Design and Development of Cell Stretching Platforms for ...

154
Design and Development of Cell Stretching Platforms for Mechanobiology Studies Author Kamble, Harshad Published 2017-09 Thesis Type Thesis (PhD Doctorate) School School of Natural Sciences DOI https://doi.org/10.25904/1912/1280 Copyright Statement The author owns the copyright in this thesis, unless stated otherwise. Downloaded from http://hdl.handle.net/10072/370968 Griffith Research Online https://research-repository.griffith.edu.au

Transcript of Design and Development of Cell Stretching Platforms for ...

Page 1: Design and Development of Cell Stretching Platforms for ...

Design and Development of Cell Stretching Platforms forMechanobiology Studies

Author

Kamble, Harshad

Published

2017-09

Thesis Type

Thesis (PhD Doctorate)

School

School of Natural Sciences

DOI

https://doi.org/10.25904/1912/1280

Copyright Statement

The author owns the copyright in this thesis, unless stated otherwise.

Downloaded from

http://hdl.handle.net/10072/370968

Griffith Research Online

https://research-repository.griffith.edu.au

Page 2: Design and Development of Cell Stretching Platforms for ...

Design and Development of Cell Stretching

Platforms for Mechanobiology Studies

Queensland Micro- and Nanotechnology Centre

School of Natural Sciences

Griffith University

Submitted in fulfillment of the requirements of the degree of

Doctor of Philosophy

(Biomedical Engineering)

by

Mr. Harshad Kamble

Bachelor of Engineering (Biomedical Engineering)

Master of Technology (Mechatronics)

September 2017

Page 3: Design and Development of Cell Stretching Platforms for ...

ii

To my shining little star, my loving partner

Twinkle George.

Page 4: Design and Development of Cell Stretching Platforms for ...

iii

Design and Development of Cell Stretching Platforms for Mechanobiology

Studies

by

Harshad Kamble

Submitted to the School of Natural Sciences on September 27th

, 2017, in partial fulfillment of

the requirements for the degree of Doctor of Philosophy in Biomedical Engineering.

ABSTRACT

Cells within the human body are continuously exposed to various mechanical stimuli due to

organ function, movement and growth. Cellular response to such mechanical stimuli is known

as a mechanobiological signalling, which is an integral part of the cell homeostasis. It is

widely accepted that maladaptation of mechanobiological signalling may lead to dysfunction

and/or disease. Thus, better understanding of mechanobiological signalling has become a key

area of interest for researchers in the field of regenerative medicine and tissue engineering.

However, complexity involved in the in vivo biological systems has been a major hurdle for

comprehensive mechanobiological investigations. This technological gap motivates

researchers to develop in vitro devices capable of introducing mechanical strain onto a cell

culture and to closely mimic the in vivo physiological conditions.

For example, various cell stretching approaches have been developed to induce mechanical

strain onto a cell culture and trigger cellular responses such as migration, proliferation and

orientation. However, very few existing cell stretching platforms fulfil the major requirement

of a robust cell stretching tool such as high experimental throughput, well-characterised and

controllable strain pattern, ease of operation, compatible with a wide range of imaging

systems and most importantly high biological relevance for systematic mechanobiological

investigation. Thus, the present thesis focuses on the development of robust cell stretching

platforms based on electromagnet and pneumatic actuations to address these existing

limitations and subsequently to establish a systematic approach for in-depth

mechanobiological investigation.

Page 5: Design and Development of Cell Stretching Platforms for ...

iv

To provide a systematic approach for detailed study, the first necessary step is quantification

of the parameter. Thus, in the Chapter three of this thesis, a novel cell stretching platform

based on a single sided uniaxial stretching approach was developed to apply tensile strain

onto the cell culture and observe cellular response of the cells towards different strains in the

same field of view with lower fabrication and operation complexity. The effectiveness of the

platform was demonstrated by observing the response of cells in culture under different strain

amplitudes. In the Chapter four, a standardised numerical tool was developed for the single-

sided uniaxial cell stretching platform. The numerical tool provided guidelines for the

optimization parameters and paved way for the development of a double-sided cell stretching

platform described in Chapter five. The developed platform was capable of investigating the

cellular behaviour for a wide range of homogenous strain amplitudes with cyclic and static

stretching conditions. Although the developed electromagnetically cell stretching platforms

provided a standardised tool for systematic mechanobiological investigation, the biological

relevance could still be improved. Thus, the Chapter six involved the development of a novel

pneumatically actuated array-based cell stretching platform, which concurrently induced a

range of cyclic strain onto the cell culture. It was developed to achieve cell patterns, which

provided an improved biological relevance for mechanobiological studies. The toroidal

shaped strain pattern was utilised to achieve circumferential cellular alignment of cells similar

to that of in vivo smooth muscle in the vascular wall. Furthermore, the dimensions of the

platform followed those of standard 96 well plates. This simple and effective design approach

ensured a high compatibility with pre-clinical tools and protocols, which is critical for high-

throughput cell-stretching assays.

Collectively, the findings for these chapters and the thesis at large, suggest the high clinical

compatibility and biological relevance of the cell stretching devices reported in this thesis

provide promising platform for systematic mechanobiological investigations.

Principal Supervisor: Prof. Nam-Trung Nguyen

Associate Supervisor: Dr. Matthew Barton

Page 6: Design and Development of Cell Stretching Platforms for ...

v

Page 7: Design and Development of Cell Stretching Platforms for ...

vi

Acknowledgement

I would like to thank and express my gratitude toward my principal supervisor, Prof. Nam-

Trung Nguyen for his invaluable guidance during my PhD candidature. His constant support,

motivation and systematic approach were the key factors which made this dissertation

possible. I was able to keep my research on track because of his expertise and scientific

insights. I would also like thank my associate supervisor, Dr. Matthew Barton for his support,

encouragement and constructive suggestions.

I am grateful to Griffith University and Queensland Micro- and Nanotechnology Centre

(QMNC) for providing me with the resources needed for my research. I would also like to

thank Associate Professor Igor Litvinyuk, Higher Degree Research (HDR) convener, for his

support. I would like to express my gratitude to Mr. Raja Vadivelu for sharing his knowledge

and helping me with biological experimentations. I would also like to acknowledge Griffith

Graduate Research School (GGRS) and QMNC center staff member for helping me with

various administrative matters.

I would like to take this opportunity to thank my beloved partner Twinkle George for her

constant support, motivation and unconditional love. Finally, I also would like to thank to

family and friends for their endless love and encouragement.

Page 8: Design and Development of Cell Stretching Platforms for ...

vii

List of Publications

1. H Kamble, MJ Barton, M Jun, S Park, NT Nguyen, "Cell stretching devices as

research tools: engineering and biological considerations" Lab on a Chip, 2016

2. H Kamble, M Jun, S Park, MJ Barton, R Vadivelu, JS John, NT Nguyen "An

electromagnetic cell-stretching device for mechanotransduction studies of olfactory

ensheathing cells" Biomedical microdevices, 2016

3. H Kamble, MJ Barton, NT Nguyen "Modelling of an uniaxial single-sided

magnetically actuated cell-stretching device."- Sensors and Actuators A: Physical,

2016

4. H Kamble , R Vadivelu , MJ Barton , K Boriachek , A Munaz , S Park , MJA.

Shiddiky , NT Nguyen "An electromagnetically actuated double-sided cell-stretching

device for mechanobiology research" Micromachines, 2017.

5. H Kamble , R Vadivelu , MJ Barton MJA. Shiddiky , NT Nguyen "Pneumatically

actuated array based cell stretching platform to achieve patterned cell culture with

predefined circumferential cellular alignment in vitro" Lab Chip, 2017 (Submitted).

Page 9: Design and Development of Cell Stretching Platforms for ...

viii

Table of Content

Table of Content ................................................................................................................ viii

List of Figures ..................................................................................................................... xii

Chapter 1: Introduction ......................................................................................................... 16

1.1 Background and motivation ....................................................................................... 16

1.2 Research scope and objective ..................................................................................... 17

1.3 Thesis framework ....................................................................................................... 18

1.4 Research questions and specific aims ........................................................................ 19

1.5 General methods and materials. ................................................................................. 23

1.5.1 Finite Element Analysis (FEA) modelling. ........................................................ 23

1.5.2 Fabrication processes .......................................................................................... 23

1.5.3 Image processing for detection and tracing of the deformation ......................... 25

1.5.4 Cell culture .......................................................................................................... 26

1.5.5 Cell fixing, immunofluorescence staining and imaging ..................................... 26

1.5.6 Image processing for the biological analysis ...................................................... 26

1.6 References .................................................................................................................. 27

Chapter 2: Cell stretching devices as research tools: engineering and biological

considerations .......................................................................................................................... 29

2.1 Introduction ................................................................................................................ 31

2.2 Actuation concepts ..................................................................................................... 33

2.2.1 Pneumatic actuation ............................................................................................ 33

2.2.2 Piezoelectric actuation ........................................................................................ 34

2.2.3 Electromagnetic actuation ................................................................................... 36

2.2.4 Optical actuation ................................................................................................. 38

2.2.5 Other actuators .................................................................................................... 38

2.3 Design considerations ................................................................................................ 40

2.3.1 Biocompatibility ................................................................................................. 41

2.3.2 Substrate properties ............................................................................................. 41

2.3.3 Control strategies ................................................................................................ 42

2.3.4 Stretching direction ............................................................................................. 43

2.3.5 Size of the actuation system ................................................................................ 43

2.4 Biological considerations ........................................................................................... 44

Page 10: Design and Development of Cell Stretching Platforms for ...

ix

2.4.1 Mechanobiology ................................................................................................. 44

2.4.2 Mechanotransduction .......................................................................................... 45

2.5. Conclusions and perspective ...................................................................................... 47

2.6 References .................................................................................................................. 50

Chapter 3: An electromagnetic cell-stretching device for mechanotransduction studies of

olfactory ensheathing cells ..................................................................................................... 56

3.1 Introduction ................................................................................................................ 58

3.2 Device concept and design ......................................................................................... 60

3.2.1 Characterisation of the permanent magnet and the electromagnet ..................... 61

3.2.2 Design and fabrication ........................................................................................ 63

3.3 Device characterisation .............................................................................................. 65

3.3.1 Characterisation of magnetic force ..................................................................... 65

3.3.2 Characterisation of stretching membrane ........................................................... 66

3.4 Characterisation of strain map.................................................................................... 66

3.4.1 Numerical analysis .............................................................................................. 67

3.4.2 Experimental analysis ......................................................................................... 68

3.5 Culturing of OECs on the stretching device............................................................... 69

3.5.1 Materials and methods ........................................................................................ 69

3.5.2 Effect of strain on the OEC morphology ............................................................ 71

3.6 Conclusions ................................................................................................................ 72

3.7 References .................................................................................................................. 73

Chapter 4: Modelling of an uniaxial single-sided magnetically actuated cell-stretching

device ........................................................................................................................................ 76

4.1 Introduction ............................................................................................................... 78

4.2 Device Concept .......................................................................................................... 80

4.3 Modelling approach.................................................................................................... 81

4.3.1 Modelling of the electromagnet .......................................................................... 81

4.3.2 Modelling of the magnetic force ....................................................................... ..84

4.3.2.1 Modelling of the permanent magnet ................................................................... 84

Page 11: Design and Development of Cell Stretching Platforms for ...

x

4.3.2.2 Modelling of the superimposed system with electromagnet and permanent

magnet ................................................................................................................. 85

4.3.2.3 Magnetic force calculations ................................................................................ 86

4.3.3 Modelling of the magnetically actuated cell-stretching device .......................... 88

4.4 Conclusion .................................................................................................................. 89

4.5 References .................................................................................................................. 90

Chapter 5. An electromagnetically actuated double-sided cell-stretching device for

mechanobiology research ....................................................................................................... 92

5.1. Introduction ................................................................................................................ 94

5.2. Materials and methods ............................................................................................... 96

5.2.1 Device design and working principle ................................................................. 96

5.2.2 Modelling and fabrication ................................................................................... 97

5.2.3 Cell culture ........................................................................................................ 100

5.2.4 Application of strain on Fibroblasts .................................................................. 100

5.2.5 Cell fixing, immunofluorescence staining and imaging ................................... 101

5.2.6 Image analysis ................................................................................................... 101

5.3 Results and discussion ............................................................................................. 101

5.3.1 Force calculation ............................................................................................... 101

5.3.2 Strain calculation .............................................................................................. 103

5.3.3 Cell area and aspect ratio .................................................................................. 105

5.3.4 Cell orientation.................................................................................................. 107

5.4 Conclusion ................................................................................................................ 110

5.5 References ................................................................................................................ 111

Chapter 6: Pneumatically actuated cell-stretching array platform to engineer cell

pattern in vitro ....................................................................................................................... 115

6.1 Introduction .............................................................................................................. 117

6.2 Materials and methods ............................................................................................. 120

6.2.1 Device design and working principle ............................................................... 120

6.2.2 Design and optimisation ................................................................................... 121

6.2.3 Fabrication process ........................................................................................... 123

6.2.4 Cell culture ........................................................................................................ 124

6.2.5 Cell Fixing, immunofluorescence staining and imaging .................................. 125

Page 12: Design and Development of Cell Stretching Platforms for ...

xi

6.3 Results and discussion .............................................................................................. 125

6.3.1 Device characterisation ..................................................................................... 125

6.3.2 Cell orientation.................................................................................................. 130

6.4 Conclusions .............................................................................................................. 133

6.5 References ................................................................................................................ 135

Chapter 7: Conclusions and perspectives ........................................................................... 139

Appendix A: MATLAB Image processing ......................................................................... 143

A.1 Custom made MATLAB program for particle detection and tracing ...................... 143

A.2 Custom made MATLAB program for curve detection and ellipse fitting ............... 146

A.2 Custom made MATLAB program for ElveFlow pressure controller ...................... 149

Page 13: Design and Development of Cell Stretching Platforms for ...

xii

List of Figures

Chapter 2

Figure 2.1 Commercially available cell stretchers: (a) ElectroForce 3100 (Bose Corporation)

(b) Stage Flexer (Flexcell international Cop.) (c) Stretch System for Microscope STB-150

(STERX Inc.). . .................................................................................................................... 31

Figure 2.2 Typical cell stretching devices with pneumatic actuation: (a) Inflated balloons with

positive pressure; (b) Actuation chamber positive pressure; (c) Two-chamber side actuation

with negative pressure; (d) Four-chamber side actuation with negative pressure; (e) Radial

stretching with circular support and negative pressure. ....................................................... 34

Figure 2.3 Typical cell stretching devices with piezoelectric actuation: (a) Radial strain with

pushing pin; (b) Linear stretching with piezoelectric linear drive; (c) MEMS translation

stage with external actuator; (d) MEMS linkage mechanism. ............................................. 35

Figure 2.4 Typical cell stretching devices with electromagnetic actuation: (a) Linear pushing

with a servo motor; (b) Radial strain with pushing by a stepper motor; (c) Stretching with

linear translation stage; (d) Stretching witch cam follower mechanism. ............................. 36

Figure 2.5 Typical cell stretching devices with optical actuation: (a) Stretching in a

microchannel with free-space laser beam ............................................................................ 37

Figure 2.6 Other actuation schemes for cell stretching: (a) Thermomechanic actuation with

shape memory alloy; (b) Dielectrophoretic actuation; (c, d) Electrostatic actuation; (e)

Hydraulic actuation.: ............................................................................................................ 39

Chapter 3

Figure 3.1 The cell-stretching device: (a) the actual device; (b) the device in the relaxed state;

(c) the device in the stretched state. ..................................................................................... 59

Figure 3.2 Measured magnetic flux density: (a) as function of a distance from the permanent

magnet; (b) as function of the voltage applied on the electromagnet. ................................. 61

Figure 3.3 Simulated results for location of actuation axis along Z axis verses average

displacement of membrane. (Inset: Optimised geometry and simulation results of FEA

model.). ................................................................................................................................ 62

Figure 3.4 Flow chart for a developed algorithm to calculate displacement and strain. .......... 63

Figure 3.5 Experimental characterisation of the device: (a) displacement versus voltage; (b)

estimated force versus voltage.. ........................................................................................... 64

Page 14: Design and Development of Cell Stretching Platforms for ...

xiii

Figure 3.6 Comparison of experimental and simulation force versus displacement results

(Inset: Actual cell stretching system displacement due to actuation) . ................................ 66

Figure 3.7 Numerically simulated strain as function of appliedforce . (Inset: Strain pattern for

membrnae with thickness of 0.2 mm) .................................................................................. 67

Figure 3.8 Comparison of Strain map over membrane: (a) simulation; (b) experiment...........68

Figure 3.9 Comparison of voltage versus simulation and experimental strain results at the

central region of membrane. (Inset: Illustration of experimental setup).. ............................ 69

Figure 3.10 Experimental results: Time verses average length of OEC over membrane. Inset:

Morphology of OECs... ........................................................................................................ 70

Figure 3.11 Experimental results: Strain verses average length of OEC after 4 h and 8 h.

