis et al., 2009), liver(Catapano et al., 2010), and bone (Rauh et al., 2011, 2012).Several bioreactor systems have been developed specificallyfor engineered tendon/ligament formation (Altman et al.,2002; Issa et al., 2011; Juncosa-Melvin et al., 2006;Nirmalanandhan et al., 2008a,b; Webb et al., 2006).Owing to force transmission being the primary functionrole of tendon/ligament, mechanical stimulation plays anessential role in the maintenance of tendon structure andfunction (Hannafin et al., 1995). Achilles tendons enduretensile loading in vivo during daily physical activities. Morecomplex structures, like the anterior cruciate ligament,undergo both tensile and torsion loading during weight-bearing knee flexion (Li et al., 2005). To restore tendonfunction after injury, engineered tendon constructs shouldbe compatible with the in vivo environment bothbiologically and mechanically (Calve et al., 2010). Cell-scaffold constructs may be ‘‘exercised’’ in an artificialmechanical environment before implantation. Indeed,programmable mechanical stimulation (PMS) is consideredto be one of the key components for tendon/ligamentforming bioreactor system (Altman et al., 2002; Issa et al.,2011; Juncosa-Melvin et al., 2006; Nirmalanandhan et al.,2008a,b; Webb et al., 2006). In this study, a bioreactorsystem was designed to apply different cyclic tensile strain onisolated rabbit Achilles tendon to establish an ex vivo model.The present aim is to evaluate the effect of differentmechanical stimulation on tendon tissue.

The mechanical stimulation of engineered tendonconstructs can result in different structures. Non-loadingof engineered neo-tendons results in loose, disorganizedmatrices with reduced collagen quantity (Jiang et al., 2011).Static mechanical stress has been shown to improve the fiberarrangement of engineered tendon constructs, but did not

produce a compacted tendon-like collagen structure in vitro(Chen et al., 2007). More promisingly, engineered tendonunder cyclic tensile strain exhibited a well-organizedcollagen fiber arrangement and structure and enhancedtensile strength, suggesting that PMS is preferable forengineered tendon formation (Jiang et al., 2011). Eventhough the importance of providing PMS in tendonbioreactor has been recognized previously (Brown andCarson, 1999; Wang et al., 2012), the optimal cyclic tensilestimulation parameters have not been defined. It is generallyassumed that mimicking the dynamic mechanical loadingcreated by daily physical activity will give the optimalloading regime in tendon bioreactor system, but this is notyet proven.

The goals of this study were: (1) to develop and validate abioreactor system; (2) to investigate the effect of differentmechanical stimulation regimes on tendon integrity. Wehypothesized that there is an optimum range of mechanicalstimulation required for tendon homeostasis and that underor over-stimulation would reduce tendon integrity.

Materials and Methods

Bioreactor Design Criteria

To develop a functional bioreactor, several design criteriawere formulated. Firstly, the bioreactor must be able toprocess more than three individual cultures simultaneouslywith transparent covers allowing visual monitoring.Secondly, the extension accuracy should be better than�100mm with the load varying as required to produce theset strain (the minimum expected load was 20N persample). Third, a construct fixation method must preventslippage during stimulation. Moreover, as the bioreactorsystem is operated in an incubator, all of the electrical andmechanical components have to be able to function in awarm, humid environment. Finally, but not least, the partsinvolved with the tissue culture should be autoclavable.

To achieve these design criteria two structural materialswere used, AISI 316 stainless steel and 6063 aluminum. Themain bioreactor components are the culture chamber, tissueclamps and hooks made of stainless steel (which can be easilysterilized) and frame. The frame was made of aluminum toreduce the weight and 2mm thick glass was used fortransparent cover. The actuator was a motor-ball screwtransmission design using a step motor and precision ballscrew to maximize the load output and positional accuracy.A programmable logic control was chosen, rather than directcomputer control, to reduce cost and allow reliableoperation in uncertain experimental environments.

Tissue Harvest and Distribution

Achilles tendons were dissected from the hindlimbs offemale New Zealand White rabbits (Oryctolagus cuniculus)

1496 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

241

aged between 14 and 15 weeks with body weight 3–5kg. Thetissue was briefly rinsed in 70% ethanol, washed in 1�phosphate-buffered saline (PBS) three times, and submergedin serum free Dulbecco’s modified Eagle’s medium (DMEM)/F-12 (GIBCO, Invitrogen, Auckland, New Zealand) supple-mented with 50mg/mL Gentamicin before further processing.

In total 24 Achilles tendons from 12 rabbits were allocatedto three groups (Fig. 1A): a loading deprivation group (8), adynamic culture (cyclic tensile strain) group (12), and acontrol group (4). In the loading deprivation group, theAchilles tendons were cultured ex vivo in growth medium(DMEM/F-12, 10% fetal bovine serum and 50mg/mLGentamicin) without mechanical loading for either 6 days(four tendons) or 2 weeks (four tendons). In the dynamicstrain group, the tendons were cultured under 3%, 6%, or 9%cyclic tensile strain for 6 days (8 h per day, 0.25Hz, Fig. 1A).The stimulating cycle was as demonstrated in Figure 1B. Thisregimen was based on previously published studies and dataon rabbit activity levels (West et al., 2004). The samples werethen fixed in 4% paraformaldehyde overnight for histologicalanalysis or directly subjected to RNA extraction. The controlgroup was made up of four tendons directly dissected fromthe animal and immediately fixed.

Histological Preparation and Assessment

For histological assessment, the tissue samples were fixed in4% paraformaldehyde overnight, dehydrated throughgraded alcohol, cleaned with xylene and embedded inparaffin blocks. Histological sections (5mm) were cut on amicrotome and stained with Haematoxylin & Eosin. Blindgeneral histological scoring was performed by three

individuals as previously described (Chen et al., 2011).Four elements (cell density, cell roundness, fiber structure,and fiber arrangement) were assessed using scoring scale of0–3, where 0 is normal and 3 severely abnormal. The averagescore of each parameter was used for comparison.