(Inset: Change in Morphology of OEC culture with and without stretching).. .................... 72

Chapter 4

Figure 4.1 Operation concept of the cell-stretching device (top view and side view): (a)

Relaxed state; (b) Actuated state. ......................................................................................... 79

Figure 4.2 The geometry of the electromagnet: (a) 2D Work plane; (b) 3D view. .................. 80

Figure 4.3 Magnetic flux density obtained with coil current of Icoil = 0.183 A: (a) With coarse

surrounding medium mesh (B= 28.6 mT); (b) With optimised mesh of the surrounding (B=

88.2 mT).. ............................................................................................................................. 82

Figure 4.4 Magnetic flux density versus applied current at the surface of the electromagnet

(inset: experimental setup and COMSOL Simulation result at 15V and 0.0911A.) ............ 83

Figure 4.5 Meshed model of the two magnets for the simulation of the magnetic force acting

on the permanent magnet: (a) 2D work plane for coupled geometry; (b) Optimised meshing

conditions for simulation.. .................................................................................................... 84

Figure 4.6 Superimposed magnetic flux density on the surface of the permanent magnet

versus the actuation current (Inset: COMSOL simulation result at 15V with 9,11 mA

yielding 203.36 mT.) ............................................................................................................ 86

Figure 4.7 Magnetic force versus applied current (Inset: COMSOL Simulation result at 15V

and 0.973A generating 1.2N force on the surface of the PM).. ........................................... 87

Figure 4.8 Simulated actuation force on the permanent magnet at different distances between

the two magnet (actuation voltage of 15V with the corresponding actuation current of 91.1

mA). ..................................................................................................................................... 88

Page 15: Design and Development of Cell Stretching Platforms for ...

xiv

Chapter 5

Figure 5.1 Schematic illustration of the working principle: (a) Steps to obtain static strain; (b)

ON state for cyclic stretching condition; (c) Actual cyclic cell stretching platform and the

PDMS device ....................................................................................................................... 96

Figure 5.2 Simulated axial displacement versus magnetic remanence. (Inset: FEA results with

Br=1.2T, selected NdFeB grade N35 magnet, PDMS device with static strain condition)…

.............................................................................................................................................. 99

Figure 5.3 Magnetic force over the voltage range of 1- 30V (Inset: Experimental setup and

FEA model for PDMS device). .......................................................................................... 102

Figure 5.4 Strain on the deformable membrane over the voltage range of 1-30V. (Inset:

Experimental arrangement, the membrane in ON and OFF state, an example for the

particle detection and tracking.) ......................................................................................... 104

Figure 5.5 Strain pattern on the membrane with a selected input of 27 V: (a) Experimental

results; (b) Simulation results.. ........................................................................................... 105

Figure 5.6 Analysis results of cell area and aspect ratio for cyclic, static and no stretch

conditions over the predefined time points (Inset: fluorescent images of the fibroblast cells

with 20x objective for corresponding time points with 50-µm scale bar).. ....................... 106

Figure 5.7 Cell orientation analysis results for cyclic stretching (1.4%.0.01Hz,50% Duty

Cycle) , static stretching (1.4%) and no stretching condition at 0 hr , 0.5 hrs ,1 hrs, 2 hrs , 3

hrs and 4 hrs . (Inset: fluorescent images of the fibroblast cells with 10x objective for 0 hr

(a-c), 0.5 hrs (d-f), 1 hrs (g-i), 2 hrs (j-l), 3 hrs (m-n) and 4 hrs (p-r) with 100-µm scale

bar)... .................................................................................................................................. 108

Chapter 6

Figure 6.1 The pneumatically actuated cell-stretching platform: (a) Schematic illustration of

the cell stretching platform in the OFF state. (b) Schematic representation of working

principle of the device with OP1 and OP3 in the ON state. (c) FEA model results

illustrating the OP1 and OP3 behaviour during ON state. (d) Lower PDMS layer (4x2)

with deformable output patterns and control pattern. (e) Actual multilayer PDMS cell-

stretching platform ............................................................................................................. 119

Figure 6.2 Simulation results of the maximum planar displacement versus the inner post

radius (Ri) versus (Inset: optimised geometry for the deformable OP1, OP2 and OP3, and

the membrane deformation results for OPs with applied pressure of 80 mbar)... .............. 122

Page 16: Design and Development of Cell Stretching Platforms for ...

xv

Figure 6.3. Fabrication process for cell stretching platform. .................................................. 123

Figure 6.4 Developed algorithm to detect the inflated membrane to fit the curve and calculates

the displacement and strain parameters. (a) Input: Reference Image and scale (pixels per

mm). (b) Image enhancement process: Normalised contrast and sharpening. (c)

Thresholding and Edge detection using canny. (d) Dilation operation and noise removal for

curve detection. (e) Fill the holes to enhance mask calculations by eliminating minor

curves. (f) Calculate mask and get minimum Y coordinates along the curve for selection.

(g) Generate symmetric reflected curve along major axis for curve fitting. (h) Apply least

Square criterion for curve fitting. (i) Calculate curve parameters and display results. ...... 125

Figure 6.5 Comparison of the simulation and experimental data: (a) Maximum displacement

versus pressure. (Inset: simulation results with pressure of 150 mbar); (b) Circumferential

strain versus pressure. (Inset: experimental setup). ........................................................... 126

Figure 6.6 Comparison of the membrane deformation for OP1, OP2 and OP3 with pressure

input 80 mabr. (a) Experimental results (actual membrane deformation in XZ plane) (b)

Simulation results (FEA model results for membrane deformation in XZ plane) .. .......... 127

Figure 6.7 Output membrane deformation pattern of OP1: (a) Experimental results. (b)

Simulation results... ............................................................................................................ 128

Figure 6.8 Averaged cell orientation in degrees over the ROI for OP1 and OP2 with

corresponding 10X cell images of each ROI and the COMSOL strain distribution map. Scale

bar is 100 µm. ....................................................................................................................... 129

Figure 6.9 Averaged cell orientation in degrees over the ROI for OP3 (a) and Control (b) with

corresponding 10X cell images of each ROI and the COMSOL strain distribution map.

Scale bar is 100 µm. ........................................................................................................... 130

Figure 6.10 Percentage of cells in given direction over the ROI and ANOVA analysis for

OP1, OP2, OP3 and Control. The cell orientation varies according to strain amplitude

(***p< 0.001, and *p< 0.05 versus control; ##p< 0.01 versus the lowest strain amplitude;

n.s., not significant). Results are expressed as means ± s.d. (n = 3) and data were analysed

using one-way ANOVA followed by post hoc Bonferroni’s test. ..................................... 131

Figure 6.11 Cytoskeleton organisation for OP1 (a), OP2 (b) and OP3 (c) and Control (d) with

20x Objective (scale bar: 50um) after 2 hours. .................................................................. 132

Chapter 7

Figure 7.1 96-well plates with cell stretching platform for high-throughput screening…….140

Page 17: Design and Development of Cell Stretching Platforms for ...

16

Chapter 1: Introduction

1.1 Background and motivation

Cells within a multicellular biological system are subjected to various mechanical stimuli

originated from their external and/or internal environment. These mechanical stimuli play an

important role in the development and maintenance of the proper cell functions (homeostasis)

such as migration, proliferation, and growth.1-3

The ability of cells to sense these mechanical

stimuli and convert them into biochemical responses is known as mechanotransduction which

is an integral part of mechanobiological signalling.4-6

At the cellular level, mechanobiological

signalling is important to modulate cell homeostasis. It plays an important role to influence

cellular activities such as cellular alignment and pattern formation during tissue regeneration

and maturation.7-10

For instance, myocytes and their internal fibres align to form myocardium

sheet in vivo due to tensile stress emerged as a result of a heart beating.11

Similarly, shear

stress caused by the change in vascular wall size results in circumferential alignment of

smooth muscle fibres within vascular walls, in vivo.12, 13

Many studies have reported similar

effects of mechanical stimuli in vitro. For example, Hirata, et al. (2008) reported that external

low tensile strain acting on fibroblast cells leads to actin polymerisation,14

while another study

observed migration and apoptosis of the fibroblast cells under high tensile strain.15

In another

study, cell reorientation was observed due to uniaxial stretching.16

For cells, responding to these mechanical cues involves a complex cascade of cellular

signalling which mainly consists of mechanosensing, signal transmission and targeted

response.17

Upon external mechanical loading, transmembrane integral mechanoreceptors,

known as integrin establishes connections between extracellular matrix (ECM) and

cytoskeleton to distribute mechanical load and induce cytoskeleton deformation.18-20

Similarly, other targeted proteins such as cadherin complexes, G-Protein and protein kinases

etc. respond to mechanical cues by mediating gene expression and protein assembly to secrete

assembled proteins into extracellular microenviorment.21-23

According to some studies, the maladaptation of this mechanobiological signalling pathway

may contribute to the pathogenesis of various diseases such as heart failure and asthma.24-27

Page 18: Design and Development of Cell Stretching Platforms for ...

17

Hence, better understanding of mechanobiological signalling has become a major area of

interest for researchers in the field of regenerative medicine and bioengineering. Study of

mechanobiological signalling pathway will provide an insight on how mechanical cues alter

cellular activities and behaviour. However, complex physiological cell microenvironment

makes it difficult to perform controlled mechanobiological investigations in vivo. Thus, most

of the mechanobiological investigations rely on the development of in-vitro cell stretching

techniques to mimic the cells’ dynamic microenvironment and provide insight into the

complex mechanobiological interplay. Indeed, various commercial and customised in vitro

cell stretching platforms have been developed over recent years.28-38

However, the

development of a reliable in vitro cell-stretching approach, which could facilitate high

throughput cell stretching assays with high in vivo biological relevance, still remains a

challenging task.

To address this technological gap, the present thesis describes the development of

Polydimethylsiloxane (PDMS)-based cell-stretching platforms to conduct mechanobiological

investigations. In Chapter 3 of this thesis, an electromagnetic based single-sided uniaxial cell

stretching was designed and developed. The electromagnetic actuation was formulated as

numerical models in Chapter 4 to optimise the operation parameters for the developed cell-

stretching device. The optimisation resulted in the subsequent development of an

electromagnetically actuated double-sided cell-stretching device described in Chapter 5,

which facilitates a wide homogenous strain range with static and dynamic operating modes.

Although, the developed cell-stretching devices provided reliable tools for in vitro

mechanobiological studies, the biological relevance could be improved. Thus, in Chapter 6

the capability of the cells to respond to external mechanical stimuli was explored to engineer a

patterned cell culture which better mimic in vivo cellular condition and facilitate highly

relevant in vitro biological models. With the knowledge gained, a novel pneumatically

actuated cell-stretching array platform was developed to engineer patterned cell culture in

vitro.

1.2 Research scope and objective

This thesis aims to introduce novel cell stretching assays with high biocompatibility and

biological/clinical relevance, which are easy to fabricate and operate and to establish a

systematic approach for detailed mechanobiological investigation.

Page 19: Design and Development of Cell Stretching Platforms for ...

18

In this thesis, electromagnetic and pneumatic actuations were utilised for cell-stretching

platforms. In the Chapter 3, 4 and 5, the thesis describes the detailed design, characterisation

and optimisation of the electromagnetic cell-stretching platform for single- and double-sided

axial stretching as reliable tools for in vitro mechanobiological studies. Then Chapter 6

utilises the gained knowledge to develop a pneumatically actuated cell stretching platform to

overcome limitations of the electromagnetic cell-stretching platform to achieve a more

accurate in vitro biological model. This cell-stretching platform was capable of engineering

patterned cell culture which closely mimics the in vivo cellular conditions. Furthermore, the

platform was designed with the same dimension as a standard 96 well plates to achieve higher

compatibility for high throughput cell stretching assays.

Collectively, the thesis mainly focused on the design, simulation and extensive

characterisation of different devices to develop highly reliable and biologically relevant cell-

stretching platforms. The biological studies described in this thesis were mainly aimed to

demonstrate the capability of the cell-stretching platforms. Therefore, the scope was limited to

the basic biological observations and in-depth biomolecular analysis of the stretched cells is

not covered.

1.3 Thesis framework

The present thesis is divided into seven chapters; Chapter 1: Introduction, Chapter 2:

Published paper 1 (Critical review): Cell stretching devices as research tools: engineering and

biological considerations; Chapter 3: Published paper 2 (Research article): An

electromagnetic cell-stretching device for mechanotransduction studies of olfactory

ensheathing cells; Chapter 4: Published paper 3 (Research article): Modelling of an uniaxial

single-sided magnetically actuated cell-stretching device; Chapter 5: Published paper 4

(Research article): An electromagnetically actuated double-sided cell-stretching device for

mechanobiological research; Chapter 6: Paper 5 submitted for publication (Research article):

Pneumatically actuated cell-stretching array platform to engineer cell pattern in vitro; Chapter

7: Conclusion and future perspective.

Page 20: Design and Development of Cell Stretching Platforms for ...

19

Chapter 1 is an introduction to the thesis topic and includes brief background, motivation,

overall research scope and objective, research questions, thesis framework with specific aims

and general methods and materials. Chapter 2 is a critical review on cell stretching devices as

research tools emphasizing engineering and biological considerations. This chapter provides

details on the different types of cell stretching devices, major design consideration and

biological requirements with a future perspective in this research area. The exhaustive

literature review provides an insight on the technological gap and possible approaches for the

development of robust cell stretching platforms. Considering the limitation of the existing

cell-stretching platforms, this thesis intended to provide a systematic approach in designing

cell-stretching devices for in-depth mechanobiological investigations with high biological

relevance. This raised the research questions and aims discussed in the following section

which were subsequently addressed in Chapter 3, 4, 5 and 6. Finally, Chapter 7 provides a

conclusion and future perspective for the research presented in this thesis.

1.4 Research questions and specific aims

Research question 1 (Chapter 3)

Various cell stretching approaches have been developed to induce homogeneous strain onto

the cell culture and establish a relationship between the strain amplitude with cellular

response. An ability to observe this cellular response under the same biological conditions for

different strain amplitudes within the same field of view would be an added advantage. Thus,

an understanding of the strain requirement for the cells under investigation is a critical factor

for mechanobiological studies. The ability to observe the effect of the different strain

amplitudes on cell morphology, behaviour and viablity could provide insight on the strain

requirement for the cell under investigation. This unique feature serves as a first step in

establishing systematic approach for mechanobiological investigation. Hence, the research

question of this chapter was: How to implement single-sided cell stretching using magnetic

force to induce heterogeneous strain onto the cell culture and how to observe the effect of the

different strain amplitudes in the same field of view?

Specific aim 1 (Chapter 3)

The study aimed to design and develop a robust cell stretching platform that could provide a

heterogeneous strain pattern and serve as a tool to establish strain requirement for the cells

Page 21: Design and Development of Cell Stretching Platforms for ...

20

under investigation. The tool facilitates a better comparison of the cellular response to

different strain amplitudes in the same field of view. The study successfully developed a well

characterised electromagnetic cell-stretching platform using the single-sided uniaxial

stretching approach. The platform was capable of inducing heterogeneous strain onto the cell

culture and facilitated the comparison of the cellular response towards different strains in the

same field of view. The platform was tested with olfactory ensheathing cells (OECs) to

observe their morphological changes for different strain amplitudes and to determine the

strain requirement for OECs.

Research question 2 (Chapter 4)

All existing cell stretching approaches have their own particular advantage. However, the

standardisation of cell stretching approaches to obtain conclusive translational models with

biological relevance is still deficient. Thus, a detailed understanding of the actuation strategy

is an important step towards standardisation of the developed electromagnetic cell-stretching

platform. Numerical system modelling and simulation with finite element analysis (FEA)

provides a design tool to obtain optimised parametric values for the operation of the

developed cell-stretching platforms. Such tools play an important role in exploring the

possible alternative strategies to achieve standardised cell stretching approaches. Therefore,

the research question of this chapter was: Can a standardised numerical model of the uniaxial

electromagnetic cell stretching platform be developed for the parametric optimisation using

FEA modelling approach?

Specific aim 2 (Chapter 4)

This study intended to develop a finite element analysis (FEA) model of the uniaxial

electromagnetic cell stretching platform for parametric optimisation and standardisation. The

study resulted in the development of the systematic FEA modelling approach for

electromagnetic actuation, coupled with structural analysis, which provided a standardised

numerical model for the uniaxial electromagnetic cell-stretching platform. The capability of

the standardised numerical model was demonstrated by optimising the distance between the

electromagnet and permanent magnet.

Page 22: Design and Development of Cell Stretching Platforms for ...

21

Research question 3 (Chapter 5)

The developed electromagnetic uniaxial single sided cell stretching platform was suitable to

obtain strain requirement for the cell type under investigation. However, a further in-depth

study of homogenous strain is critical for a robust cellular response analysis.

The next logical step was to optimise the cell-stretching platform using existing numerical

tools to achieve a homogenous strain pattern, which will further improve the capability to

systematically investigate the cellular response for mechanobiological studies. Moreover,

previous studies have shown that different stretching conditions induce different cellular

responses. Therefore, an ideal cell stretching platform should incorporate a variety of

stretching modes into a single device, thus facilitating experimental flexibility for more

exhaustive cell stretching studies. For instance, the comparative study of the homogeneous

static or dynamic strain is critical for observing cellular alignment and reorientation in the

fields of tissue engineering and medication testing. An optimised cell-stretching platform

capable of inducing homogeneous strain onto the cell culture that facilitates different

stretching mode was needed to address the above mentioned drawbacks. Hence, the research

question of this chapter was: How can an active magnetic repulsion force be utilised to

develop a cell stretching platform with homogenous strain and different stretching modes?

Specific aim 3 (Chapter 5)

The aim of this study was to optimise an electromagnetic uniaxial single sided cell stretching

platform to incorporate different stretching modes with homogeneous strain pattern. Building

on the gained knowledge, and utilising a previously developed standardised numerical model

of the uniaxial single-sided cell stretching platform the effect of axially aligned two

permanent magnets was explored to achieve the homogeneous strain condition. This study

resulted in the successful development of an electromagnetically actuated double-sided cell

stretching platform to introduce homogenous strain onto the membrane using two axially

aligned electromagnetic actuators. The versatility of the cell-stretching platform is

demonstrated through its capability of introducing homogeneous strain on to the cell culture

and facilitating both static and dynamic stretching modes. The platform was tested with

fibroblast cells to observe their cellular behaviour under homogeneous static and dynamic

stretching conditions.

Page 23: Design and Development of Cell Stretching Platforms for ...

22

Research question 4 (Chapter 6)

Different cells within a human body are known to have different morphological and

behavioural patterns in vivo, depending on their anatomical location. The developed cell

stretching platforms systematically provide well standardised tools for the strain requirement

of cells under investigation and in-depth mechanobiological studies. However, to further

improve mechanobiological understanding, more accurate tools are needed to accurately

mimic the complex in vivo biological environment. Although various cell stretching studies

have shown the effectiveness of external mechanical stimuli to reorient and realign the cells,

very few have explored the cells’ mechanobiological signalling to engineer patterned cell

culture that provides an additional advantage to closely mimic the in vivo pattern condition of

the cells. Thus, a new approach is required to engineer patterned cell culture to better mimic

in-vivo cellular conditions and enhance the biological relevance of the cell-stretching

platform. The research question of this chapter was: How to utilise mechanotransduction

capability of the cells to actively manipulate the cell orientation and obtain patterned cell

culture?

Specific aim 4 (Chapter 6)

This study aimed to introduce an effective approach by exploring the mechanotransduction

capability of the cells to engineer patterned cell culture similar to obtain highly relevant in

vitro biological models. Moreover, for improving experimental flexibility an array based

pneumatic actuation approach was designed for this study. The study also aimed to improve

the compatibility of the developed platform by maintaining the dimension of the cell

stretching platform the same as the F bottom 96 well plate which was the simple and effective

approach. The detailed study resulted in the development of the pneumatically actuated cell-

stretching array platform, which can concurrently induce a strain onto the cell culture array to

engineer circumferential cellular pattern in vitro to mimic in vivo circumferential cellular

alignment of the smooth muscle fibres due to the shear caused by the changes in vascular wall

sizes. The study successfully demonstrated the capability of the platform to engineer cell

culture patterns and obtain biological relevant models in vitro. Additionally, array-based

approach further enhanced the compatibility of the developed cell stretching platform with

clinical tools.

Page 24: Design and Development of Cell Stretching Platforms for ...

23

1.5 General methods and materials.

This section is a brief description of the basic methods and materials that apply to all the

research work carried out for this thesis. All specific methods and materials for each chapter

have been elaborated in details within the respective chapter.

1.5.1 Finite Element Analysis (FEA) modelling.

For the purposes of this thesis, COMSOL Multiphysics 5.2 was utilised as a FEA modelling

platform for system behavior study and parametric optimisation. In the conceptual phase of

the each developed cell-stretching platform, the reference geometry was adopted in COMSOL

and the geometric optimisation was carried out by varying one parameter while keeping

others constant. For each cell-stretching platform, the key optimisation parameters were

observed and critically investigated to finalise the design. In most cases, structural mechanics,

AC/DC physics was utilised for formulating the numerical model. The appropriate materials

were selected for the reference geometry to obtain a realistic FEA model. Necessary study

with the required constraints and input values were implemented to achieve the realistic

simulation results. Finally, the geometry was meshed and solved to observe system behavior.

Parametric sweep option in COMSOL was extensively used to implement variation in inputs

for parametric optimisation.

1.5.2 Fabrication processes

In this thesis, conventional machining with cutting tools, soft lithography, PDMS casting, spin

coating, plasma bonding and 3D printing approaches have been utilised to fabricate different

parts of the developed cell stretching platforms. In brief, master moulds for replicating PDMS

devices in electromagnetic cell stretching platforms are fabricated utilising conventional

milling, while soft lithography was opted for pattern replication in the development of the

pneumatic array-based cell stretching platform. Additionally, spin coating was utilised to

achieve a 200-µm thick deformable PDMS membrane whereas, mounting platforms and

constrain clip for electromagnetic cell stretching platforms were fabricated using 3D printing.

Conventional machining. For replicating the optimised geometry of the PDMS device for

electromagnetic cell stretching platforms, the master mould was fabricated from aluminum, a

non-magnetic material. First, all the parts were designed in SolidWorks 2013 considering

Page 25: Design and Development of Cell Stretching Platforms for ...

24

fabrication tolerance. Each part was then carefully milled and drilled to achieve the design

specification. Finally, all the parts were screwed and assembled to obtain the master mould

for the PDMS device replication.

Soft lithography. In the case of the pneumatic array-based cell stretching platform, the

master mould with the pattern was designed in CleWin 3.0 software. The optimised pattern

was printed on the 28 mm x 22 mm plastic mask for replication. For preparing the master

mould, a cleaned silicon wafer was pre baked at 80oC for 30 min. The baked wafer was then

spin coated with SU-8 50 photoresist at 3,300 rpm for 30 sec to achieve 100-m thickness. To

avoid SU-8 cracking, a series of pre baking steps (65°C for 10 min, 95°C for 30 min, 50°C for

30 minutes) were performed. Subsequently, the UV exposure at 9.8mW/m cm2 for 35 seconds

allowed pattern transfer. Finally, the SU-8 was developed in 1-methoxy-2-propanol acetate

for 15 minutes after post baking at 80°C for 30 mins.

PDMS casting. If not stated otherwise, in general, a clear and degassed PDMS and cross

linker (Sylgard 184 elastomer kit, Dow Corning, USA) was prepared at a volume ratio of 10:1

for PDMS device casting. The mixture was carefully poured onto the master mould and again

degassed for 15 mins to remove any remaining air bubbles. The mould was then cured for 2

hours at 80C in a vacuum oven. The cured PDMS device was inspected and carefully peeled

off from the master mould. Finally, the PDMS device was then cleaned using isopropanol and

DI water.

Spin coating. In this thesis, considering the plastic nature of the PDMS material used for

membrane fabrication thickness of the membrane was kept comparatively high at 200 µm to

limit the PDMS membrane degradation. The deformable membrane was fabricated with a

clear degassed PDMS-crosslinker mixture in 10:1 volume ratio. The mixture was spin coated

at 400 rpm for 2 mins to achieve the 200-µm thickness.

Plasma bonding. In this thesis, unless stated otherwise, the deformable PDMS membrane is

plasma bonded to the PDMS devices using oxygen plasma (Harrick Plasma). Prior to surface

activation, the surface of both membrane and device was thoroughly cleaned using

isopropanol and deionised (DI) water to ensure the dust free surface. The cleaned surface of

Page 26: Design and Development of Cell Stretching Platforms for ...

25

the membrane and device was then activated using oxygen plasma for 45 seconds followed by

bonding and 1 hour curing at 80°C in a vacuum oven.

3D printing. Throughout this research work SolidWorks 2013 was utilised as a design tool.

In the case of electromagnetic cell stretching platforms, to provide necessary constraint and

alignments for the PDMS devices, electromagnets and permanent magnets, 3D CAD model of

the mounting platforms and constrain clips was designed in SolidWorks 2013.

Cells are sensitive to the temperature changes thus, while designing the mounting platform the

electromagnet heating issue was critical. To facilitate cooling the platforms were designed to

minimise the electromagnet surface covering. Moreover, cooling tubes and heat sink

mounting was included to avoid heating issue. After optimisation and considering the

fabrication tolerance the designs were finalised and imported into Eden 260V 3D printer

(Stratasys Ltd.) in standard STereoLithography (STL) format. Finally, the mounting platforms

and constrain clips were 3D printed with photopolymer resin.

1.5.3 Image processing for detection and tracing of the deformation.

In this thesis, custom-made algorithms based on image processing techniques were developed

and implemented in MATLAB to detect and trace axial and planar membrane deformation. In

general, deformation was recorded using a digital camera (EO Edmund Optics). A minimum

of three image frames for each data set was analysed to obtain averaged parametric values.

Furthermore, each experiment was repeated at least three times to maintain the integrity of the

results.

For axial deformation, random points were marked on the region of interest (ROI). The axial

deformation (XY-plane) was recorded using a digital camera (EO Edmund Optics). The

images were then analysed using an algorithm mainly based on Digital Image Correlation

(DIC) and MATLAB Image Processing Toolbox to detect the marked points and to obtain

effective displacement and to calculate the strain profile (Appendix A1). In the case of planar

deformation, the side view (XZ-plane) of the deformation was captured using a digital camera

(EO Edmund Optics) with 1x magnification mounted on a XY stage (Zaber Technologies).

The planar deformation was quantified by using a custom-made algorithm to detect the

Page 27: Design and Development of Cell Stretching Platforms for ...

26

deformation curve, curve fitting then provides the displacement and strain parameters

(Appendix A2).

1.5.4 Cell culture

Unless otherwise stated, the standard cell culture protocol utilised in this thesis is as follows:

The cell stretching platform was sterilised with 80% ethanol followed by 20 mins UV

exposure to avoid contamination. The device was subsequently treated with fresh media and

incubated in a standard incubator for 1 h at 37oC and 5% CO2. The cells were cultured in a

standard T75 flask with DMEM/F12 (Gibco, Thermo Fisher Scientific) medium enriched with

10% FBS and 1% antibiotics. The cells were incubated at 37oC and 5% CO2 in a standard

incubator to achieve 70% confluency. For seeding, 70% confluent T75 flask was trypsinised

with TrypLE Express (Thermo Fisher Scientific) and the cells were resuspended in the fresh

media. Finally, aliquots of the optimised seeding density were used to seed the device.

Moreover, the cell stretching platform was incubated for 24 h in a standard incubator at 37oC

and 5% CO2 to ensure proper cell growth and cell adhesion onto the membranes.

1.5.5 Cell fixing, immunofluorescence staining, and imaging

Stretched cells were immediately fixed using 4% Paraformaldehyde (PFA) followed by 3x

Phosphate buffer solution (PBS) washes. The cells were stored in cold PBS solution at 4oC

untill further staining and imaging. ActinGreenTM 488 and NucBlueTM ReadyProbeTM

Reagent (ThermoFisher Scientific) were used to stain nuclei and actin fibres of the cells,

respectively. The standard staining protocol was optimised and adapted for the developed

cell-stretching platforms. Finally, inverted phase contrast or fluorescent microscope with

appropriate magnification was used for imaging.

1.5.6 Image processing for the biological analysis.

For the biological analysis, at least three images were obtained and analysed along each

region of the device. Triplicates of each biological experiment were performed to confirm the

repeatability of the results. In general, post processing of the images obtained during the

experiment was performed using ImageJ (NIH) to enhance the image quality for quantitative

analysis. The post processing steps mainly involved FFT bandpass filtering, sharpening,

enhancing image contrast and thresholding.

Page 28: Design and Development of Cell Stretching Platforms for ...

27

1.6 References

1. B. D. Hoffman and J. C. Crocker, Annu. Rev. Biomed. Eng., 2009, 11, 259-288.

2. C. L. Happe and A. J. Engler, Circ. Res., 2016, 118, 296-310.

3. M. Aragona, T. Panciera, A. Manfrin, S. Giulitti, F. Michielin, N. Elvassore, S.

Dupont and S. Piccolo, Cell, 2013, 154, 1047-1059.

4. R. P. Brandes, N. Weissmann and K. Schroder, Antioxid. Redox Signal., 2014, 20,

887-898.

5. D. J. Leong, J. A. Hardin, N. J. Cobelli and H. B. Sun, in Skeletal Biology and

Medicine Ii: Bone and Cartilage Homeostasis and Bone Disease, ed. M. Zaidi,

Blackwell Science Publ, Oxford, 2011, vol. 1240, pp. 32-37.

6. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

7. F. Klingberg, B. Hinz and E. S. White, The Journal of pathology, 2013, 229, 298-309.

8. B. D. Hoffman, C. Grashoff and M. A. Schwartz, Nature, 2011, 475, 316-323.

9. S. Y. Chew, R. Mi, A. Hoke and K. W. Leong, Biomaterials, 2008, 29, 653-661.

10. S. Hoehme, M. Brulport, A. Bauer, E. Bedawy, W. Schormann, M. Hermes, V. Puppe,

R. Gebhardt, S. Zellmer, M. Schwarz, E. Bockamp, T. Timmel, J. G. Hengstler and D.

Drasdo, Proc Natl Acad Sci U S A, 2010, 107, 10371-10376.

11. P. Helm, M. F. Beg, M. I. Miller and R. L. Winslow, Ann N Y Acad Sci, 2005, 1047,

296-307.

12. M. B. Chan-Park, J. Y. Shen, Y. Cao, Y. Xiong, Y. Liu, S. Rayatpisheh, G. C. Kang

and H. P. Greisler, J. Biomed. Mater. Res. A, 2009, 88, 1104-1121.

13. M. Morioka, H. Parameswaran, K. Naruse, M. Kondo, M. Sokabe, Y. Hasegawa, B.

Suki and S. Ito, PLoS One, 2011, 6, e26384.

14. H. Hirata, H. Tatsumi and M. Sokabe, J. Cell Sci., 2008, 121, 2795-2804.

15. W. Qian, Z. Xu and Z. Yi, JMiMi, 2013, 23, 015002.

16. J. H. Wang, P. Goldschmidt-Clermont, J. Wille and F. C. Yin, J Biomech, 2001, 34,

1563-1572.

17. C. R. Ethier and C. A. Simmons, Introductory Biomechanics: From Cells to

Organisms, Cambridge University Press, 2007.

18. C. Tamiello, A. B. Buskermolen, F. P. Baaijens, J. L. Broers and C. V. Bouten, Cell

Mol Bioeng, 2016, 9, 12-37.

Page 29: Design and Development of Cell Stretching Platforms for ...

28

19. X.-m. Liu, K. J. Peyton, A. R. Shebib, H. Wang and W. Durante, Biochem.

Pharmacol., 2011, 82, 371-379.

20. P. S. Howard, U. Kucich, R. Taliwal and J. M. Korostoff, J. Periodontal Res., 1998,

33, 500-508.

21. J. H. Wang and B. P. Thampatty, Biomech Model Mechanobiol, 2006, 5, 1-16.

22. M. A. Wozniak and C. S. Chen, Nat. Rev. Mol. Cell Biol., 2009, 10, 34-43.

23. T. M. Maul, D. W. Chew, A. Nieponice and D. A. Vorp, Biomech Model

Mechanobiol, 2011, 10, 939-953.

24. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

25. D. E. Ingber, Ann Med, 2003, 35, 564-577.

26. C. Hahn and M. A. Schwartz, Nat. Rev. Mol. Cell Biol., 2009, 10, 53-62.

27. P. A. Janmey and R. T. Miller, J Cell Sci, 2011, 124, 9-18.

28. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog. Biophys. Mol. Biol., 2014, 115, 93-102.

29. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

30. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

31. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

32. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

33. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

34. O. Chaudhuri, S. H. Parekh, W. A. Lam and D. A. Fletcher, Nat. Methods, 2009, 6,

383-387.

35. Q. Wang, X. Zhang and Y. Zhao, Sensors Actuators B: Chem., 2014, 194, 484-491.

36. W. W. Ahmed, T. Wolfram, A. M. Goldyn, K. Bruellhoff, B. A. Rioja, M. Moller, J.

P. Spatz, T. A. Saif, J. Groll and R. Kemkemer, Biomaterials, 2010, 31, 250-258.

37. S. P. Arold, J. Y. Wong and B. Suki, Ann. Biomed. Eng., 2007, 35, 1156-1164.

38. K. Sato, S. Kamada and K. Minami, IJMS, 2010, 52, 251-256.

Page 30: Design and Development of Cell Stretching Platforms for ...

29

Chapter 2: Cell stretching devices as research tools: engineering and

biological considerations

Cells within the human body are subjected to continuous, cyclic mechanical strain

caused by various organ functions, movement, and growth. Cells are well known to

have the ability to sense and respond to mechanical stimuli. This process is referred to

as mechanotransduction. A better understanding of mechanotransduction is of great

interest to clinicians and scientists alike to improve clinical diagnosis and

understanding of medical pathology. However, the complexity involved in the in vivo

biological systems creates a need for better in vitro technologies, which can closely

mimic the cells' microenvironment using induced mechanical strain. This technology

gap motivates the development of cell stretching devices for better understanding of

the cellular response to mechanical stimuli. This Chapter is a critical review on the

engineering and biological considerations for the development of such cell stretching

devices. The chapter discusses different types of stretching concepts, major design

consideration and biological aspects of the cell stretching and provides a perspective

for the future development in this research area. The exhaustive literature review

provided an insight on the technological gap and possible approaches for the

development of robust cell stretching platforms during the thesis.

Page 31: Design and Development of Cell Stretching Platforms for ...

30

Statement of Contribution to Co-authored Published Paper

Page 32: Design and Development of Cell Stretching Platforms for ...

31

2.1 Introduction

Cells experience various kinds of mechanical forces within a normally functioning

body. These forces play an important role for the development of cells as well as for

the regulation of their homeostatic activities. For instance, endothelial cells in blood

vessels are subjected to both shear stress due to the blood flow and cyclic strain due to

the blood pressure. Mechanical forces acting on cells are well known to induce

intracellular biochemical signals, which play important roles in cellular behaviours

such as proliferation, growth and migration.1-3

A cell can sense mechanical stimuli and

convert these stimuli into biological responses through a series of cellular processes,

which is known as mechanotransduction.4-6

Mechanotransduction plays significant

roles in the regulation of the cellular activities and can significantly influence cell

activities.7 Abnormalities in these processes may contribute to the pathogenesis of

several diseases, such as cancer, asthma, heart failure, etc.8, 9

Thus,

mechanotransduction has been a major research interest in the field of regenerative

medicine and bioengineering. Due to the complexity of the in-vivo cellular

environment, most mechanotransduction research relies on the development of in-vitro

techniques with integrated in-vivo like stimuli. Thus, in-vitro cell-stretching assays are

vital for further understanding the dynamics of mechanotransduction.10-12

The

development of in-vitro techniques that better mimic the pathological in-vivo

environment of cells will facilitate researchers to gain greater insight into

mechanotransduction and will significantly improve our understanding of physiology

Fig.2.1: Commercially available cell stretchers: (a) ElectroForce 3100 (Bose

Corporation) (b) Stage Flexer (Flexcell international Cop.) (c) Stretch System for

Microscope STB-150 (STERX Inc.).

Linearguide

Loading

TissueSample

Flow channel

Cell culture

Strain

Vacuum

Loading Post

Loading Post

Membrane

Holder

Stretched Membnrae

Microscope assess

Cell culture Drive

Motor

Fixed Holder Movable Holder(C)

(b)(a)

Page 33: Design and Development of Cell Stretching Platforms for ...

32

and cellular biology for clinical diagnosis and subsequent treatments.