TUNEL Assay

The terminal deoxynucleotidyl transferase-mediated deox-yuridine triphosphate nick end labeling (TUNEL) assay(Cell Death Detection-AP Kit, Roche, Switzerland) wasperformed to identify apoptotic cells by labeling the nuclearDNA fragments. Paraffin-embedded sections were dewaxedand rehydrated using graded ethanol (100%, 95%, 80%, and70%). The tissue was digested with 20mg/mL proteinase K(Qiagen, Hilden, Germany) for 15min, incubated withTUNEL reaction mixture for 60min and the convert-AP(alkaline phosphatase) reagent for another 30min, accord-ing to the manufacturer’s instruction. After rinsing withPBS, the sections were visualized with Diaminobenzidine(DAB) kit (DAKO, Glostrup, Denmark) for 5min, whichstained the positive cells brown; and counterstained withhematoxylin.

Immunostaining for Type III Collagen

To evaluate the extracellular matrix (ECM) turnover intendons with and without dynamic tensile strain, proteinexpression of type III collagen was evaluated. Paraffin-embedded sections were dewaxed and rehydrated. Theendogenous peroxidase was blocked with 3% hydrogenperoxide (H2O2) for 10min, and the sections were then

Figure 1. A: Number and distribution of samples for each of the 12 rabbits. B: Daily mechanical stimulation pattern used in this study. C: Mechanical input signal from the

control system.

Wang et al.: Programmable Mechanical Stimulation 1497

Biotechnology and Bioengineering

242

incubated with 5% fetal bovine serum in 1� PBS for 30min,followed by monoclonal anti-collagen type III (036K4782,Sigma, St Louis, MO; 1:100) in a damp chamber overnight at48C. Biotinylated goat antimouse IgG at a dilution of 1:200(Sigma) was used as secondary type III antibody. PBSwas usedto wash the sections between solutions and reagents. Sectionswere developed to visualize the positive staining using theDAB kit, and counterstained with hematoxylin solution.

Quantitative Real-Time Polymerase ChainReaction (Q-PCR)

The effect of mechanical stimulation on collagen geneexpression was examined in Achilles tendon after theprogrammed dynamic training. Approximately 200mg oftendon tissue was dissected from the middle region, and thetotal ribonucleic acid (RNA) was extracted using TRIzol1

reagent (Invitrogen, Carlsbad, CA). The RNA concentrationwas quantified and using SmartSpecTM 3000 (Bio-Rad,Hercules, CA), and A260/A280 was measured to evaluate theRNA quality. The RNA (500ng) was then reverse-transcribedinto complementary deoxyribonucleic acid (cDNA) usingSuperScript1 III First-Strand Synthesis System (Invitrogen,USA) according to the manufacturer’s instructions.

The cDNA was amplified and quantified by Q-PCR. Forthis reaction, 5mL of iQTM SYBR1 Green Supermix (Bio-Rad), 1mL of cDNA, 0.5mL of each forward and reverseprimer and 3mL of DEPC water were combined and mixed.The real-time PCR was performed as previously described(Lin et al., 2008). The primer sequences are listed in Table I.

Statistical Analyses

All statistics have been analyzed using the prism 5.03 softwarepackage (GraphPad Software, Inc., San Diego, CA). Data arepresented as mean� standard deviation (SD). Statisticalanalysis was performed by one-way analysis of variance(ANOVA). P values less than 0.05 were considered significant.

Results

Bioreactor for Tendon Engineering

By constructing a bioreactor to the design criteria, we havegenerated a bioreactor that is able to simultaneously apply

pre-programmed uniaxial stimulation on a maximum of sixtendons, cultured separately in their own chamber (toprevent cross contamination) (Fig. 2A). As Achilles tendonsrequire only tensile stimulation (Louis-Ugbo et al., 2004),system performance was tested with rabbit Achilles tendons.The system was actuated by a step motor-ball screwtransmission system (SMBS) using VEXTA step motor(PX245M-01AA, Oriental Motor, Kabushiki Kaisha, Tokyo,Japan) and precision ball screw (SG0601, KSS Co., Ltd.,Ojiya, Niigata, Japan), which produced 198N peak force anda positional accuracy of �5mm. Using programmable logiccontrollers (MELSEC FX1S, Mitsubishi, Marunouchi,Chiyoda-ku, Tokyo, Japan), as shown as Figure 2A,displacement, frequency, and stimulation period are allprogrammable. The culture chamber contains six indepen-dent vessels with individual capacity of 20mL (width¼2 cm, length¼ 9 cm). Individual transparent glass coversminimized the risk of cross contamination between testedsamples, and provided a window for monitoring eachtendon (Fig. 3). This system provides two samplestabilization methods: tissue clamps (Fig. 2C) and hooks(Fig. 2D). Soft and flat scaffolds, like a collagen and silkconstruct, can be fixed using the tissue clamps, whereasrobust and slippery tissue, such as bone/tendon and muscle/tendon junctions, can be tied to the hooks using suture. Thewhole system was placed inside a water-jacketed incubator(Forma Scientific, Marietta, OH) at 378C with 5% CO2, and95% air.

Using the final bioreactor design, Achilles tendonsisolated from rabbits were placed into the system tooptimize of condition of dynamic culture. Cyclic tensilestrain at 3% 6%, and 9% (0.25Hz for 8 h per day for 6 days)were applied, to a total of 12 rabbit Achilles tendons. Overallculture period was over 3 weeks with no contaminationfound during the culture period. Clear vision was achievedthrough transparent glass cover. Both the tissue clamp andtissue hook worked effectively with no slippage observed.