Micropipette, tweezers, atomic force microscopes (AFM) and micro posts with

integrated magnets are some of the most common techniques that have been

traditionally used for cell stretching in clinical diagnosis.13-17

Recently, more elegant

techniques have been pursued commercially, such as Flexcell (Flexcell International

Corporation); Strex Systems for cell Stretching (STREX Inc.) and ElectroForce

(Bose Corporation), (Fig.2.1). Flexcell’s Stage Flexer has been used ubiquitously due

to its well-characterised strain profile, homogenous strain pattern and adaptability of

stretching modes.18-20

Apart from these commercial systems, several other custom-

made cell-stretching devices have been reported over the last decade.21-24

However,

most of these devices have low throughput and generate a non-linear strain profile.

Contemporary research on cell stretching has been significantly influenced by the

state-of-the-art technologies of Micro-Electro-Mechanical Systems (MEMS) and

microfluidics.25-27

Recent advances in MEMS and microfluidics technologies have

facilitated complex operations such as trapping cells, creating more realistic

microenvironments and providing direct observation for quantifying cell behaviour.28-

30 MEMS and microfluidics technologies are both expected to play a significant role in

the future to recreate the cellular microenvironment and provide a more accurate

physiological model in-vitro.

Several review articles have evaluated the different actuation techniques and methods

used for cell stretching.31-34

Although, most of these articles are focused upon different

parameters such as cell stretching methods, cell mechanics or actuation, whereas very

few discuss the engineering as well as biological considerations involved. Furthermore,

most reviews were published more than five years ago, thus warranting an updated

review for this exciting field. In the present review, we provide an in-depth exploration

of cell stretching systems with a focus on the major design considerations and

biological aspects involved in the development of cell stretching device and provide

future perspective for the design, fabrication technologies and materials required for

the development of cell-stretching devices.

Page 34: Design and Development of Cell Stretching Platforms for ...

33

2.2 Actuation concepts

This section discusses the different types of actuation that have been reported in the

literature, where stretching concepts are classified based on the different actuation

techniques used for cell stretching. Furthermore, we elaborate on the suitability of the

stretching concept and other major design considerations for achieving the delicate task

of cell stretching.

2.2.1 Pneumatic actuation

Pneumatic actuators have been widely used to induce mechanical stress or strain to

cells in-vitro. This actuation concept holds substantial advantages such as simple setup,

homogenous strain actuation and no direct contact with the cells and/or the media,

which is important to avoid contamination. The majority of devices using pneumatic

actuation techniques are based on the deformation of a thin membrane with controlled

actuation pressure. The cells are cultured directly onto this membrane.28, 35-37

Actions with both positive and negative pressure sources have been exploited for cell

stretching. For instance, Shimizu, et al. (2011) utilized positive air pressure for

inflating serially connected balloons, (Fig. 2.2a).38

Furthermore, the pressure drop in

microchannels was utilized to deliver a wide range of strain magnitudes within a single

device. Similarly, Heo, et al. (2013) fabricated pneumatic a microactuator consisting of

pneumatic chambers.39

The balloon-like expansion of the chambers stretches the cells

cultured on the membrane (Fig. 2.2b). Huang and Nguyen (2013) developed a multi-

layered PDMS device, in which vacuum source was used to pull the membrane by

deforming the wall attached to the sides of the membrane (Fig. 2.2c). 21

Tremblay, et al. (2014) further advanced the above concept and designed a similar

multi-layered PDMS device with four low-pressure compartments for biaxial

stretching.40

In this device, the independently controlled negative pressure was used to

pull the membrane and deform each compartment to stretch the cell culture in both

horizontal and vertical directions (Fig. 2.2d). Kreutzer, et al. (2014) developed a

circular device with a thin membrane and computer controlled vacuum pressure in the

cavity between the two PDMS shells to deform the membrane.41

This induced a

Page 35: Design and Development of Cell Stretching Platforms for ...

34

symmetrical radial stress onto the inner shell and subsequently stretches the cells

grown on the membrane (Fig. 2.2e).

2.2.2 Piezoelectric actuation

Another popular actuation approach to induce stretching has been piezoelectric micro-

and nanomanipulators. Piezoelectric manipulators using a high displacement resolution

have been included in a number of studies to induce cell stretching, where the key

advantages is the precision and broad range of controllable strains. Whilst the high

resolution of displacement further provides an active tool to control loading during the

process of cell stretching. However, some piezoelectric actuators require a direct

physical contact with cells, which severely limits its applications. Although indirect

stretching through microstructures can be used to overcome this problem, loading

limitations still remain. Nonetheless, it is common to use Micro-Nano manipulators

Stretchedcells

Inlet

Flow channel

Cell culture

Inflated BalloonBalloon structure

Outlet

Flow channel

Cells seeding

Flow channel

MembraneVacuumchamber

Vacuum Vacuum

Stretched cells Stretched membraneStretchedmembrane

Glass Cell cultureConnector

VacuumMembrane

Stretched cells

CellsPDMSsheet

Fibronectin

Pneumaticcompartment

Stretched cells

Pressure

Glass 1

Lowpressurechannel

Lowpressurechannel

Fluidicchannelwithout

cells

Cellsseeding

23 4

Deformablewalls

10um membrane

Vacuumactuation

compartments

Lowpressure

Lowpressure

Stretched cells Stretchedmembrane

(a)

(e)

c)(

(b) (d) Fig.2.2: Typical cell stretching devices with pneumatic actuation: (a) Inflated

balloons with positive pressure; (b) Actuation chamber positive pressure; (c) Two-

chamber side actuation with negative pressure; (d) Four-chamber side actuation with

negative pressure; (e) Radial stretching with circular support and negative pressure.

Page 36: Design and Development of Cell Stretching Platforms for ...

35

which are externally linked to the on-chip microstructures such as micro intendant or

microplate for on-chip cell stretching.42-45

Kamotani, et al. (2008) developed a device with an array of miniature cell stretching

chambers which included microwells with a flexible bottom membrane placed over a

piezoelectrically actuated pin.46

Each pin was independently actuated by individual

piezoelectric actuators using a customised computer program that pushes the bottom

membrane to achieve radial strain on the cells (Fig. 2.3a). Deguchi, et al. (2015) used

two tandemly arrayed piezoelectric actuators to stretch a PDMS membrane, with each

consisting of a chamber with a base membrane (Fig. 2.3b).47

Fior, et al. (2011)

fabricated a MEMS device consisting of microstructure linkages connected to the cell

stretching area.23

This device used an externally controlled piezoelectric actuator to

displace the micro linkages and subsequently displace the membrane (Fig. 2.3c). Sato,

et al. (2010) designed and fabricated a device consisting of elastic transparent micro-

chambers and a MEMS micro-linkage mechanism,48

where six 2mm2 mm devices

Fig.2.3: Typical cell stretching devices with piezoelectric actuation: (a) Radial strain

with pushing pin; (b) Linear stretching with piezoelectric linear drive; (c) MEMS

translation stage with external actuator; (d) MEMS linkage mechanism.

Reservior

Microwell

Cellculture

100ummembrane

Pin

PZT actuator

Computer control

Stretchedmembrane

Stretchedcells

Strain

(a)

Piezoelectricactuator Linear drive

Linear stage

Device holder

S tra in

Cell culture

Base

Spring stage

YX (b)

Micro manipulator

Controller

PZT

Microscope

Microwire

MEMS device

YX

c)(

Stretchingsubstrate

Connector

Supportframe

Stretchingarm

Slider

Pin

(d)

Micromanipulator

Page 37: Design and Development of Cell Stretching Platforms for ...

36

were fabricated on a 22mm glass plate. Each device consisted of a silicone rubber

microchamber, stretching arms, a slider to drive stretching arm, bearing to transmit

linear motion into rotation and micro-needle connected to micro-actuators to push the

slider and to drive stretching arm (Fig.2.3d).

2.2.3 Electromagnetic actuation

Electromagnetic actuators provide another sophisticated alternative for inducing

mechanical stress on cells. Controlled electromagnetic motors have been used to

achieve the desired stretching effect,49-51

however, the ongoing need for lubrication and

the possibility for device erosion are major concerns for cell contamination. Yet,

advantages such as high precision and intuitive programmability with a relatively

simple setup for both static and dynamic loading are both attractive features with this

actuation concept.

Fig.2.4: Typical cell stretching devices with electromagnetic actuation: (a) Linear

pushing with a servo motor; (b) Radial strain with pushing by a stepper motor; (c)

Stretching with linear translation stage; (d) Stretching witch cam follower

mechanism.

Linearactuator

Supportframe

Membraneholding

ring

Microscopestage

Movable plate

Membrane

Microscope access Indenter ring

Linear guide

(a)

Cell culture

Indenter ring

Membraneholding

ring

Cell culture Hetero-

genous

strain

Restriction walls

Homogenousstrain

(Str

etc

he

d)

(Un

str

etc

he

d)

(b)

Clamping module

Mounting platform

Membraneassemble

stage

Stretchingmodule with

cam follower

MotorControl

PDMS cell culture

well

Microscopestage

(d)

StageMovable holder

MountingBase

Micro-pattern

Fixedholder

PDMS substrate

Linear actuator

Linear guide

c)(

Page 38: Design and Development of Cell Stretching Platforms for ...

37

Because of the above advantages, many custom-made cell-stretching devices have used

electromagnetic actuators. Huang, et al. (2010) designed a mechanical cell stretching

system based on the indenter design and drove it with a servomotor.52

The device

consists of a membrane holder ring fixed to a mobile plate, where the indenter ring

placed on the base is smaller than the membrane holder ring. The controlled vertical

downward motion of the mobile plate leads to the stretching of membrane (Fig. 2.4a).

Ursekar, et al. (2014) adapted a similar construction with a stepper motor as the

actuator.24

The authors addressed the issue of heterogamous strain and fabricated a 1-

mm thick and 10-mm tall cylindrical wall on the membrane to confine the cells and to

induce a homogeneous strain (Fig. 2.4b). Chang, et al. (2013) used a translation stage

for motorised stretching of neural stem cells,53

where the flexible stretching substrate

with micropatterns was clamped with one end fixed to the base and other end

connected to the translation stage. Linear motion of translation stage actuated by the

motor transfers the strain to the substrate and results in the stretching (Fig. 2.4c).

Fig. 2.5: Typical cell stretching devices with optical actuation: (a) Stretching in a

microchannel with free-space laser beam

50:50 splitter

Waveguide Microchannel

Connectors

Piezo

Function generator

CW Yb laser

(1070nm)

CCDcamera

Lense

Lightsource

Laserdiode

Imagingobjective

Y

X

Microfluidicdevice

Opticalstretcher

sizemeasurement

MirrorMirror

(a)

(b)

Page 39: Design and Development of Cell Stretching Platforms for ...

38

Shao, et al. (2013) introduced a cam-follower mechanism, driven by an electrical

motor,54

this device consisted of a clamping module with one fixed and one movable

L-shaped plate. The L-plate-cam-follower linkage stretches the PDMS membrane

clamped to the movable L-shaped plate (Fig. 2.4d).

2.2.4 Optical actuation

Optical actuation is another approach for cell stretching. Being a non-contact actuation

technique, optical actuation avoids interference with the cell culture and maintains

sterility. Optical tweezers or optical stretchers have been reported in many clinical

studies,55-58

where a simple optical stretcher incorporates a high power diode laser with

a wavelength and beam width smaller than the cells. The laser beam induces a force

onto the cells along the axis of the laser which causes deformation of the cells.22, 59, 60

A major limitation of this type of technique is that the flow in the microchannel cannot

be controlled precisely. Therefore the cells will settle in the micro-channel and may

form clusters, which disturb flow, ultimately causing non-uniform stretching and

quantification difficult.

Sraj, et al. (2010) demonstrated a method to stretch swollen red blood cells (RBCs)

with an optical trap by a single-beam, diode laser.61

The authors fabricated a

microfluidic channel in PDMS bonded to glass with a cross section of 15 m

150 m, using a single-mode laser diode (830 nm, 200 mW) to optically trap the RBCs

in the microchannel and to induce the deformation. The shape and orientation of the

cells were determined by taking a measurement in a trap and after exiting the trap (Fig.

2.5a). Whereas, Nava, et al. (2015) employed an acoustophoretic method using a

piezoelectric transducer to pre-focus the flow in a microfluidic device by integrating

optical waveguide separated by a distance of 25 m for cell manipulation and

stretching (Fig. 2.5b).62

2.2.5 Other actuators

Apart from the major actuation techniques discussed above, other concepts such as

electro-thermal, dielectrophoretic and electrostatic actuations have also shown

potential

Page 40: Design and Development of Cell Stretching Platforms for ...

39

Electrothermal actuation incorporates a V-shaped beam which is subjected to thermal

expansion stretching along the apex.63-65

Although this technique has been extensively

used for micromanipulation in MEMS technology, it has not yet been exploited due to

the thermosensitivity of cells. However, a number of researchers have employed the

thermal expansion properties of shape memory alloy (SMA) for cell stretching. For

instance, Iwadate and Yumura (2009) used SMA actuators for cyclic cell stretching,

this system consists of four parallel connected SMA coils with one end fixed and the

other connected to the stretched substrate.66

A power supply and a fan were used to

heat and cool the SMA for cyclic stretching. A sequential square wave of 2.6 V was

applied to the SMA to induce cyclic heating and subsequently cyclic stretching of the

membrane (Fig. 2.6a).

Dielectrophoretic actuation is based on cells experiencing a force when exposed to the

high-frequency electric field. This actuation method has been used for cell separation

and positioning, but only a few studies reported cell stretching using this method.

Guido, et al. (2012) fabricated a microfluidic device incorporating two 250-um wide

Fig. 2.6: Other actuation schemes for cell stretching: (a) Thermomechanic actuation

with shape memory alloy; (b) Dielectrophoretic actuation; (c, d) Electrostatic

actuation; (e) Hydraulic actuation.

“X”

str

uctu

re

Quadrant

Folded spring

Comb drive

Input PDMSdevice Output

Glass slide

ConnectorsMicrochannel Microelectrode

Tetherbeam

Fixed finger

Movable finger

Comb drive

Anchor

Movable plate

Fixed plate

Membrane

Microchannel

Wate

r in

let

Output membrane

Upper substrate

Intermediate membraneChannel at upper layer

Middle substrate

Channel at lower layer

Lower substrate

Connector

Strain

PowerCircuit

SMA coil

Movable holder

Membrane

Cell culture

Coverslip

Fixedholder

Microscope

(a)

(b)

(d) (e)

c)(

Page 41: Design and Development of Cell Stretching Platforms for ...

40

indium tin oxide microelectrode separated by a microchannel of 20-um width. 67

Cells

were trapped in the channel, and an electric field (6 Vrms square wave, 15 MHz) was

applied to microelectrodes to induce cell deformation (Fig. 2.6b).

Electrostatic forces were utilised for micromanipulation in many MEMS applications.

An electrostatic actuator primarily consists of a comb drive with a linear range of few

micrometres.10, 68-70

Scuor, et al. (2006) designed and fabricated a biaxial cell

stretcher, which included a circular four quadrant sectional disk and linkages. 68

This

device was designed in such a way, that the quadrants moved simultaneously in a

horizontal and vertical direction to achieve biaxial stretching with the comb drive

connected with linkages to provide linear actuation (Fig. 2.6c).Whilst Shen, et al.

(2008) used a comb drive to induce uniaxial tension on hydrated collagen fibrils.69

The

MEMS device includes collagen fibrils mounted between the fixed pad and the

movable pad, which was further connected to a comb drive. Upon actuation, the

movable fingers of the comb drive pull the movable pad and induce tension to the

collagen fibrils (Fig. 2.6d). In another study utilising electrostatic forces, Wang, et al.

(2014) designed and fabricated a bi-layered microfluidic device made of PDMS and

selected water as the medium for hydraulic actuation. PDMS is porous and permeable

to gas, pneumatic actuation may not create consistent strain on the cells.70

Their device

consisted of four outlet channels and intermediate membranes connected to the fluidic

inlet with a single pump to achieve identical strain patterns, with an intermediate

membrane that subsequently deforms the cells on the outlet channels (Fig. 2.6e).

2.3 Design considerations

Thus far it is evident that several actuation techniques and customised designs can be

used to induce strain to cells in vitro for mechanotransduction analysis. Stretching

parameters such as: magnitude; direction of strain; frequency and time intervals have a

different influence on cell behaviour.71-73

In order to design and develop a well-

characterised cell stretching device, it is important to consider all factors contributing

towards these aforementioned parameters. Apart from these key parameters, other

factors will be discuss in the following section.

Page 42: Design and Development of Cell Stretching Platforms for ...

41

2.3.1 Biocompatibility

Designing a cell-stretching device must take into account factors that are required to

maintain viable cells such as sufficient fresh culture media. Moreover, the

biocompatibility of the device must be warranted to avoid contamination, especially in

the case of contact-based cell stretching techniques. For instance, cell-stretching

techniques with electrostatic and electrothermal actuators66, 69

face serious problems

with the exposure to the cell’s aqueous environment. Electrostatic actuators such as

comb drives are susceptible to reduction in the initial stroke when exposed to cell

culture due to the electrically conductive media. In the presence of liquid media, an

electrothermal actuator may also present limitations due to the fluctuations in

temperature it causes to the cell culture.

Therefore, non-contact based stretching techniques such as pneumatic, magnetic and

optical all seem advantageous options. Although no actuation method can be

considered completely exemplary for stretching, as each technique has its inherent

advantage and disadvantages. Thus, the suitable actuation technique depends on the

specific characteristics of the cells tested.

2.3.2 Substrate properties

Substrate properties such as elasticity, roughness and wettability are factors which

have to be considered when designing cell stretching devices. It is known that some

cells such as a fibroblasts or a smooth muscle cells have sensitive surface adhesion

mechanisms and thus require a desirable surface stiffness in order to attach and

proliferate.74

Gray, et al. (2003) reported a distinct behavioural change in fibroblast cells between

soft and hard substrate stretching devices.75

Tondon and Kaunas (2014) further

demonstrated the significance of substrate rigidity in cell behaviours to applied strain

with 2D and 3D substrate platforms.76

Palchesko, et al. (2012) described

phenotypical changes within cells in response to the surface’s roughness and

wettability.77

Although many surface adhesion proteins such as collagen, fibronectin,

laminin, and matrigel etc. have been used to maintain the desired adhesion

mechanisms, substrate properties still play an important role in cell behaviour. The best

Page 43: Design and Development of Cell Stretching Platforms for ...