Histological Examination of Tendon UndergoingDifferent Cyclic Tensile Strains

Compared with the native tendon (Fig. 4A), the collagenfiber arrangement changed from parallel to moderately wavyafter 6 days in the absence of applied loading. Cell densitywas increased with slightly rounded nuclei. After 2 weekswithout loading, progressive ECM disruption was observed.

Table I. Rabbit specific primer sequences for the genes used.

Target genes Forward (50–30) Reverse (30–50)

COL1A1 TTGCCCTTCCTTGATATTGC CCTCTTTCGCCCACAATTTA

COL3A1 AAGCCCCAGCAGAAAATTG TGGTGGAACAGCAAAAATCA

MMP-1 GGCTCAGTTCGTCCTCACTC CAGGTCCATCAAAGGGAGAA

MMP-3 GCTTTGCTCAGCCTATCCAC ACCTCCAAGCCAAGGAACTT

MMP-12 ATGCCAGGGAGACCAGTATG AAAGCATGGGCTATGACACC

36B4 ACCCAAAATGCTTCATCGTG CAGGGTTGTTTTCCAGATGC

1498 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

243

Collagen fiber arrangement had no discernible pattern.Fragmentation of collagen fibers and rounded tenocytenuclei were observed microscopically (Fig. 4C). Histologicalassessment showed a time dependent increasing grading infiber arrangement, fiber structure and cell roundness,indicating progressive divergence from the native (control)tissue. Cell density was also increased after 6 days withoutloading (Fig. 4D–G).

In the dynamic culture (cyclic tensile strain) group,histology revealed that the 3% strain stimulated group hadslightly disrupted ECM structure and significant cellularmorphology changes (Fig. 5B). These tendon samplesgraded 1.5 in fiber arrangement, 1.1 in fiber structure and1.3 in cell roundness (Fig. 5E–G). However, in the tendonsthat underwent 6% strain, the ECM orientation and cellular

morphology appeared normal (Fig. 5A and C). That is, therewas no statistical difference of the histology scoring betweenthe 6% group and native tissue (Fig. 5E–H). In the 9% strainstimulated group, the tendon samples were found to bepartially torn and graded 1.45 in fiber structure and 0.8 infiber arrangement (Fig. 5E and F). Increased cell numberand slightly rounded tenocyte nuclei were observed aroundthe rupture site (Fig. 5D, G, and H).

Impact of Cyclic Tensile Strain on Apoptosis of Tenocytes

To investigate the impact of mechanical force on theapoptosis of tenocytes, we used the TUNEL assay to examinethe cellular apoptosis rate in specimens. In the nativetendons, the average cell apoptosis fraction (as a proportionof total cell number) was less than 10% (Fig. 6A and G). Theapoptosis fraction progressively increased over time whenthe tissues were cultured without loading (35% in 6 days and95% in 2 weeks) (Fig. 6G). Cell viability in 3% and 6%stimulated Achilles tendon was similar to that in nativetendon (Fig. 6D and E) without statistical significance(Fig. 6G), whilst cell apoptosis rate increased to 45% whenthe stimulation strain reach 9% (Fig. 6F and G).

Type III Collagen Turnover in Achilles Tendon UnderDynamic Culture

To investigate the ECM turnover in tendon undergoingdifferent cyclic tensile strain, type III collagen immunohis-tochemistry was done. No positive type III collagen wasobserved in the core of healthy tendon using

Figure 2. A: The over view of the bioreactor system. B: Schematic illustration of the bioreactor system. C: Exploded view of the tissue clamps assembly. D: Exploded view of

the tissue hooks assembly.

Figure 3. Bioreactor chamber with three rabbit Achilles’ tendons cultured

inside.

Wang et al.: Programmable Mechanical Stimulation 1499

Biotechnology and Bioengineering

244

Figure 4. H&E staining assessment and histological scoring of loading deprivation culture group. A: Native Achilles tendon structure (day 0). B: Achilles tendon histology after

6 days loading deprivation culture with minor wavy collagen structure. C: Achilles tendon histology after 2 weeks culture displays disorganized collagen structure with round

tenocytes. D: Fiber arrangement, scale of 0–3, 0 represents compacted and parallel, and 3 represents no identifiable pattern. E: Fiber structure, scale of 0–3, 0 represents continue,

long fiber, and 3 represents severe fragmented. F: Cell roundness, scale of 0–3, 0 represents long spindle shape, and 3 represents severely rounding. G: Cell density, scale of 0–3, 0

represents normal pattern, and 3 represents severely increase Result are expressed as the mean� SD (�P< 0.05, ��P< 0.01, ���P< 0.001). Scale bar¼ 50mm.

1500 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

245

Figure 5. H&E staining assessment and histological scoring of 6 days dynamic culture group. A: Native Achilles tendon structure. B: Three percent stimulated group with

wavy collagen structure and minor rounded tenocytes. C: Six percent stimulated group with similar morphology compared with the native group. D: Nine percent stimulated group

with severe tear collagen structure, and rounded nuclei around the tear region (arrow). E: Fiber arrangement, scale of 0–3, 0 represents compacted and parallel, and 3 represents no

identifiable pattern. F: Fiber structure, scale of 0–3, 0 represents continue, long fiber, and 3 represents severely fragmented. G: Cell roundness, scale of 0–3, 0 represents long

spindle shape, and 3 represents severely rounding. H: Cell density, scale of 0–3, 0 represents normal pattern, and 3 represents severely increase. Result is expressed as the

mean�SD (�P< 0.05, ��P< 0.01, ���P< 0.001). Scale bar¼ 50mm.

Wang et al.: Programmable Mechanical Stimulation 1501

Biotechnology and Bioengineering

246

Figure 6. TUNEL assay. A: Native Achilles tendon with less than 10% cell apoptosis. B: Six days loading deprivation group has increased apoptotic tenocytes pointed out by

arrows. C: In 2 weeks loading deprivation group, almost all tenocytes are under apoptosis D: Three percent stimulated group with minor increased apoptotic cells (arrow). E: Six

percent stimulated group with similar cell apoptosis rate as native sample. F: Nine percent stimulated group with massive cell apoptosis around the damaged site (pointed by

arrows). G: Quantified cell apoptosis fraction of each group. Result are expressed as the mean�SD (�P< 0.05, ��P< 0.01, ���P< 0.001). Scale bar¼ 50mm.

1502 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

247

immunochemistry (Fig. 7A). After 6 days of either noloading or 3% strain, the Achilles tendons expressed onlyminor and moderate collagen type III respectively(Fig. 7B and C), whilst relatively higher positive stainingwere found in the 9% strain group (Fig. 7E). No type IIIcollagen expression was observed in the 6% cyclic strainstimulated group (Fig. 7D), which was similar to the nativetendon.

Impact of Cyclic Tensile Strain on ECM RemodelingGene Expression

Q-PCR was carried out to examine the gene expression oftype I (COL1A1), type III collagen (COL3A1), and MMP-1,3, 12 (matrix metalloproteinases). As shown in Figure 8, thetissues subjected to 6% cyclic tensile strain had the highestexpression of COL1A1 (Fig. 8A), whereas 3% stimulatedgroup had the lowest expression. Type III collagen(COL3A1) was suppressed by minor and moderate loading(3% and 6%), yet upregulated in the 9% strain stimulatedgroup (Fig. 8B). Gene expression of MMP1, 3, and 12 werehighly upregulated by 3% strain stimulation compared tothe other groups.

Discussion

A previous study has suggested that engineered neo-tendonsin a non-loading environment result in loose, disorganizedand less collagen-rich matrices (Jiang et al., 2011). Staticmechanical loading has been shown to improve the fiberarrangement of engineered tendon constructs, but did notproduce a compacted tissue structure in vitro (Chen et al.,2007). Engineered neo-tendons with PMS had a well-organized fiber arrangement and structure and enhancedtensile strength, suggesting that dynamic mechanicalloading is preferable for engineered tendon formation(Jiang et al., 2011). Although PMS has been recognized as acritical regulator for tendon formation (Brown and Carson,1999), the optimal PMS parameters are yet to be defined.

In this study we designed a uni-axial bioreactor systemwith a programmable stimulation regimen. Achilles tendonsfrom rabbits were used to evaluate the impact of differentPMS conditions on tendon tissue. Our result showed that3% cyclic tensile strain in the bioreactor did not preventmatrix deterioration, whilst at 6% cyclic tensile strainstructural integrity and cellular function was maintained. At9% cyclic tensile strain there was massive rupture of thecollagen bundles. However, these results are for an already

Figure 7. Collagen type III immunohistochemistry. A: Native Achilles tendon with minor collagen type III expression. B: Loading deprivation group has minor increased

collagen type III expression and some positive staining tenocytes pointed out by arrow. C: Three percent stimulated group with increased collagen type III expression. D: Six

percent stimulated group with similar collagen type III expression as native sample. E: Nine percent stimulated group with massive increased collagen type III expression and

positive staining tenocytes (pointed by arrows). Scale bars¼ 50mm.

Wang et al.: Programmable Mechanical Stimulation 1503

Biotechnology and Bioengineering

248

mature tendon and the loading regimes may be different forengineered constructs at various stages of growth.Furthermore, the frequency of loading may also have animportant role in maintaining tendon homeostasis and thisneed to be better qualified by in vivo measurements ofAchilles tendon loading. Nevertheless, the current resultsdemonstrate that PMS can be achieved in this bioreactor,and that an optimal cyclic tensile loading (around 6% cyclictensile strain) is required to maintain healthy tendonhomeostasis, at least over the time period of the culturesused in this study.

In recent years, many studies of tendon engineering havebeen conducted using custom-made bioreactor system, andthe results have been promising (Abousleiman et al., 2009;Androjna et al., 2007; Chen et al., 2010; Doroski et al., 2010;Issa et al., 2011; Juncosa-Melvin et al., 2006; Nguyen et al.,2009; Nirmalanandhan et al., 2008a,b; Saber et al., 2010;Webb et al., 2006; Woon et al., 2011). Several commercialtendon engineering bioreactors are available now, such asBose1 ElectroForce1 BioDynamic1 and LigaGen system.To develop a functional bioreactor system, several factors

needed to be considered, including material, mechanicaldesign, tissue fixation method, and operating environment(Wang et al., 2012). Compared with commercial bioreactorsin Bose and LigaGen, the current design has severaladvantages. Firstly, our device can condition multiplesamples simultaneously but separately. It can be easilycleaned and sterilized in standard autoclaves. Beingconstructed from AISI 316 stainless steel it can go throughmany sterilization cycles without corrosion or degradation.The device can produce a maximum force of 198N, which issufficient for most tendon scaffold and tendon tissue fromsmall scale animals such as are commonly used in tissueengineering research. Most commercial bioreactor systemshave problems in holding thick and slippery tissue duringPMS, where as our novel tissue hook design demonstratedexcellent tissue fixation. Finally, the actuation system isdesigned to operate inside a standard incubator. It is able tofunction in a humid environment and the horizontal designis easy to move and will only occupy a small space in theincubator. As the tendon’s main function is mechanical, thisbioreactor system is able to provide high-accuracy PMS,

Figure 8. Quantitative PCR results. A: COL1A1 expression in 6% stimulated group is significantly higher than other groups. B: COL3A1 expression in 9% stimulated group is

significantly higher than other groups. Three percent and 6% group had much lower COL3A1 expression compared to the other groups. C–E: Three percent stimulated group has

significantly high expression on MMP1 MMP3 and MMP12, respectively. Result is expressed as the mean�SD (�P< 0.05, ��P< 0.01, ���P< 0.001). Scale bar¼ 50mm.