42

practise is considering substrate properties and cell adhesion mechanisms when

choosing a suitable substrate. Moreover, adhesion of cells to the substrate also

contributes towards the magnitude of strain experienced by the cells.

2.3.3 Control strategies

The force applied to the cells is the main factor that directly affects

mechanotransduction, and is therefore one of the most important parameters. The

magnitude of the required force depends on the cell type being investigated.

Techniques such as optical tweezers and atomic force microscopy provide an actuation

stress ranging from pN/m2 to N/m

2.13, 17

An inaccurate strain could change the

behaviour, morphology of the cells and lead to irreversible damage.71

Thus, for any

cell stretching system, it is imperative that the stress induced by the actuators has high

accuracy, resolution, and repeatability. System control plays an important role to

maintain these parameters in the desirable range. The controller needs to induce a

desirable magnitude of the force to maintain the well-controlled mechanical

stimulation of living cells. For instance, when designing a pneumatic actuator system,

apart from the general loading control factors: such as air compressibility; leakage; and

gas permeability; the elasticity of the membrane also needs to be considered.

It is interesting to notice that most of the commercially available sophisticated cells

stretching systems have pneumatic, electromagnetic or piezoelectric actuator systems.

Compatibility of these actuation systems with the readily available control system

platform such as NI LabVIEW provides a selection edge over other actuation systems

such as optical actuation, electrostatic or eletrothermal actuation systems. Flexcell

which is acknowledge as one of the most sophisticated commercially available cell

stretching device has a pneumatic actuation system. Control systems with a high

resolution, precision and repeatability are commercially available for these actuation

techniques. Pneumatic actuation system seems to be an attractive candidate for design

as it will be able to provide a non-contact actuation with precise actuation control

system. Considering these control factors and comparing the data of various actuation

techniques indicate that pneumatic actuation systems hold a major advantage over the

other actuation techniques.

Page 44: Design and Development of Cell Stretching Platforms for ...

43

2.3.4 Stretching direction

Stretching direction is a further important parameter when designing a cell-stretching

device. Common examples include uniaxial, biaxial and radial stretching. However,

depending on the assay requirement, different directional types can be included in the

same design. As cells randomly oriented themselves over the in vitro membrane and

not always in the direction of strain, they may experience less strain than that actually

induced. Thus, cell behaviour varies according to the stretching direction,78, 79

although

it has been widely acknowledge that the cellular orientation can be induced using

uniaxial cyclic strain.80, 81

Ahmed, et al. (2010) reported the response of skeletal myoblasts after cyclic stretching

at 00, 45

0, 90

0 directional orientation using micro contact printed fibronectin lines.

50

The strongest development was observed after cyclic stretching of myoblast at 450.

Myoblast at 00 showed a random orientation, and myoblast at 90

0 to pattern did not

show any significant alignment. Chen, et al. (2013) reported that delayed focal

adhesion disassembly in the human bone osteosarcoma epithelial cell line U2OS under

sustained uniaxial stretching occurred in the long axis of the cell, perpendicular to the

stretching direction. This result hints that orientation-specific responses depend on

molecules such as focal adhesion kinase Src, which play an important role in

the mechanotransduction pathway.82

Therefore stretch direction should be considered

when designing a device.

2.3.5 Size of the actuation system

The size of the stretching membrane also needs to be considered during the designing

phase. Although macro scale devices provide high resolution and precision,

unfortunately they consume signicant amounts of reagent. On the other hand,

microdevices require less culture media but have limited precision and loading

capability. Thus, in many reported works, microactuators with macroscopic control

systems are often preferred. The advancement in MEMS and microfluidics will lead to

high throughput on-chip actuation systems, which might overcome the existing the

problems.30, 37

Page 45: Design and Development of Cell Stretching Platforms for ...

44

2.4 Biological considerations

Cells in the human body are constantly interacting and adapting to a variety of physical

forces placed upon them. Some of these forces include: the compressive forces in a

loaded joint; the sheer forces inside the pumping heart or blood vessel; the tensile

forces driven by contracting muscles and the episodic expansion forces within the

lungs. Therefore, to maintain tissue homeostasis cells within tissue must proliferate,

differentiate, migrate, modulate gene expression and remodel, subject to these forces

and thus prevent injury or detrimental pathologies, such as tendinopathies,

cardiomyopathies and atherosclerosis.83-85

But it is unclear how cells actually respond

to such mechanical forces (mechanobiology) and how they are able to transduce these

biophysical forces into biomolecular events (mechotransduction) for adaption.

Comprehension of how mechanical force influences tissue homeostasis will ultimately

lead to advancements in biological therapeutics i.e. bioengineering, biomaterials and

regenerative medicine. The following section will provide a brief review on some

research that has been explored in these two areas of cellular biology, while providing

some specific examples of in vitro cellular and tissue/organ mechanobiological

research.

2.4.1 Mechanobiology

In a way, Empedocles over 2 millennia ago was right, all sensations we perceive from

the world is by physical touch. Yet biological science is only beginning to understand

how physical forces influence cellular physiology, while physicians and scientists are

just uncovering the mechanical basis aetiologies for some diseases such as

osteoporosis, stroke and asthma. However, before we can understand mechanobiology,

we must first explore the architecture of tissue itself. Tissue is a complicated sandwich

of cells, reinforced by a lattice work of protein fibres (extracellular matrix: ECM),

which is mirrored inside each individual cell.86

Thus, being a sandwich of organelles

inside a membrane held together by a sea of filamentous cytoskeletons. This

intracellular cytoskeletal framework is not a passive device,9, 87

rather a complex

transducer for the forces that pass onto and through it.

Page 46: Design and Development of Cell Stretching Platforms for ...

45

The cell’s cytoskeleton can further distribute and balance physical forces through their

adherence to the protein fibres in the ECM, which creates a complicated viscoelastic

phenomenon within cells and tissues.88, 89

Trepat, et al. (2007) demonstrated that the

cytoskeleton can actually transform between fluid and solid phases depending on the

mechanical loads exerted on it.12

Therefore, mechanical loading of the cytoskeleton

and the ECM will affect the appearance and behaviour of cells. Ingber (2003),

reinforces this notion and suggests that physical forces play a critical role in the

development, differentiation and maintenance of all living cells/tissue.9 For example,

chondrocyte ECM gene expression and proliferation was the highest under cyclic

compressive loads compared to constant loads.90

Osteoblasts proliferate and lay down

more tissue under a constant frequency (1Hz) and low stain.91, 92

Repetitive stretch and

relaxation of skeletal muscle cells increased their elasticity and bulk.93, 94

Endothelial

cells respond most efficiently to shear forces,95

and stretch force differentiates

ligamentous fibroblasts.84

But how does a mechanical force result in these phenotypic

changes? This leads us to mechanotransduction.

2.4.2 Mechanotransduction

The ability of the cell to maintain its integrity when subjected to mechanical forces is

considerably due to the cytoskeleton’s viscoelastic nature. Mechanical loads are

reciprocally displaced through the cytoskeleton and across the cell membrane into the

ECM via integral mechanoreceptors known as “integrins”. Integrins function as load

elements adhering to actin cytoskeletal elements at focal points on the membrane

which connect to the cell’s nucleus.96-98

The ECM-integrin-cytoskeletal pathway is

currently the most researched and understood of the mechanotransduction pathways,84

which connects to cellular components such as: G-proteins; cadherin complexes;

GTPases; mechanical-gated ion channels; protein kinases; and transcription factors

etc., thus converting physical stimuli into biochemical signals.

It has been demonstrated experimentally that when integrins are mechanically

deformed, the cellular response is to produce a localised intracellular biochemical

transduction, upregulate and recruit focal adhesion proteins to the point of stress and

re-orientate the cytoskeletal architecture to withstand additional stress.99

Kaunas, et al.

Page 47: Design and Development of Cell Stretching Platforms for ...

46

(2005) reported the effect of cyclic mechanical stretch and signalling pathways on

orientation of Bovine aortic endothelial cells (BAECs) stress fibers.100

The stress fibers

align in the direction a stretch in the absences of Rho activity. However, in presence of

intact Rho signalling pathway, stress fibers orients perpendicular to the stretching

direction.

Thereby, the application of an external stress will directly affect the morphology and

sensitively of the focal adhesions for the characteristics of that particular stress,

subsequently creating a more efficient biochemical modulation for intracellular signal

transduction.8 For example, when endothelial cells were exposed to a mechanical

stress, similar to the fluid shear forces found within a blood vessel, the cells elongated

and the focal adhesions along with the cytoskeleton orientated themselves

perpendicular to this stress.101, 102

In this case, a subset of the GTPase proteins played a

pivotal role in the signal transduction pathway and thus facilitated the production of

focal adhesion complexes. Weng et al. (2016) showed that cell mechanical

homeostasis in fibroblasts was dependent upon subcellular (rheostasis) cytoskeletal

tension and focal adhesions,103

while Shao et al. (2014) reported that a novel

mechanotransduction process in endothelial cells was influenced by cell shape via

cytoskeleton and focal adhesions under cyclic stretching. 104

Different stretch

magnitudes and frequency by the pulsating vessel will result in different signalling

properties by endothelial cells.105

Therefore, the actual mechanical forces themselves

are necessary to maintain a proper functioning of the vascular system and irregularities

can cause inappropriate cellular activities and consequential cardiovascular

abnormalities.98, 106-108

The highly localised arrangement of focal adhesions within the

cytoskeletal projections (cilia) of hair cells within the vestibular apparatus, respond

ideally to ECM alterations to fluid dynamics, although in this case the focal adhesions

are linked to mechanical ion gates.9, 109

Many in vitro studies have reported the

different effects strain has on a monoculture. However, due to the complex in vivo

biological environment, with respect to mechanical forces in such systems as the

respiratory system organ-on-a-chip research has become an exciting contemporary

avenue of research.

Page 48: Design and Development of Cell Stretching Platforms for ...

47

Boccafoschi, et al. (2007) observed the effect of different magnitude (1-25%) and

frequency (0.25-3 Hz) of cyclic stretch on human lung fibroblast.110

Results showed

that at a magnitude of 1%, fibroblast aligned to stretch direction, but at and 2%

achieved the highest orientation to strain whilst no significant alteration was seen from

5-20%. Also it was observed that the frequency change did not influence the

percentage of cells oriented perpendicular to the stretch direction. Further increase in

strain at 25% resulted in cellular death. Recently, Cui, et al. (2015) showed that 1-5%

cyclic stretching at frequencies of 0.01-10 Hz increased spreading and stress fiber

formation of fibroblasts on soft substrates (k=2.3 nN/μm).9 It was also found that these

stretched-induced cellular behaviours were linked to biochemical responses by two

related transcription factors, MRTF-A (myocardin-related transcription factor-A) and

YAP (Yes-associated protein). The lung-on-a-chip device has an ideal platform for

mechanotransduction studies. The platform is a microfluidic device consisting of two

micro-channels separated by 10 um PDMS membrane. Epithelial and endothelial cells

can be seeded separately in microchannel and cultured on ECM coated pours PDMS

membrane stretched using pneumatic actuation to mimics the alveolar capillary

interface.28

2.5 Conclusions and perspective

In this chapter, we first discussed the possible actuation concepts for cell stretching

devices. Pneumatic actuation is the most common concept that relies on external

vacuum or pressure to stretch a flexible membrane with cells cultured on it. The main

advantage of pneumatic actuation is that it induces a homogenous mechanical strain to

the cells whilst the device only needs tubing to connect the actuation chambers on the

device to the external vacuum/pressure supply, and therefore does not interfere with

the cell culturing process. Piezoelectric actuation is another viable method as it does

not induce heat although due to the need of a relatively high voltage, piezoelectric

actuation is not suitable for an integrated solution where the actuators may be in

contact with wet and conducting medium. Electromagnetic actuators such as

servomotor or stepper motor can be precisely controlled, however, the ongoing need

for lubrication and the possibility for device erosion are major concerns for cell

contamination.

Page 49: Design and Development of Cell Stretching Platforms for ...

48

Besides these three major actuation concepts, other actuation schemes such

electrothermal, electrostatic, optical and dielectrophoretic concepts can all be utilised

for designing cell stretching devices. Electrothermal and electrostatic actuators could

be implemented in the same way as piezoelectric and electromagnetic actuators. Due to

the heat and high voltage involved, these actuators have to be placed outside of the cell

culture chamber, e.g. coupled mechanically with a linkage mechanism. Optical and

diectrophoretic actuations produce relatively small forces and are therefore suitable for

in-situ single-cell stretching. Monitoring and controlling single-cell stretching is a big

challenge as the culturing condition has to be met at the same time. Furthermore, these

actuation concepts may induce optical and electrostatic stimuli to the cells, which

cannot be separated from mechanical stretching and make designing

mechanotransduction experiments difficult.

Both state-of-the-art commercially available and experimental stretching devices are

limited to simple induction of mechanical stretching to the cells. Both microfluidics

and MEMS technologies have not been fully utilised for making these devices yet.

Following are a few directions for their improvement to make cell-stretching devices

more versatile and enabling tools.

First, most devices reported in the literature are only able to induce one strain rate in a

single experiment. To improve the throughput, the stretching device should have the

same format as a standard well plate array with a range of programmable strain rate.

An 35 array of pneumatic cell stretching devices with three different strain rates have

been demonstrated before.20

In the future, the same concept could be scaled up to suit

the standard format of 96-well, 384-well and 1536-well microplates.

Second, all reported devices relied on the geometry and applied force/pressure to

generate a fixed strain pattern on the cells. As the response to mechanical stimuli may

help cells to differentiate and to form functional tissue, a programmable two-

dimensional or possibly three-dimensional strain map could be designed to create

different tissue type on the same stretching device. The programmability of the strain

Page 50: Design and Development of Cell Stretching Platforms for ...

49

distribution would allow for a new method of bottom-up tissue printing where cells

differentiate themselves based on mechanical cues.

Third, most of the reported devices used manual processes for seeding of cells and

delivering and exchanging culture media. Although most stretching devices based on

microfluidics have the capability to manipulate cells and liquids in microchannels,

none of them actually have pumping and valving integrated for automated handling of

cells and culture medium. In the future, if the handling process can be automated, cell-

stretching devices can be set up and used at a low cost and minimum manual labor.

Finally, the complexity of cell stretching devices may be enhanced with an additional

microfluidic network for creating different concentrations of chemicals such as growth

factors. The ability to induce both mechanical and chemical stimuli to the cells would

bring the capability of these devices to the next level. Cell-stretching devices with

these improvements will enable us to understand better the important processes of

many diseases associated with defects in mechanotransduction.

Page 51: Design and Development of Cell Stretching Platforms for ...

50

2.6 References

1. B. D. Hoffman and J. C. Crocker, Annu. Rev. Biomed. Eng., 2009, 11, 259-288.

2. C. L. Happe and A. J. Engler, Circ. Res., 2016, 118, 296-310.

3. M. Aragona, T. Panciera, A. Manfrin, S. Giulitti, F. Michielin, N. Elvassore, S.

Dupont and S. Piccolo, Cell, 2013, 154, 1047-1059.

4. R. P. Brandes, N. Weissmann and K. Schroder, Antioxid. Redox Signal., 2014, 20,

887-898.

5. D. J. Leong, J. A. Hardin, N. J. Cobelli and H. B. Sun, in Skeletal Biology and

Medicine Ii: Bone and Cartilage Homeostasis and Bone Disease, ed. M. Zaidi,

Blackwell Science Publ, Oxford, 2011, vol. 1240, pp. 32-37.

6. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

7. J. H. C. Wang, J. Biomech., 2006, 39, 1563-1582.

8. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

9. D. E. Ingber, Ann Med, 2003, 35, 564-577.

10. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6, 6333.

11. A. E. Ekpenyong, N. Toepfner, E. R. Chilvers and J. Guck, Biochim. Biophys. Acta,

2015, 1853, 3105-3116.

12. X. Trepat, L. Deng, S. S. An, D. Navajas, D. J. Tschumperlin, W. T. Gerthoffer, J. P.

Butler and J. J. Fredberg, Nature, 2007, 447, 592-595.

13. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

14. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

15. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

16. Y. L. Sun, Z. P. Luo, A. Fertala and K. N. An, J. Biomech., 2004, 37, 1665-1669.

17. O. Chaudhuri, S. H. Parekh, W. A. Lam and D. A. Fletcher, Nat. Methods, 2009, 6,

383-387.

18. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog. Biophys. Mol. Biol., 2014, 115, 93-102.

Page 52: Design and Development of Cell Stretching Platforms for ...

51

19. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

20. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

21. Y. Huang and N. T. Nguyen, Biomed. Microdevices, 2013, 15, 1043-1054.

22. K. B. Roth, K. B. Neeves, J. Squier and D. W. M. Marr, Biomed. Opt. Express, 2015,

6, 807.

23. R. Fior, S. Maggiolino, M. Lazzarino and O. Sbaizero, Microsys. Technol., 2011, 17,

1581-1587.

24. C. P. Ursekar, S. K. Teo, H. Hirata, I. Harada, K. H. Chiam and Y. Sawada, PLoS

One, 2014, 9, e90665.

25. Y. Sei, K. Justus, P. LeDuc and Y. Kim, Microfluid. Nanofluid., 2014, 16, 907-920.

26. J. D. Caplin, N. G. Granados, M. R. James, R. Montazami and N. Hashemi, Adv

Healthc Mater, 2015, 4, 1426-1450.

27. B. Jiang, W. Zheng, W. Zhang and X. Jiang, Sci China Chem, 2013, 57, 356-364.

28. D. Huh, B. D. Matthews, A. Mammoto, M. Montoya-Zavala, H. Y. Hsin and D. E.

Ingber, Science, 2010, 328, 1662-1668.

29. H. J. Kim, D. Huh, G. Hamilton and D. E. Ingber, Lab Chip, 2012, 12, 2165-2174.

30. W. Zheng, B. Jiang, D. Wang, W. Zhang, Z. Wang and X. Jiang, LChip, 2012, 12,

3441.