1504 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

249

enabling different loading regimes including displacement,frequency and stimulation period. Once programmed it isable to operate reliably for weeks without computer support.

It has been well known that tensile mechanical loading isessential for tendon integrity, and loading deprivation cancause tendinopathy-like morphology such as disorientatedcollagen fibers and rounded cell nuclei (Hannafin et al.,1995). Yet the specific effect of different loading conditionson tendon has not been well elucidated. To study the effectof PMS on tendon/ligament, both in vitro and in vivomodels have been used. The Flexcell1 system has beenapplied to examine the in vitro gene expression profiles oftenocytes under tensile or shear force (Skutek et al., 2001;Wang et al., 2003; Yang et al., 2005). These systems canprecisely control the mechanical regimen and monitor thecell behaviors, but only use a monolayer of cells, not anorganized tissue. Cells isolated from their extracellular nicheand cultivated in a monolayer may not necessarily replicatetheir original behavior in vivo. In vivo models, such asAchilles tendon in rats or rabbit, have been commonly usedas study models. Tendon loading is mainly induced byelectrical stimulation on muscle (Asundi et al., 2008;Nakama et al., 2005), treadmill training on animal andpartial tendon transaction (Smith et al., 2008). However,mechanical stimulation on the tendon varies betweenindividuals, making precise quantification of load next toimpossible. In this study, we have demonstrated a bioreactorsystem that not only can be used as an device for tendon/ligament engineering, but also serves as an ex vivo culturemodel for the study of tendon biology. Although thebiomechanics of Achilles tendon is more complicated underin vivo environment (West et al., 2004), by applying PMS,the effect of different cyclic tensile strain on tendon tissuecould be systematically analyzed on a tissue level.

The present study has shown for the first time that only anarrow range of PMS can stimulate the anabolic effect oftendon tissue. This suggests a proposed model describingthe effect of mechanical loading, from under-load to over-load, on tendon homeostasis. This model is illustrated inFigure 9 in which the black curve presents the typicalmechanical response caused by increasing tensile strain intendon. The breakage of collagen fibers increases withincreasing strain, leading to eventual rupture at the failurepoint. The red curve, on the other hand, proposes thebiological repair response of the tendon induced by tensilestrain. Mechanical loading is required for the biologicalactivity of tenocytes, yet the regime needs to be within acertain range to promote matrix product and net tissuerepair. By combining these two curves, three crossover areasare seen, which can be divided according to their metabolicstatus into: A (anabolic) zone and C (catabolic) zones. In theA zone, where the tensile loading is within a certain range,the biological repair response caused by PMS exceeds theinduced damage. In the C zone, there is either too little loadto stimulate matrix production or too much mechanicaldamage caused by excess tensile loading. In our study the 3%strain stimulated group (under-loaded) was located in C

zone. When the tendons were subjected to insufficienttensile strain, early tendinopathologic changes (includingincreased cellular number, type III collagen expression anddisorientated collagen fibers) could be observed. Thesemorphological changes may be attributed by the upregula-tion of MMPs expression in tendon under 3% tensileloading. MMPs are zinc-dependent endopeptidases that ableto degrade all the components of the ECM. MMP-1 isprimarily responsible for the cleavage of collagen, includingtype I, II, and III (Krane et al., 1996). MMP-3 and MMP-12are other members of the MMP family that have a widerange of substrate specificities. Moreover, MMP-12 is able toactivate MMP-3, and MMP-3 has the ability to activateother MMPs such as MMP-1, MMP-7, and MMP-9 (Chen,2004; Knauper et al., 1996). Studies have shown thatdecreased mechanical force on tendon fiber would lead tothe deprivation of cytoskeletal tension in tenocytes and alterthe mechanostransduction in the cells (Arnoczky et al.,2008). Insufficient loading of tenocyte can initiate thedegenerative cascade by expressing a pattern of catabolicgene, including MMPs (Arnoczky et al., 2004, 2008;Egerbacher et al., 2008; Lavagnino et al., 2003, 2006). Inthis study, we have confirmed that under the underloadingcondition, the phenotype of the tenocytes changed fromspindle to round, and the expression levels of catabolicgenes, such as MMP-1, 2, and 12, were upregulated. Theelevated catabolic factors may contribute to the furthermicro-damage of the collagen fiber. Conversely, althoughthe tendons in the 6% strain stimulated group weresubjected to more mechanical damage, the repair mecha-nism was able to maintain tissue homeostasis (i.e., there is anarrow ‘‘net anabolic window’’ between two net cataboliczones). In the 9% group (overloaded), mechanical damageoverwhelmed the tendon tissue repair and caused partialrupture, which is consistent with our proposed model.

The identification of an optimal PMS range for tendonwill not only guide us to a better tissue-engineering pathway

Figure 9. Matrix damage (black curve) and production (red broken) affected by

mechanical stimulation. A zone indicates a small range of anabolic effect caused

by mechanical stimulation. Matrix production stimulated by mechanical loading

overcomes the matrix damage that the tendon is able to maintain its structural

integrity. C zone indicates two catabolic zones that matrix damage caused by

mechanical stretching overcomes the matrix production result in tendinopathy.

Wang et al.: Programmable Mechanical Stimulation 1505

Biotechnology and Bioengineering

250

for tendon regeneration, but also provide an insight intotendon homeostasis under both physiological and patho-logical conditions. Clinically, chronic tendinopathy ofAchilles tendon is usually conservatively managed byeccentric overload training. The calf muscle is eccentricallyloaded with the knee held straight or the soleus muscle ismaximally loaded with the knee bent (Alfredson et al.,1998). Although positive impacts of eccentric overloadtraining on tendinopathy treatment have been recorded, theoutcomes have varied widely (Alfredson and Lorentzon,2003; Alfredson et al., 1998; Fahlstrom et al., 2003; Rooset al., 2004; Shalabi et al., 2004; Silbernagel et al., 2001;Stanish et al., 1986). In the study of Alfredson et al. (1998)the overall improvement in pain score was up to 94%,though in the study of Silbernagel et al. reported only 29%(Silbernagel et al., 2001). Although the mechanism ofeccentric overload training on tendinopathy managementremains unclear, the present study has provided a possibleexplanation of the mechanism of eccentric loading onchronic tendinopathy. The high variation of clinicaloutcomes might be due to poor control of the loadingregimen. Given that there is only a narrow range of tensilestrain where eccentric exercise could provide an anaboliceffect on tendon, clinical treatment protocols will need toaccurately control the magnitude of load applied throughthe injured tendon.

Finally, it is noted that nutrient supply might be altered byadvective transport associated with mechanical stimulation.Although a previous study on canine flexor tendonssuggested that mechanical stimulation did not alter smallnutrient uptake compared to static culture, and the tendonrepair improvements derived from passive motion is notrelated to an increase of nutrient transport (Hannafin andArnoczky, 1994), under different loading regimes, or indifferent tendon types, this may not be the case. Moreover,the transport of large molecules, such as growth factors andmatrix components, may be affected by mechanicalstimulation, as has been shown in studies in related tissues(Bonassar et al., 2000; Zhang et al., 2007). Therefore, thetendon stimulated by dynamic mechanical loading in thisstudy might be attributed to the enhanced nutrient transport,compared to the static control. Consequently it is difficult toisolate the direct effect of mechanical stimulation on tendonresponse from potential enhanced advective transportprocesses. Further study would be required to measure thenutrition transport in tendon under PMS.

Conclusion

We have developed and validated a uniaxial bioreactorsystem. The results in our study showed tendons subjectedto 6% of cyclic tensile strain were able to maintain theirstructural integrity and cellular function. Our data suggestedthat tendon is a mechano-sensing tissue where the netanabolic action of cyclic tensile strain only occurs over anarrow range. In our study, mechanical underloading (3%

cyclic tensile strain) and overloading (9% cyclic tensilestrain) produced a net negative impact on the rabbit Achillestendon tissue.

References

Abousleiman R.I., Reyes Y, McFetridge P, Sikavitsas V. 2009. Tendon tissue

engineering using cell-seeded umbilical veins cultured in a mechanical

stimulator. Tissue Eng Part A 15(4):787–795.

Alfredson H, Lorentzon R. 2003. Intratendinous glutamate levels and

eccentric training in chronic Achilles tendinosis: A prospective study

using microdialysis technique. Knee Surg Sports Traumatol Arthrosc

11(3):196–199.

Alfredson H, Pietila T, Jonsson P, Lorentzon R. 1998. Heavy-load eccentric

calf muscle training for the treatment of chronic Achilles tendinosis.

Am J Sports Med 26(3):360–366.

Altman GH, Lu HH, Horan RL, Calabro T, Ryder D, Kaplan DL, Stark P,

Martin I, Richmond JC, Vunjak-Novakovic G. 2002. Advanced biore-

actor with controlled application of multi-dimensional strain for tissue

engineering. J Biomech Eng 124(6):742–749.

Androjna C, Spragg RK, Derwin KA. 2007. Mechanical conditioning of cell-

seeded small intestine submucosa: A potential tissue-engineering strat-

egy for tendon repair. Tissue Eng 13(2):233–243.

Arnoczky SP, Lavagnino M, Egerbacher M, Caballero O, Gardner K,

Shender MA. 2008. Loss of homeostatic strain alters mechanostat

‘‘set point’’ of tendon cells in vitro. Clin Orthop Relat Res

466(7):1583–1591.

Arnoczky SP, Tian T, Lavagnino M, Gardner K. 2004. Ex vivo static tensile

loading inhibits MMP-1 expression in rat tail tendon cells through a

cytoskeletally based mechanotransduction mechanism. J Orthop Res

22(2):328–333.

Asundi KR, King KB, Rempel DM. 2008. Evaluation of gene expression

through qRT-PCR in cyclically loaded tendons: An in vivo model. Eur J

Appl Physiol 102(3):265–270.

Bonassar LJ, Grodzinsky AJ, Srinivasan A, Davila SG, Trippel SB. 2000.

Mechanical and physicochemical regulation of the action of insulin-like

growth factor-I on articular cartilage. Arch Biochem Biophys

379(1):57–63.

Brown CH, Jr., Carson EW. 1999. Revision anterior cruciate ligament

surgery. Clin Sports Med 18(1):109–107.

Calve S, Lytle IF, Grosh K, Brown DL, Arruda EM. 2010. Implantation

increases tensile strength and collagen content of self-assembled tendon

constructs. J Appl Physiol 108(4):875–881.

Catapano G, Patzer JF 2nd, Gerlach JC. 2010. Transport advances in

disposable bioreactors for liver tissue engineering. Adv Biochem Eng

Biotechnol 115:117–143.

Chen J, Yu Q,WuB, Lin Z, Pavlos NJ, Xu J, OuyangH,Wang A, ZhengMH.

2011. Autologous tenocyte therapy for experimental Achilles tendino-

pathy in a rabbit model. Tissue Eng Part A 17(15–16):2037–2048.

Chen JL, Yin Z, Shen WL, Chen X, Heng BC, Zou XH, Ouyang HW. 2010.

Efficacy of hESC-MSCs in knitted silk-collagen scaffold for tendon

tissue engineering and their roles. Biomaterials 31(36):9438–9451.

Chen JM, Willers C, Xu J, Wang A, Zheng MH. 2007. Autologous tenocyte

therapy using porcine-derived bioscaffolds for massive rotator cuff

defect in rabbits. Tissue Eng 13(7):1479–1491.