31. T. D. Brown, J. Biomech., 2000, 33, 3-14.

32. D. Desmaële, M. Boukallel and S. Régnier, J. Biomech., 2011, 44, 1433-1446.

33. B. D. Riehl, J. H. Park, I. K. Kwon and J. Y. Lim, Tissue Eng Part B Rev, 2012, 18,

288-300.

34. C. Moraes, G. Mehta, S. C. Lesher-Perez and S. Takayama, Ann. Biomed. Eng., 2011,

40, 1211-1227.

35. J. M. Mann, R. H. Lam, S. Weng, Y. Sun and J. Fu, Lab Chip, 2012, 12, 731-740.

36. C. Moraes, J. H. Chen, Y. Sun and C. A. Simmons, Lab Chip, 2010, 10, 227-234.

37. C. S. Simmons, J. Y. Sim, P. Baechtold, A. Gonzalez, C. Chung, N. Borghi and B. L.

Pruitt, J Micromech Microeng, 2011, 21, 54016-54025.

38. K. Shimizu, A. Shunori, K. Morimoto, M. Hashida and S. Konishi, Sensors Actuators

B: Chem., 2011, 156, 486-493.

39. Y. J. Heo, E. Iwase, I. Shimoyama, T. Kan and K. Matsumoto, Micro &amp; Nano

Letters, 2013, 8, 865-868.

Page 53: Design and Development of Cell Stretching Platforms for ...

52

40. D. Tremblay, S. Chagnon-Lessard, M. Mirzaei, A. E. Pelling and M. Godin,

Biotechnol. Lett, 2014, 36, 657-665.

41. J. Kreutzer, L. Ikonen, J. Hirvonen, M. Pekkanen-Mattila, K. Aalto-Setala and P.

Kallio, Med. Eng. Phys., 2014, 36, 496-501.

42. B. P. Chan, C. H. Li, K. L. Au-Yeung, K. Y. Sze and A. H. Ngan, Ann. Biomed. Eng.,

2008, 36, 1254-1267.

43. P. Fernández, P. A. Pullarkat and A. Ott, Biophys. J., 2006, 90, 3796-3805.

44. E. A. Peeters, C. W. Oomens, C. V. Bouten, D. L. Bader and F. P. Baaijens, J

Biomech, 2005, 38, 1685-1693.

45. D. B. Serrell, J. Law, A. J. Slifka, R. L. Mahajan and D. S. Finch, Biomed.

Microdevices, 2008, 10, 883-889.

46. Y. Kamotani, T. Bersano-Begey, N. Kato, Y. C. Tung, D. Huh, J. W. Song and S.

Takayama, Biomaterials, 2008, 29, 2646-2655.

47. S. Deguchi, S. Kudo, T. S. Matsui, W. Huang and M. Sato, AIP adv., 2015, 5, 067110.

48. K. Sato, S. Kamada and K. Minami, IJMS, 2010, 52, 251-256.

49. S. P. Arold, J. Y. Wong and B. Suki, Ann. Biomed. Eng., 2007, 35, 1156-1164.

50. W. W. Ahmed, T. Wolfram, A. M. Goldyn, K. Bruellhoff, B. A. Rioja, M. Moller, J.

P. Spatz, T. A. Saif, J. Groll and R. Kemkemer, Biomaterials, 2010, 31, 250-258.

51. C. Neidlinger-Wilke, H. J. Wilke and L. Claes, J. Orthop. Res., 1994, 12, 70-78.

52. L. Huang, P. S. Mathieu and B. P. Helmke, Ann. Biomed. Eng., 2010, 38, 1728-1740.

53. Y. J. Chang, C. J. Tsai, F. G. Tseng, T. J. Chen and T. W. Wang, Nanomedicine, 2013,

9, 345-355.

54. Y. Shao, X. Tan, R. Novitski, M. Muqaddam, P. List, L. Williamson, J. Fu and A. P.

Liu, Rev. Sci. Instrum., 2013, 84, 114304.

55. N. Bellini, F. Bragheri, I. Cristiani, J. Guck, R. Osellame and G. Whyte, Biomed Opt

Express, 2012, 3, 2658-2668.

56. G. Jochen, A. Revathi, C. C. Cunningham and K. Josef, J. Phys.: Condens. Matter,

2002, 14, 4843.

57. I. Sraj, A. C. Szatmary, S. A. Desai, D. W. Marr and C. D. Eggleton, Phys. Rev. E

Stat. Nonlin. Soft Matter Phys., 2012, 85, 041923.

58. F. Lautenschlager, S. Paschke, S. Schinkinger, A. Bruel, M. Beil and J. Guck, Proc

Natl Acad Sci, 2009, 106, 15696-15701.

Page 54: Design and Development of Cell Stretching Platforms for ...

53

59. L. Boyde, A. Ekpenyong, G. Whyte and J. Guck, Appl. Opt., 2012, 51, 7934-7944.

60. C.-W. Lai, S.-K. Hsiung, C.-L. Yeh, A. Chiou and G.-B. Lee, Sensors Actuators B:

Chem., 2008, 135, 388-397.

61. I. Sraj, C. D. Eggleton, R. Jimenez, E. Hoover, J. Squier, J. Chichester and D. W. M.

Marr, JBO, 2010, 15, 047010.

62. G. Nava, F. Bragheri, T. Yang, P. Minzioni, R. Osellame, I. Cristiani and K. Berg-

Sørensen, Microfluid. Nanofluid., 2015, 19, 837-844.

63. G. F. Christopher, J. M. Yoo, N. Dagalakis, S. D. Hudson and K. B. Migler, Lab Chip,

2010, 10, 2749-2757.

64. N. M. Elman, B. C. Masi, M. J. Cima and R. Langer, Lab Chip, 2010, 10, 2796-2804.

65. Y. Zhu, A. Corigliano and H. D. Espinosa, JMiMi, 2006, 16, 242-253.

66. Y. Iwadate and S. Yumura, Biotechniques, 2009, 47, 757-767.

67. I. Guido, C. Xiong, M. S. Jaeger and C. Duschl, Microelectron. Eng., 2012, 97, 379-

382.

68. N. Scuor, P. Gallina, H. V. Panchawagh, R. L. Mahajan, O. Sbaizero and V. Sergo,

Biomed. Microdevices, 2006, 8, 239-246.

69. Z. L. Shen, M. R. Dodge, H. Kahn, R. Ballarini and S. J. Eppell, Biophys J, 2008, 95,

3956-3963.

70. Q. Wang, X. Zhang and Y. Zhao, Sensors Actuators B: Chem., 2014, 194, 484-491.

71. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6.

72. M. Koike, H. Shimokawa, Z. Kanno, K. Ohya and K. Soma, J. Bone Miner. Metab.,

23, 219-225.

73. N. Haghighipour, M. Tafazzoli-Shadpour, M. A. Shokrgozar and S. Amini, Artif.

Organs, 2010, 34, 481-490.

74. E. A. Vogler, Biomaterials, 2012, 33, 1201-1237.

75. D. S. Gray, J. Tien and C. S. Chen, J. Biomed. Mater. Res. A, 2003, 66, 605-614.

76. A. Tondon and R. Kaunas, PLoS One, 2014, 9, e89592.

77. R. N. Palchesko, L. Zhang, Y. Sun and A. W. Feinberg, PLoS One, 2012, 7, e51499.

78. K. Kurpinski, J. Chu, C. Hashi and S. Li, Proc. Natl. Acad. Sci. U. S. A., 2006, 103,

16095-16100.

Page 55: Design and Development of Cell Stretching Platforms for ...

54

79. C. Tamiello, A. B. C. Buskermolen, F. P. T. Baaijens, J. L. V. Broers and C. V. C.

Bouten, Cell. Mol. Bioeng., 2016, 9, 12-37.

80. B. Liu, M. J. Qu, K. R. Qin, H. Li, Z. K. Li, B. R. Shen and Z. L. Jiang, Biophys J,

2008, 94, 1497-1507.

81. A. Livne, E. Bouchbinder and B. Geiger, Nat. Commun., 2014, 5, 3938.

82. Y. Chen, A. M. Pasapera, A. P. Koretsky and C. M. Waterman, Proc Natl Acad Sci U

S A, 2013, 110, E2352-2361.

83. S. M. Gopalan, C. Flaim, S. N. Bhatia, M. Hoshijima, R. Knoell, K. R. Chien, J. H.

Omens and A. D. McCulloch, Biotechnol. Bioeng., 2003, 81, 578-587.

84. J. H. Wang and B. P. Thampatty, Biomechanics and modeling in mechanobiology,

2006, 5, 1-16.

85. X. Trepat, L. Deng, S. S. An, D. Navajas, D. J. Tschumperlin, W. T. Gerthoffer, J. P.

Butler and J. J. Fredberg, Nature, 2007, 447, 592-595.

86. M. Yoshigi, E. B. Clark and H. J. Yost, Cytometry Part A, 2003, 55A, 109-118.

87. Y. Chen, A. M. Pasapera, A. P. Koretsky and C. M. Waterman, Proc. Natl. Acad. Sci.

U. S. A., 2013, 110, E2352-E2361.

88. N. F. Jufri, A. Mohamedali, A. Avolio and M. S. Baker, Vasc Cell., 2015, 7, 8.

89. N. von Offenberg Sweeney, P. M. Cummins, Y. A. Birney, J. P. Cullen, E. M.

Redmond and P. A. Cahill, Cardiovasc. Res., 2004, 63, 625-634.

90. R. L. Smith, S. F. Rusk, B. E. Ellison, P. Wessells, K. Tsuchiya, D. R. Carter, W. E.

Caler, L. J. Sandell and D. J. Schurman, J Orthop Res, 1996, 14, 53-60.

91. D. Kaspar, W. Seidl, C. Neidlinger-Wilke, A. Beck, L. Claes and A. Ignatius, J

Biomech, 2002, 35, 873-880.

92. R. L. Duncan and C. H. Turner, Calcif Tissue Int, 1995, 57, 344-358.

93. C. A. Powell, B. L. Smiley, J. Mills and H. H. Vandenburgh, Am J Physiol Cell

Physiol, 2002, 283, C1557-1565.

94. H.-S. Yu, J.-J. Kim, H.-W. Kim, M. P. Lewis and I. Wall, J. Tissue Eng., 2015, 6,

2041731415618342.

95. P. F. Davies, Physiol Rev, 1995, 75, 519-560.

96. C. Tamiello, A. B. Buskermolen, F. P. Baaijens, J. L. Broers and C. V. Bouten, Cell

Mol Bioeng, 2016, 9, 12-37.

Page 56: Design and Development of Cell Stretching Platforms for ...

55

97. X.-m. Liu, K. J. Peyton, A. R. Shebib, H. Wang and W. Durante, Biochem.

Pharmacol., 2011, 82, 371-379.

98. P. S. Howard, U. Kucich, R. Taliwal and J. M. Korostoff, J. Periodontal Res., 1998,

33, 500-508.

99. N. Wang, J. P. Butler and D. E. Ingber, Science, 1993, 260, 1124-1127.

100. R. Kaunas, P. Nguyen, S. Usami and S. Chien, Proc. Natl. Acad. Sci. U. S. A., 2005,

102, 15895-15900.

101. N. F. Jufri, A. Mohamedali, A. Avolio and M. S. Baker, Vasc Cell, 2015, 7, 8.

102. A. M. Collinsworth, C. E. Torgan, S. N. Nagda, R. J. Rajalingam, W. E. Kraus and G.

A. Truskey, Cell Tissue Res, 2000, 302, 243-251.

103. S. Weng, Y. Shao, W. Chen and J. Fu, Nat. Mater., 2016, DOI: 10.1038/nmat4654.

104. Y. Shao, J. M. Mann, W. Chen and J. Fu, Integr. Biol. (Camb.), 2014, 6, 300-311.

105. C. L. Happe and A. J. Engler, Circ. Res., 2016, 118, 296-310.

106. I. S. Joung, M. N. Iwamoto, Y. T. Shiu and C. T. Quam, Microvasc Res, 2006, 71, 1-

11.

107. Z. Hu, Y. Xiong, X. Han, C. Geng, B. Jiang, Y. Huo and J. Luo, PLoS One, 2013, 8,

e71359.

108. H. Lal, S. K. Verma, H. B. Golden, D. M. Foster, A. M. Holt and D. E. Dostal,

Molecular Signaling Mechanisms of Myocardial Stretch: Implications for Heart

Disease, Springer, Dordrecht, 2010.

109. W. Zheng, B. Jiang, D. Wang, W. Zhang, Z. Wang and X. Jiang, Lab Chip, 2012, 12,

3441-3450.

110. F. Boccafoschi, M. Bosetti, S. Gatti and M. Cannas, Cell adhesion & migration, 2007,

1, 124-128.

Page 57: Design and Development of Cell Stretching Platforms for ...

56

Chapter 3: An electromagnetic cell-stretching device for

mechanotransduction studies of olfactory ensheathing cells

First step in establishing systematic approach for mechanobiological investigation is to

understand the strain requirement for the cells under investigation. Indeed, various cell

stretching approaches have been developed to induce homogeneous strain onto the cell

culture and establish a relationship between the strain amplitude with cellular response.

However, an ability to observe this cellular response under the same biological

conditions for different strain amplitudes within the same field of view would be an

added advantage. Thus, this Chapter focuses on the development of the novel cell

stretching platform with heterogeneous strain pattern to facilitate a better comparison

of the cellular response to different strain amplitudes in the same field of view. The

study successfully developed a well characterised electromagnetic cell-stretching

platform using the single-sided uniaxial stretching approach. The platform was capable

of inducing heterogeneous strain onto the cell culture and facilitated the comparison of

the cellular response towards different strains in the same field of view. The platform

was tested with olfactory ensheathing cells (OECs) to observe their morphological

changes for different strain amplitudes and to determine the strain requirement for

OECs.

Page 58: Design and Development of Cell Stretching Platforms for ...

57

Statement of Contribution to Co-authored Published Paper

Page 59: Design and Development of Cell Stretching Platforms for ...

58

3.1 Introduction

Cells in the human body are under constant influence from the various mechanical forces

place upon them. The conversion of mechanical stimuli such as rigidity and stretching into

intracellular biomolecular signaling is known as a mechanotransduction 1, 2

. Mechanical force

is well known to be one of the most significant stimuli to influence cell growth 3, behaviour

and morphology 4. Furthermore, abnormalities in mechanotransduction have been reported to

lead to diseases such as asthma, heart failure and cancer 5, 6

. Thus, the development of

techniques to best mimic the cells’ physical microenvironment has been a major interest

within the research community of mechanotransduction. Nevertheless, the complexity of

mechanobiology associated with the in-vivo environment has led to the development of in-

vitro systems. Cell stretching is the most common approach through the use of clinical tools

such as micropipettes or tweezers 7-11

. However, these techniques are not suitable for long-

term experiments. Flexcell (Flexcell International Corporation), Strex Systems for cell

Stretching (STREX Inc.) and ElectroForce (Bose Corporation) are the few commercially

available cell stretching systems. Flexcell is one of the most sophisticated commercially

available cell stretching devices 12-14

and the BioFlex (Flexcell International Corporation)

system uses a pneumatic control to inflate or deflate elastic membrane for stretching cultured

cells.

Several other custom-made devices reported in the literature utilize other approaches, such as:

electromagnetic, electrostatic or optical actuation techniques. A common approach is

mechanically pulling an elastic membrane using computer controlled motors 15-17

. Illustrated

by Shao, et al. (2013) who used a controlled motor to drive a cam follower assembly, that was

connected to a flexible polydimethylsiloxane (PDMS) membrane for cyclic cell stretching

[17]. Micro-electromechanical systems (MEMS) integrated microactuators have also been

widely used for cell stretching 19, 20

, with Deguchi, et al. (2015) utilizing piezoelectric

actuators to stretch the PDMS device that has integrated microchambers and membranes for

cell attachment [20]. Another common technique for cell stretching is pneumatic actuation to

deform a thin PDMS membrane 22-25

, Huang and Nguyen (2013) fabricated a multilayered

PDMS device and used a controlled vacuum source to achieve uniaxial deformation of cells

cultured on a flexible membrane [25]. However, pneumatically actuated systems may lead to

uncertain experimental results due to air leakage and increased gas permeability into the

Page 60: Design and Development of Cell Stretching Platforms for ...

59

membrane. Finally, optical trapping, shape memory alloy (SMA) actuation 27

and

dielectrophoresis 28

, are other methods which have their own advantages, but these devices

rarely fulfill the general requirements of cell stretching such as a high experimental

throughput, well-characterised strain pattern, ease of operation, compatible with a wide range

of imaging systems and biocompatibility with the cells under investigation.

The effect of strain on cell morphology and behaviour may potentially lead to improved

culture conditions of olfactory ensheathing cells (OECs). In this study, OECs was of interest

and the system was designed to understand the primary effect of stretching on OEC

morphology and physiology. OECs are the glia cells of the primary olfactory system and are

located in the olfactory nerves within the nasal cavity and the outer layer of the olfactory bulb.

OECs provide support and maintenance for the olfactory sensory neurons and guide the

olfactory sensory axons to their targets within the olfactory bulb 29

. Primary sensory neurons

project their dendrites to the surface of the olfactory epithelium and are directly exposed to

external toxic substances and pathogens and as result frequently die. However, neural stem

cells in the basal layer of the olfactory epithelium give rise to new neurons and the OECs

release numerous growth factors which promote the growth of newly generated axons. The

ability of OECs to aid facilitate axon growth 30

and integrate peripheral sensory axons into the

PDMS Device

Membrane

Electromagnet

Permanent magnet

(b) OFF state Top view

Actuation axis

Stretched membrane

Deformedfront wall

PDMS device

Permanent Magnet

Electromagnet

Air vent

Mountingplatform

Power SupplyConnectors

Holder

Membrane

a)(

c) ON state Top view

Constrained back wall

(

8mm

Fig. 3.1 The cell-stretching device: (a) the actual device; (b) the device in the relaxed state; (c) the device in the stretched state.