Chen X, Qi YY, Wang LL, Yin Z, Yin GL, Zou XH, Ouyang HW. 2008.

Ligament regeneration using a knitted silk scaffold combined with

collagen matrix. Biomaterials 29(27):3683–3692.

Chen YE. 2004. MMP-12, an old enzyme plays a new role in the pathogen-

esis of rheumatoid arthritis? Am J Pathol 165(4):1069–1070.

Dennis RG, Smith B, Philp A, Donnelly K, Baar K. 2009. Bioreactors for

guiding muscle tissue growth and development. Adv Biochem Eng

Biotechnol 112:39–79.

Doroski DM, Levenston ME, Temenoff JS. 2010. Cyclic tensile culture

promotes fibroblastic differentiation of marrow stromal cells encapsu-

lated in poly(ethylene glycol)-based hydrogels. Tissue Eng Part A

16(11):3457–3466.

1506 Biotechnology and Bioengineering, Vol. 110, No. 5, May, 2013

251

Egerbacher M, Arnoczky SP, Caballero O, Lavagnino M, Gardner KL. 2008.

Loss of homeostatic tension induces apoptosis in tendon cells: An in

vitro study. Clin Orthop Relat Res 466(7):1562–1568.

FahlstromM, Jonsson P, Lorentzon R, Alfredson H. 2003. Chronic Achilles

tendon pain treated with eccentric calf-muscle training. Knee Surg

Sports Traumatol Arthrosc 11(5):327–333.

Hannafin JA, Arnoczky SP. 1994. Effect of cyclic and static tensile loading on

water content and solute diffusion in canine flexor tendons: An in vitro

study. J Orthop Res 12(3):350–356.

Hannafin JA, Arnoczky SP, Hoonjan A, Torzilli PA. 1995. Effect of stress

deprivation and cyclic tensile loading on the material and morphologic

properties of canine flexor digitorum profundus tendon: An in vitro

study. J Orthop Res 13(6):907–914.

Issa RI, Engebretson B, Rustom L, McFetridge PS, Sikavitsas VI. 2011. The

effect of cell seeding density on the cellular and mechanical properties

of a mechanostimulated tissue-engineered tendon. Tissue Eng Part A

17(11–12):1479–1487.

Jiang Y, Liu H, Li H,Wang F, Cheng K, Zhou G, ZhangW, YeM, Cao Y, Liu

W, Zou H. 2011. A proteomic analysis of engineered tendon formation

under dynamic mechanical loading in vitro. Biomaterials 32(17):4085–

4095.

Juncosa-Melvin N, Shearn JT, Boivin GP, Gooch C, Galloway MT, West JR,

Nirmalanandhan VS, Bradica G, Butler DL. 2006. Effects of mechanical

stimulation on the biomechanics and histology of stem cell-collagen

sponge constructs for rabbit patellar tendon repair. Tissue Eng

12(8):2291–2300.

Knauper V, Lopez-Otin C, Smith B, Knight G, Murphy G. 1996. Biochemical

characterization of human collagenase-3. J Biol Chem 271(3):1544–1550.

Krane SM, Byrne MH, Lemaitre V, Henriet P, Jeffrey JJ, Witter JP, Liu X,

Wu H, Jaenisch R, Eeckhout Y. 1996. Different collagenase gene

products have different roles in degradation of type I collagen. J

Biol Chem 271(45):28509–28515.

Lavagnino M, Arnoczky SP, Egerbacher M, Gardner KL, Burns ME. 2006.

Isolated fibrillar damage in tendons stimulates local collagenase mRNA

expression and protein synthesis. J Biomech 39(13):2355–2362.

LavagninoM, Arnoczky SP, Tian T, Vaupel Z. 2003. Effect of amplitude and

frequency of cyclic tensile strain on the inhibition of MMP-1 mRNA

expression in tendon cells: An in vitro study. Connect Tissue Res

44(3–4):181–187.

Lea RB, Smith L. 1972. Non-surgical treatment of tendo achillis rupture. J

Bone Joint Surg Am 54(7):1398–1407.

Leppilahti J, Orava S. 1998. Total Achilles tendon rupture. A review. Sports

Med 25(2):79–100.

Li G, Defrate LE, Rubash HE, Gill TJ. 2005. In vivo kinematics of the

ACL during weight-bearing knee flexion. J Orthop Res 23(2):

340–344.

Lin Z, Fitzgerald JB, Xu J, Willers C, Wood D, Grodzinsky AJ, Zheng MH.

2008. Gene expression profiles of human chondrocytes during passaged

monolayer cultivation. J Orthop Res 26(9):1230–1237.

Louis-Ugbo J, Leeson B, Hutton WC. 2004. Tensile properties of fresh

human calcaneal (Achilles) tendons. Clin Anat 17(1):30–35.

Nakama LH, King KB, Abrahamsson S, Rempel DM. 2005. Evidence of

tendon microtears due to cyclical loading in an in vivo tendinopathy

model. J Orthop Res 23(5):1199–1205.

Nguyen TD, Liang R, Woo SL, Burton SD, Wu C, Almarza A, Sacks MS,

Abramowitch S. 2009. Effects of cell seeding and cyclic stretch on the

fiber remodeling in an extracellular matrix-derived bioscaffold. Tissue

Eng Part A 15(4):957–963.

Nirmalanandhan VS, Rao M, Shearn JT, Juncosa-Melvin N, Gooch C,

Butler DL. 2008a. Effect of scaffold material, construct length and

mechanical stimulation on the in vitro stiffness of the engineered

tendon construct. J Biomech 41(4):822–828.

Nirmalanandhan VS, Shearn JT, Juncosa-Melvin N, Rao M, Gooch C, Jain

A, Bradica G, Butler DL. 2008b. Improving linear stiffness of the cell-

seeded collagen sponge constructs by varying the components of the

mechanical stimulus. Tissue Eng Part A 14(11):1883–1891.