Page 61: Design and Development of Cell Stretching Platforms for ...

60

olfactory bulb of the central nervous system make them prime candidates for spinal cord

injury transplantation therapies 31

. A recent human trial has shown the potential of OEC

transplantation in therapeutic and regenerative medicine for spinal cord injury 32-34

. Although

the results were promising, the complexities for post transplantation survival of cells have

been a major concern. Thus, better understanding of the OEC in-vivo physical environment

holds the key to obtaining better post transplantation survival results using OECs.

In-vitro mechanotransduction studies can serve as a tool to better understand the functionality

of OECs in their physical environment. We developed a cell stretching system that provides a

heterogeneous strain pattern over a deformable membrane. The heterogeneous strain pattern

obtained was utilized to observe the effect of predictable strain on OECs, and to determine the

optimal strain requirement for the cell growth. The present chapter reports a novel cell-

stretching device based on the single sided uniaxial stretching approach using magnetic

actuation. We present the mechanical design, fabrication, simulation and experimental

characterisation of the device. Alongside preliminary cell culture results which will be utilized

for further optimisation of the design and the control of the device. The device allows for the

active manipulation of growth and orientation of OECs in culture.

3.2 Device concept and design

A single sided uniaxial stretching device was designed for application of strain to cultured

OECs and at the same time allowing microscope access for live imaging. Figure 3.1 depicts

the device concept and the actual cell stretching system. This system provided advantages

such as ease of fabrication, replication and a heterogeneous strain pattern compared to

existing homogenous cell stretching devices. The strain pattern also allows for a better

comparison of cell response towards different strains in the same field of view which is an

added advantage.

Figure 3.1a depicts the actual system for OEC stretching. The PDMS device incorporates an

NdFeB permanent disc magnet with a diameter of 15 mm and a thickness of 2 mm [JL

Magnet, Seoul, Korea] embedded into the front wall of the PDMS. The deformable membrane

with a thickness of 200 µm was bonded at the bottom of the device after a treatment with

oxygen plasma. The electromagnet [JL Magnet, Seoul, Korea] is controlled by a

programmable DC power supply [MK Power, Seoul, Korea] and serves as the actuator in the

Page 62: Design and Development of Cell Stretching Platforms for ...

61

system. The mounting platform provides a necessary constraint to the PDMS device and also

the axial alignment of the permanent magnet and the electromagnet (Fig. 3.1b). The magnetic

force generated by the embedded magnet and the electromagnet pulls the front wall, while the

back wall is constrained (Fig. 3.1c). The unconstrained front wall is deformed as illustrated in

Figure 3.1c. As the membrane is bonded to the bottom of the device, the membrane is

stretched uniaxially in the direction of the force. OECs cultured on the membrane are

consequently stretched with the induced strain.

3.2.1 Characterisation of the permanent magnet and the electromagnet

The magnetic fields of the NdFeB disc magnet and the electromagnet were first characterised

to determine the operation parameters for the device. Moreover, to design the mounting

platform it was necessary to determine the acceptable minimum distance between the magnet

and electromagnet in the off-state, which does not create any visible attraction between the

permanent magnet and electromagnet. A gauss meter (Hirst Magnetic Instrument Ltd.) was

used to measure the magnetic flux density as a function of distance and voltage for the

permanent magnet and the electromagnet, respectively.

0 1 2 3 440

60

80

100

120

140

160

180

200

Distance(cm)

Magn

etic

fie

ld (m

T)

0 5 10 15 20 25 30 350

10

20

30

40

50

60

70

80

90

Voltage(V)

2m

m

15mm

40mm

25

mm

Body Coil

Core

(a) (b)

Current CoilCore

Actuationaxis

Actuationaxis

Fig. 3.2 Measured magnetic flux density: (a) as function of a distance from the

permanent magnet; (b) as function of the voltage applied on the electromagnet.

Page 63: Design and Development of Cell Stretching Platforms for ...

62

The permanent magnet was mounted onto the linear stage of a syringe pump; the gauss meter

was placed perpendicular to the center axis of the permanent magnet. Reading for the

magnetic flux density was taken for a distance ranging from 0 to 4 cm with a step of 0.1mm

(Fig. 3.2a). The results of this characterisation experiment provided a guideline for the

acceptable distance between the magnet and the electromagnet, 5 mm was chosen for

designing the mounting platform. Furthermore, the same experimental setup was chosen for

the characterisation of the electromagnet with a diameter of 25 mm and height of 40 mm.

The gauss meter was placed at a distance of 5 mm as selected from the characterisation of the

permanent magnet. Readings were taken for the magnetic flux density with a voltage range

of 1 to 30V. Figure 3.2b shows that the flux density increases almost linearly with increasing

voltage. A driving voltage up to 15 V can provide a field strength up to 50 mT that can

control the movement of the permanent magnet placed 5 mm apart.

8 10 12 14 16 18 20

2

4

6

8

10

12

Location Magnet Z axis (mm)

Ave

rage

Dis

pla

ce

me

nt

of

me

mb

ran

e (

mm

)

Force

Strain

Constrained back wall

Front WallSide wall

Membrane

Optimised Permanent

magnet position

Electromagnet

Actuation axis

Z

YX(0 0 0)

O

Fig. 3.3 Simulated results for location of actuation axis along Z axis verses average

displacement of membrane. (Inset: Optimised geometry and simulation results of FEA

model.)

Page 64: Design and Development of Cell Stretching Platforms for ...

63

3.2.2 Design and fabrication

In the conceptual design phase, a

simple Finite element analysis

(FEA) model of PDMS device was

created in COMSOL Multiphysics

4.3b to predict the device behavior

and to optimise the device

geometry. The location of the

actuation axis along the z axis was

the key parameter of the

optimisation. Reference FEA model

geometry with a total volume of

30x20x28 mm3 was built with a

front wall thickness of 5 mm,

constrained wall thickness of 8mm,

side wall thickness of 2 mm,

membrane thickness of 0.2 mm and

the permanent magnet embedded in

the font wall at 15 mm and 2.5 mm

along X and Y axis from origin O.

Figure 3.3 shows the geometry of

the FEA model. Boundary

conditions were implemented to

constrain the back wall and all other

parts were made free. An outward

force of 0.5 N was applied to the

surface magnet to represent magnetic force. The device was optimised by varying one

geometry parameter while keeping other parameters constant. Using parametric sweep

function, location of the magnet center along Z axis was varied form 8 mm to 20 mm

with step of 0.5 mm over front wall. Figure. 3.3 shows the location of the permanent

Input : Reference Image,

Pixel length Frame/secImage Sequence

Threshold and Edge detection

Dilation operation for particle detection

Noise removal to enhance

particle marking

Mark detectable edges

Convert Image into RGB and make

marked

edges green

Generate grid over reference image

Take input from user for

Region of interest (ROI)

Calculate offset from centre for

selected ROI with reference

image

For Image Sequence

Calculate Displacement

and Strain

Display results

ParticleDisplacment

8m

m

Fig. 3.4 Flow chart for a developed algorithm to

calculate displacement and strain.

Page 65: Design and Development of Cell Stretching Platforms for ...

64

magnet along Z axis verses average displacement of membrane. The parametric study

provided optimised permanent magnet position at 8.5 mm along Z axis.

The fabrication process of the cell stretching system consisted of three noteworthy steps: 1) a

master mold was fabricated with aluminum by a milling machine (CSI Tech, Pyeongtaek,

Korea). 2) the mounting platform and holder was made of plastics using a 3D printer

(Former’s farm, Pusan, Korea) and Autodesk Inventor Professional 2015 Student Edition. 3)

the PDMS device and PDMS membrane were made of PDMS (Sylgard 184 elastomer kit,

Dow Corning). Fifteen gram PDMS mixture was prepared by mixing PDMS prepolymer and

its cross-linker at 30:1 ratio. The mixture was poured into the master mold and degassed

again for 15 minutes to remove air bubbles. The master mold was closed carefully with a

cylindrical cap on mold facing inside to create a cavity for the magnet placement. After

curing the master mold for two h at 80 °C, the mold was opened and the permanent magnet

was placed onto the cavity created by the cylindrical cap. To ensure the placement 10:1

PDMS and cross linker mixture was used as glue and pasted on magnet. After ensuring

placement of the magnet the master mold was closed and again cured at 80 °C for 30

minutes. Replicated PDMS devices were then inspected and carefully peeled out of master

mold. To prepare the deformable membrane PDMS and cross linker was mixed in 10:1 ratio.

PDMS membrane was spin coated at 400 rpm for 2 min to achieve 0.2 mm membrane

Voltage (V)

0

1

2

3

4

0 2 4 6 8 10 12 14 16

0

0.5

1

1.5

Voltage (V)

Forc

e (N

)D

ispla

cem

ent (m

m)

(b)

(a)

PD

MS

Devic

e

Permeantmagnet

Camera

Electromagnet

PowerSupply

15m

m D

ia

Fig. 3.5. Experimental characterisation of the device: (a) displacement versus voltage;

(b) estimated force versus voltage.

Page 66: Design and Development of Cell Stretching Platforms for ...

65

thickness and cured for 2 h at 80 °C in a vacuum oven. PDMS device and membrane was

cleaned with isopropanol and Deionized (DI) water and plasma bonded for 45 seconds

followed by 1 h curing at 80 °C in a vacuum oven to ensuring proper bonding.

3.3 Device characterisation

3.3.1 Characterisation of magnetic force

A three-dimensional (3D) finite element analysis (FEA) model of PDMS was utilized to

calculate the spring constant (K) of the experimental setup. With the known spring constant,

the force generated by the embedded permanent magnet and the electromagnet was then

estimated. The geometry of the device was replicated in COMSOL Multiphysics 4.3b. The

Young’s modulus for the device material (PDMS mixed with 30:1 volume ratio) and the

membrane material (PDMS mixed with 10:1 volume ratio) was selected as 0.145 MPa and

50 Kpa, respectively 35, 36

. The back wall was fixed with the other parts able to move freely.

The magnet embedded in the wall was considered as the region of interest (ROI) and

displacement of the ROI was obtained by applying outwards force to the surface of the

permanent magnet to mimic the magnetic force acted on it with the electromagnet turned on.

The applied force was varied from 0.1 to 10 N to obtain the corresponding average

displacement. Considering the elastic nature of the PDMS material, displacement obtained

was further utilized by calculating the apparent spring constant K using Hook’s Law;

𝐹 = −𝐾. 𝑥 (1)

Where F is the applied force in N, x is the displacement in m and K is the apparent spring

constant in N/m.

For experimental magnetic force calculation, we chose the same ROI as considered for the

simulation and randomly marked points over ROI. The embedded permanent magnet was

deflected by supplying the electromagnet with a voltage ranging from 1 to 15 V. The

displacement of the ROI was recorded using a digital camera. An algorithm was developed

to calculate the displacement of the marked point from the captured images. The algorithm

was implemented using Digital Image Correlation (DIC) functions and Matlab Image

Processing Toolbox. Figure. 3.4 shows the detailed flowchart and example of the step output

of the algorithm implemented in Matlab R2014a.

Page 67: Design and Development of Cell Stretching Platforms for ...

66

An average surface displacement of ROI was calculated for each input and utilized simulated

average spring constant value (0.339 kN/m) to calculate the total force generated over the

surface based on Hook’s law. Figure. 3.5 shows the experimental results of the relationship

between the displacement as well as the estimated force and the applied voltage, respectively.

For comparative studies, experimentally estimated force (Fig. 3.5b) was further utilized as an

input to a comprehensive simulation model to obtain the corresponding displacement. Figure.

3.6 shows the simulation and experimental results.

3.3.2 Characterisation of stretching membrane

The numerical model was further utilized to optimise the membrane thickness. Figure 3.7

shows the average strain over the membrane as a function of thickness t ranging from 0.04 to

0.4 mm with a step of 0.04 mm. The simulation results provided a guideline for the actual

membrane thickness to be fabricated. Considering fabrication stability and suitability

membrane size was fixed at 0.2 mm.

3.4 Characterisation of strain map

A well characterised strain map on the membrane is important to quantify the strain effect on

OECs in culture. Considering the heterogeneity of the strain map whilst understanding the

Fig.3. 6 Comparison of experimental and simulation force versus displacement results

(Inset: Actual cell stretching system displacement due to actuation)

0 0.5 1 1.50

0.5

1

1.5

2

2.5

3

3.5

4

4.5

Force(N)

Dis

pla

cme

nt(m

m)

Experimental Results

Simulation Results

Mounting Platform

Power SupplyPDMS

DevicePermanentMagnet

Membrane

Electro-magnet

Deformed Front wall

Constrained Back Wall

Strain

Page 68: Design and Development of Cell Stretching Platforms for ...

67

location of the cells with the corresponding strain are crucial for the evaluation of the cell

response. The strain pattern on the membrane was investigated both numerically and

experimentally.

3.4.1 Numerical analysis

The 3D FEA model previously used for design optimisation was further utilized to verify the

experimental results. Experimentally calculated force over operation range was selected as an

input to FEA model to obtain strain over the ROI. For graphical comparison, experimentally

calculated force (0.801N) for the 12V was selected as an input to FEA model to obtain strain

map over the membrane. Figure. 3.8a shows the simulation results. Results displayed an

increasing semi elliptical pattern of strain map in the direction of force over the membrane.

The central region of membrane near the deformable front wall experiences the maximum

strain upon actuation while membrane region near constrained wall experiences the least

strain. It is evident from the results that OECs in culture will experience different magnitude

of strains depending on their location upon the membrane. Thus, morphological and

numerical changes observed in the OECs will mainly depend on their location. For instance,

cells located at maximum strain regions near the front wall are expected to have significant

0.06

0.05

0.04

0.03

0.02

0

0.01

0.5 1

t= 0

.04

mm

t= 0.2 mm

t= 0.4 mm

Force (N)

0

2.5

2

1.5

1

x10-2

Ave

rage

Str

ain

over

mem

bra

ne (

1)

Fig. 3.7 Numerically simulated strain as function of applied force. (Inset: Strain pattern

for membrane with thickness of 0.2 mm)

Page 69: Design and Development of Cell Stretching Platforms for ...

68

morphology changes. While, the region of membrane near the constrained wall will have a

comparable morphology to the control cells.

Thus, cells present at any location on the membrane will experience approximately the same

strain as estimated characterisation results. Also from experimental and simulation results, it

was observed that the curvature of the membrane over the working range is very low

allowing minimum out of plane deformation and thus does not affect real-time imaging

capability of the system.

3.4.2 Experimental analysis

The same algorithm described in force calculation section (Fig. 3.4) was utilized to calculate

offset displacement of the selected particle and calculate Eulerian strain. The membrane was

divided into a 5x12 matrix of image regions. A minimum of three samples were taken from

each region to calculate the strain pattern. Figure. 3.8b shows the experimental results of the

strain field over the deformable membrane with actuation force generated by an applied

voltage of 12 V.

Fig. 3.8 Comparison of Strain map over membrane: (a) simulation; (b) experiment.

Simulation results:

Experimental results:

(a)

(b)

Page 70: Design and Development of Cell Stretching Platforms for ...

69

A relationship between strain over the ROI and voltage is established using same

experimental platform. Figure. 3.9 shows the comparison between simulation and

experimental results, where the relation between strain and voltage is linear. Thus

experimental matches well to simulated results and results show consistency over the

working voltage range of over the 10-15 V range allowing estimation of strain range

over 1% to 2.5%. Applied voltages of 12 V and 15 V yielded an average strain of

1.69% and 2.10 %, respectively.

3.5 Culturing of OECs on the stretching device

3.5.1 Materials and methods

Based on the above characterisation results, we considered the stretching condition with 15V

(ɛavg=2.1%, ɛmax. =2.9%) and a frequency 0.01Hz at 50% duty cycle for the OEC stretching

experiment. For OEC cultures, the device was sterilized using ethanol and dionised water

0 5 10 15

0

0.005

0.01

0.015

0.02

0.025

Voltage (V)

Str

ain

(1)

Permanent Magnet

Electromagnet

Camera

MarkedMembrane

PDMS Device

Power SupplyMounting Platform

Experimental results

Simulation results

Fitted line

8mm

Fig. 3.9. Comparison of voltage versus simulation and experimental strain results at the

central region of membrane. (Inset: Illustration of experimental setup).

Page 71: Design and Development of Cell Stretching Platforms for ...

70

followed by low-intensity UV exposure for 20 minutes. The membrane was then treated with

50-l matrigel (Corning) for at least one hour, to improve OEC adhesion to the membrane.

Immortalised OECs 37

were cultured under standard cell culture condition. DMEM/F12

culture media was enriched with FBS (10% vol/vol), forskolin (2 μM), pituitary extract (20

μg ml-1

), FGF and EGF (10 ng) and gentamycin (5% vol/vol) for the cell culture. All OECs

were cultured in a humidified atmosphere of 5% CO2 at 37 °C. Cells were passaged after

every 2 days at 70-75% of confluence. Seeding for the device was optimised considering

various factors such as strain on cells, imaging capability, homogenous distribution of cells

over the membrane and confluence rate (70-80% after 24 h). A sagging membrane deformity

due to cell culture media was another factor which was a challenge for seeding optimisation.

As cell culture imposed an additional load on the membrane, it affected the strain pattern and

led to inconsistent results. To overcome this problem, after successive experimentation, the

seeding density was optimised as 50,000 cells in a volume of 250 to 500 l of media to

obtain 70-80% confluence rate after 24 h with negligible sacgging of the membrane. A

05 10 15 20 25

20

30

40

50

60

70

Time (Hour)

Avg

. L

eng

th o

f O

EC

over

mem

bra

ne (

um

)

Plateau phase

Exponential Growth phase

4 Hours

8 Hours

12 Hours

16 Hours 24 Hours

20 Hours

Fig. 3.10. Experimental results: Time verses average length of OEC over membrane.

Inset: Morphology of OECs.

Page 72: Design and Development of Cell Stretching Platforms for ...