Peach MS, James R, Toti US, Deng M, Morozowich NL, Allcock HR,

Laurencin CT, Kumbar SG. 2012. Polyphosphazene functionalized

polyester fiber matrices for tendon tissue engineering: In vitro evalua-

tion with human mesenchymal stem cells. Biomed Mater 7(4):045016.

Rauh J, Milan F, Gunther KP, Stiehler M. 2011. Bioreactor systems for bone

tissue engineering. Tissue Eng Part B Rev 17(4):263–280.

Rauh J, Milan F, Gunther KP, Stiehler M. 2012. Bioreactor systems for bone

tissue engineering. Tissue Eng Part B Rev 17(4):263–280.

Roos EM, Engstrom M, Lagerquist A, Soderberg B. 2004. Clinical improve-

ment after 6 weeks of eccentric exercise in patients with mid-portion

Achilles tendinopathy—A randomized trial with 1-year follow-up.

Scand J Med Sci Sports 14(5):286–295.

Saber S, Zhang AY, Ki SH, Lindsey DP, Smith RL, Riboh J, Pham H, Chang

J. 2010. Flexor tendon tissue engineering: Bioreactor cyclic strain

increases construct strength. Tissue Eng Part A 16(6):2085–2090.

Sahoo S, Ouyang H, Goh JC, Tay TE, Toh SL. 2006. Characterization of a

novel polymeric scaffold for potential application in tendon/ligament

tissue engineering. Tissue Eng 12(1):91–99.

Shalabi A, Kristoffersen-Wilberg M, Svensson L, Aspelin P, Movin T. 2004.

Eccentric training of the gastrocnemius-soleus complex in chronic Achilles

tendinopathy results in decreased tendon volume and intratendinous

signal as evaluated by MRI. Am J Sports Med 32(5):1286–1296.

Silbernagel KG, Thomee R, Thomee P, Karlsson J. 2001. Eccentric overload

training for patients with chronic Achilles tendon pain—A randomised

controlled study with reliability testing of the evaluation methods.

Scand J Med Sci Sports 11(4):197–206.

Skutek M, van Griensven M, Zeichen J, Brauer N, Bosch U. 2001. Cyclic

mechanical stretching modulates secretion pattern of growth factors in

human tendon fibroblasts. Eur J Appl Physiol 86(1):48–52.

Smith LL. 2000. Cytokine hypothesis of overtraining: A physiological

adaptation to excessive stress? Med Sci Sports Exerc 32(2):317–331.

Smith MM, Sakurai G, Smith SM, Young AA, Melrose J, Stewart CM,

Appleyard RC, Peterson JL, Gillies RM, Dart AJ, Sonnabend DH, Little

CB. 2008. Modulation of aggrecan and ADAMTS expression in ovine

tendinopathy induced by altered strain. Arthritis Rheum 58(4):1055–1066.

StanishWD, Rubinovich RM, Curwin S. 1986. Eccentric exercise in chronic

tendinitis. Clin Orthop Relat Res (208): 65–68.

Stoll C, John T, Endres M, Rosen C, Kaps C, Kohl B, Sittinger M, Ertel W,

Schulze-Tanzil. 2010. Extracellular matrix expression of human teno-

cytes in three-dimensional air-liquid and PLGA cultures compared

with tendon tissue: Implications for tendon tissue engineering. J

Orthop Res 28(9):1170–1177.

Wang JH, Jia F, Yang G, Yang S, Campbell BH, Stone D, Woo SL. 2003. Cyclic

mechanical stretching of human tendon fibroblasts increases the produc-

tion of prostaglandin E2 and levels of cyclooxygenase expression: A novel

in vitro model study. Connect Tissue Res 44(3–4):128–133.

Wang T, Gardner B, Lin Z, Rubenson J, Wang A, Kirk TB, Xu J, Smith D,

Lloyd D, Zheng MH. 2012. Bioreactor design for tendon/ligament

engineering. Tissue Eng Part B Rev. [Epub ahead of print] DOI:

10.1089/ten.teb.2012.0295

Webb K, Hitchcock RW, Smeal RM, LiW, Gray SD, Tresco PA. 2006. Cyclic

strain increases fibroblast proliferation, matrix accumulation, and

elastic modulus of fibroblast-seeded polyurethane constructs. J Bio-

mech 39(6):1136–1144.

West JR, Juncosa N, Galloway MT, Boivin GP, Butler DL. 2004. Characteriza-

tion of in vivo Achilles tendon forces in rabbits during treadmill locomo-

tion at varying speeds and inclinations. J Biomech 37(11):1647–1653.

Woon CY, Farnebo S, Schmitt T, Kraus A, Megerle K, Pham H, Yan X,

Gambhir SS, Chang J. 2012. Human Flexor Tendon Tissue Engineering:

Revitalization of Biostatic Allograft Scaffolds. Tissue Eng Part A

18(23–24):2406–2417.

Woon CY, Kraus A, Raghavan SS, Pridgen B.C., Megerle K, PhamH, Chang J.

2011. Three-dimensional-construct bioreactor conditioning in human

tendon tissue engineering. Tissue Eng Part A 17(19–20):2561–2572.

Yang G, ImHJ,Wang JH. 2005. Repetitive mechanical stretchingmodulates

IL-1beta induced COX-2, MMP-1 expression, and PGE2 production in

human patellar tendon fibroblasts. Gene 363:166–172.

Zhang L, Gardiner BS, Smith DW, Pivonka P, Grodzinsky A. 2007. The

effect of cyclic deformation and solute binding on solute transport in

cartilage. Arch Biochem Biophys 457(1):47–56.

Wang et al.: Programmable Mechanical Stimulation 1507

Biotechnology and Bioengineering

252