71

relationship between lengths of the cells over the time was established to evaluate the effect

of the applied strain on the cells. For imaging and evaluation purposes the membrane was

divided into a 3x6 matrix of image regions. Image for each region was taken after ever 4 h.

Obtained images were analyzed using ImageJ, for analysis, each image was subdivided into

5 sub-regions with 20 cells analysed from each region; the experiment was repeated three

times. To calculate the average length of OECs in culture over the membrane, the length for

all cells were measured. Figure.3.10 shows the results for the average length of the OECs

with respect to time. Two distinct phases can be observed (depicted by the dashed vertical

line in Fig.3.10) in the results, OEC growth is exponentially increased during first 8 h and

reaches a plateau phase after 10 h.

3.5.2 Effect of strain on the OEC morphology

Considering the results (Fig.3.10), 8 h cyclic stretching experiment was designed to observe

the OECs’ behavior. To apply strain OECs were seeded onto the membrane with optimised

seeding density (50,000 cells in 250 l of media) and allowed to attach to the membrane up

to 3 h. Then cyclic stretch was applied to the membrane with selected stretching parameters.

The strain was transferred to the OECs and changes in morphology of OECs were observed

and noted. A significant morphological change was observed in the high strain regions after

4 h (Fig.3.11). Figure 3.11 shows the length of OECs versus the degree of strain, along with

OEC morphology after four and 8 h, along with the control (without stretching). The results

show that the length of the OEC decreases linearly with the increase in strain, whilst OECs

retained their morphology even after the stretching was ceased.

From the results, it is evident that the OEC morphology changes depending on the amount of

strain experienced and the time subjected to the strain. Interestingly, it was observed that

OECs retained their changed morphology after stretching. OECs have a plastic morphology

and can rapidly alter their shape in vitro and in vivo 38, 39

, in particular, OECs can extend and

retract branches particularly when interacting with other cells. When co-cultured with

neurons, the axons of the neurons extend along the branches of the OECs and hence

increasing the length of OECs would promote axon extension 30

. Our results demonstrate that

increasing strain results in the length of the OECs decreasing suggesting that they are either

are unable to maintain focal adhesions or that the altering physical relationship with the

neighbouring cells disrupts the cells’ ability to distinguish directional cues from growth-

Page 73: Design and Development of Cell Stretching Platforms for ...

72

promoting factors within the extracellular environment. These results suggest that it is

important to further understand how the changes in cell-cell interactions influence cell

adhesion and growth. For example, after transplantation the constant movement of cells may

decrease the successful integration of cells with the host tissue and thereby results in poorer

therapeutic outcomes.

3.6 Conclusions

A device was designed, developed and characterised to apply known strain to OEC culture

and investigate mechanotransduction behavior. Primary experimentations of OEC culture

shows that the length OECs decreases with increase in strain experienced because of the fact

that either they are unable to maintain their focal adhesion or morphology change has

distrusted OECs ability to distinguished direction. The observations obtained from the

experimental results illustrate the effectiveness of the developed platform and provides

guidelines for the future exploration of the OECs mechanotransduction studies.

00.5 1 1.5 2

20

30

40

50

60

70

80

Avg.

Le

ngth

of O

EC

(u

m)

After 4 Hours of stretching

After 8 Hours of stretching

Strain (1)

Control Stretching Control Stretching

After 8 HoursAfter 4 Hours

OEC Morphology

0.591%

0.937%

2.12%

1.83%

0.591%

2.12%

1.83%

0.937%

Fig.3.11 Experimental results: Strain verses average length of OEC after 4 h and 8 h.

(Inset: Change in Morphology of OEC culture with and without stretching)

Page 74: Design and Development of Cell Stretching Platforms for ...

73

3.7 References

1. R. P. Brandes, N. Weissmann and K. Schroder, Antioxid. Redox Signal., 2014, 20,

887-898.

2. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

3. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6.

4. J. H. C. Wang and B. P. Thampatty, Biomechanics and modeling in mechanobiology,

2006, 5, 1-16.

5. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

6. D. E. Ingber, Ann Med, 2003, 35, 564-577.

7. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

8. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

9. Y. L. Sun, Z. P. Luo, A. Fertala and K. N. An, J. Biomech., 2004, 37, 1665-1669.

10. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

11. M. A. Teitell, S. Kalim, J. Schmit and J. Reed, Biomechanics of Single Cells and Cell

Populations, 2010.

12. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog. Biophys. Mol. Biol., 2014, 115, 93-102.

13. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

14. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

15. S. P. Arold, J. Y. Wong and B. Suki, Ann. Biomed. Eng., 2007, 35, 1156-1164.

16. W. W. Ahmed, T. Wolfram, A. M. Goldyn, K. Bruellhoff, B. A. Rioja, M. Moller, J.

P. Spatz, T. A. Saif, J. Groll and R. Kemkemer, Biomaterials, 2010, 31, 250-258.

17. C. Neidlinger-Wilke, H. J. Wilke and L. Claes, J. Orthop. Res., 1994, 12, 70-78.

18. Y. Shao, X. Tan, R. Novitski, M. Muqaddam, P. List, L. Williamson, J. Fu and A. P.

Liu, Rev. Sci. Instrum., 2013, 84, 114304.

19. Y. Kamotani, T. Bersano-Begey, N. Kato, Y. C. Tung, D. Huh, J. W. Song and S.

Takayama, Biomaterials, 2008, 29, 2646-2655.

Page 75: Design and Development of Cell Stretching Platforms for ...

74

20. K. Sato, S. Kamada and K. Minami, IJMS, 2010, 52, 251-256.

21. S. Deguchi, S. Kudo, T. S. Matsui, W. Huang and M. Sato, AIP adv., 2015, 5, 067110.

22. D. Huh, B. D. Matthews, A. Mammoto, M. Montoya-Zavala, H. Y. Hsin and D. E.

Ingber, Science, 2010, 328, 1662-1668.

23. J. M. Mann, R. H. Lam, S. Weng, Y. Sun and J. Fu, Lab Chip, 2012, 12, 731-740.

24. C. Moraes, J. H. Chen, Y. Sun and C. A. Simmons, Lab Chip, 2010, 10, 227-234.

25. C. S. Simmons, J. Y. Sim, P. Baechtold, A. Gonzalez, C. Chung, N. Borghi and B. L.

Pruitt, J Micromech Microeng, 2011, 21, 54016-54025.

26. Y. Huang and N. T. Nguyen, Biomed. Microdevices, 2013, 15, 1043-1054.

27. Y. Iwadate and S. Yumura, Biotechniques, 2009, 47, 757-767.

28. I. Guido, C. Xiong, M. S. Jaeger and C. Duschl, Microelectron. Eng., 2012, 97, 379-

382.

29. J. A. Ekberg, D. Amaya, A. Mackay-Sim and J. A. St John, Neurosignals, 2012, 20,

147-158.

30. L. C. Windus, F. Chehrehasa, K. E. Lineburg, C. Claxton, A. Mackay-Sim, B. Key

and J. A. St John, Cell. Mol. Life Sci., 2011, 68, 3233-3247.

31. J. A. Ekberg and J. A. St John, Anat Rec (Hoboken), 2014, 297, 121-128.

32. J. A. K. Ekberg and J. A. St John, Neural Regener. Res., 2015, 10, 1395-1396.

33. P. Tabakow, G. Raisman, W. Fortuna, M. Czyz, J. Huber, D. Li, P. Szewczyk, S.

Okurowski, R. Miedzybrodzki, B. Czapiga, B. Salomon, A. Halon, Y. Li, J. Lipiec, A.

Kulczyk and W. Jarmundowicz, Cell Transplant., 2014, 23, 1631-1655.

34. S. Wang, J. Lu, Y. A. Li, H. Zhou, W. F. Ni, X. L. Zhang, S. P. Zhu, B. B. Chen, H.

Xu, X. Y. Wang, J. Xiao, H. Huang, Y. L. Chi and H. Z. Xu, Cell transplantation,

2016, 25, 141-157.

35. D. H. Kim, N. Lu, R. Ma, Y. S. Kim, R. H. Kim, S. Wang, J. Wu, S. M. Won, H. Tao,

A. Islam, K. J. Yu, T. I. Kim, R. Chowdhury, M. Ying, L. Xu, M. Li, H. J. Chung, H.

Keum, M. McCormick, P. Liu, Y. W. Zhang, F. G. Omenetto, Y. Huang, T. Coleman

and J. A. Rogers, Science, 2011, 333, 838-843.

36. R. N. Palchesko, L. Zhang, Y. Sun and A. W. Feinberg, PLoS One, 2012, 7, e51499.

37. J. Tello Velasquez, M. E. Watts, M. Todorovic, L. Nazareth, E. Pastrana, J. Diaz-

Nido, F. Lim, J. A. Ekberg, R. J. Quinn and J. A. St John, PLoS One, 2014, 9,

e111787.

Page 76: Design and Development of Cell Stretching Platforms for ...

75

38. L. C. Windus, C. Claxton, C. L. Allen, B. Key and J. A. St John, Glia, 2007, 55, 1708-

1719.

39. Z. H. Huang, Y. Wang, X. B. Yuan and C. He, Exp Cell Res, 2011, 317, 2823-2834.

Page 77: Design and Development of Cell Stretching Platforms for ...

76

Chapter 4: Modelling of an uniaxial single-sided magnetically actuated cell-

stretching device

It is evident from the literature review that the actuation method is a key factor in the design

and development of the robust cell stretching platform. Thus, in-depth understanding of the

actuation strategy was an important next step towards the standardisation of the developed

electromagnetic cell-stretching platform. Numerical system modelling and simulation with

finite element analysis (FEA) provides a design tool to obtain optimised parametric values for

the operation of the developed cell-stretching platforms. Furthermore, such tools play an

important role in exploring the possible alternative strategies to further optimise existing

electromagnetic cell stretching approaches. Hence, this Chapter intended to develop a finite

element analysis (FEA) model of the uniaxial electromagnetic cell stretching platform for the

parametric optimisation and standardisation. This Chapter reports the systematic modelling

approach and optimisation of the developed electromagnetically actuated cell-stretching

device. The Chapter describes the numerical simulation of the actuation system consisting of

a permanent magnet and an electromagnet. The capability of the standardised numerical

model was demonstrated by optimising the distance between the electromagnet and

permanent magnet of the developed cell stretching platform. The obtained results suggested

that this actuation system is capable of precisely predicting the behavior of the developed cell-

stretching platform.

Page 78: Design and Development of Cell Stretching Platforms for ...

77

Statement of Contribution to Co-authored Published Paper

Page 79: Design and Development of Cell Stretching Platforms for ...

78

4.1 Introduction

Cells Mechanotransduction is the process of converting mechanical stimuli into biological

electrochemical signals, which guides cell behavior, growth and morphology of cells1-4

.

Mechanotransduction has been reported as an important factor for maintaining cellular

homeostasis5-7

. Therefore, dysfunction or irregularities in cellular mechanotransduction may

well result in diseases such as heart failure, asthma and cancer 8, 9

. For this reason,

mechanotransduction has become an increasing interest for researchers in fields such as

bioengineering and regenerative medicine. However, controlling the complexities associated

with a cell’s physical in-vivo microenvironment is a hubris task, which has led to the

development of in-vitro devices. In-vitro cell stretching devices are designed to mimic the

cells physical microenvironment and provide a greater insight into the complex

mechanotransduction mechanism.

Micropipettes or tweezers are two common in-vitro stretching methods used for introducing

mechanical force into cells10-14

.

A number of commercial cell-stretching platforms are

currently available. For example, Flexcell (Flexcell International Corporation) is considered

the most elegant cell-stretching platform in the market. This system incorporates pneumatic

actuators to induce a homogenous strain to a membrane, where cells are cultured15-17

. Strex

Systems (STREX Inc.) and ElectroForce (Bose Corporation) are other cell-stretching systems,

which also have been widely used18-21

. Apart from commercial cell-stretching systems, several

custom-made cell-stretching devices have been reported in the literature over the last

decade22-26

.

Common methods for stretching cells grown on an elastic membrane include:

electromagnetic, piezoelectric, optical and pneumatic actuators22, 27-31

. Ursekar, et al. (2014)

developed a cell-stretching device with indenter design and utilized a stepper motor to induce

homogenous strain onto a thin membrane. Nava, et al. (2015) introduced optical actuation for

cell-stretching purpose. Huang and Nguyen (2013) utilized pneumatically actuated

multilayered microfluidic device for uniaxial cell stretching. Fior, et al. (2011) designed an

microelectromechanical systems (MEMS) device with externally controlled piezoelectric

Page 80: Design and Development of Cell Stretching Platforms for ...

79

actuation system. Apart from these common methods, di-electrophoresis, electro-thermal and

electrostatic actuations have also been reported for cell-stretching applications32-34

.

All existing cell-stretching methods have some advantages, but very few devices fulfill the

general requirements for robust cell stretching research, such as: precise strain patterning,

compatibility with a variety of microscopes and/or imaging systems (for analysis) and high

experimental throughput. System modelling and simulation are also imperative, to obtain

optimized parametric values in the design and the operation of a cell-stretching devices.

System optimisation with finite element analysis (FEA) will help to achieve the above

requirements and the optimal operation of a cell-stretching device.

Our previous report on the single-sided uniaxial magnetically actuated cell-stretching system

elucidates the design, simulation and characterisation of the developed stretching device

platform with preliminary experimental observations of olfactory ensheathing cells in vitro 26

.

The previous model incorporates a simple structural model. The present chapter focuses on

Fig. 4.1 Operation concept of the cell-stretching device (top view and side view): (a)

Relaxed state; (b) Actuated state.

Constrained wall

Strain

Electromagnet

Permanent magnet

Top view

Deformed front wall

Constrained wall

Membrane with cell culture

Deformed membrane

with cell culture

Actuation axisA A

A-A

A A

A-A

Side view

Front wall

Rela

xe

d s

tate

Actu

ate

d s

tate

(a)

(b)

Page 81: Design and Development of Cell Stretching Platforms for ...

80

the coupled simulation of an electromagnetic actuator using an electromagnet and a

permanent magnet for the uniaxial cell-stretching device. Experimental data subsequently

verified the simulation results, which was demonstrated through optimizing the distance

between the permanent magnet and the electromagnet

4.2 Device Concept

The present numerical study uses the previously reported cell stretching device design as an

example (Fig. 4.1). The cell-stretching system includes a PDMS device and a permanent

magnet (PM) with a diameter of 15 mm and thickness of 2 mm embedded into the front wall

of the device. The permanent magnet is actuated by an electromagnet (EM) controlled by the

programmable direct-current (DC) power supply. A mounting platform provides a support for

the PDMS device and maintains the alignment of the PM and EM along the actuation axis.

The 200-µm thick deformable PDMS membrane was bonded to the bottom of the PDMS

device using oxygen plasma. Figure 4.1 illustrates the operation concept of the actuation

system. Once the EM is activated, the magnetic force acting on the PM deforms the front wall

of the PDMS device, consequently causing the membrane to be stretched uniaxially along the

actuation axis.

Surrounding region

Finely meshed surrounding

medium (Air)

Coilwinding

Core

Actuationaxis

Point of measurement (POM)

(a)

(b)

2D work plane revolution axis

5 mm

Fig. 4.2 The geometry of the electromagnet: (a) 2D Work plane; (b) 3D view.

Page 82: Design and Development of Cell Stretching Platforms for ...

81

4.3 Modelling approach

4.3.1 Modelling of the electromagnet

The finite element analysis (FEA) model was implemented in COMSOL Multiphysics 5.2

(COMSOL, Inc., MA, USA) utilising AC/DC and structural mechanics modules for the

optimisation of our existing cell stretching device 26

. For computational simplicity, the

transient nature of the problem is modelled in two quasi-static steps. The first step includes

modelling of the actuation system and the calculation of the magnetic force. The second step

includes coupling of the magnetic force from the first step with a structural model to

determine the strain on the membrane.

The calculation of the magnetic force was further divided into two subtasks. The first task

includes modelling the EM and its validation with experimental data to confirm the stability

of the model. The second task introduces the PM into the already optimised EM model

environment (task 1) to obtain a superimposed actuation condition of the complete cell-

stretching device. Once more, experimental data was used to validate the results of the

second task.

The linear relationship between the actuating current I and the generated magnetic flux

density was used to model EM in COMSOL Multiphysics 5.2. Considering the EM as an

induction coil in an ideal state and neglecting environmental disturbance, the flux density is

determined as:

𝐵 = µ 𝑁 𝐼 , (1)

where B is the magnetic flux density (in T), I is the current (in A); µ and N are the magnetic

constant and the number of turns, respectively.

A static study in AC/DC module with magnetic fields was considered for modelling the EM

in COMSOL 35

. The geometry mainly consists of the EM core, the coil and the surrounding

medium for magnetic field propagation. The dimension of the EM model was taken from the

actual EM [MK magnets, Seoul, Korea] of our existing cell-stretching device. For better

visualisation of the simulated magnetic field, a 2D work plane modelling approach was first

considered. Subsequently, the 360° revolution option in COMSOL was utilised to build the

Page 83: Design and Development of Cell Stretching Platforms for ...

82

3D EM model with the core and the coil winding. The EM measured 40 mm in length with a

core diameter of 10 mm and an outer coil diameter of 25 mm (Fig. 4.2). The surrounding

medium was confined in a cylinder that had a diameter of 15 mm and length of 52 mm.

Furthermore, iron, copper and room air were selected to constitute the cylinder core, the coil

and the medium domain, respectively. The cylindrical coordinate system was used. Multi turn

coil study was selected for the EM modelling assuming the following equation;

𝐽e = (𝑁 𝐼coil

𝐴) 𝑒coil (4)

where A is total cross section area of the coil domain (in m2), 𝐽𝑒 is current density in the

direction of wires (in A/m2) and ecoil is local coil direction. The experimentally obtained coil

current 𝐼coil (in A), the number of turns N and the conductivity of copper of 6x107 S/m were

considered as input parameters for the model.

The centre of the electromagnet face was considered as the point of measurement to evaluate

the magnetic flux density. Following the same experimental setup reported in our previous

Fig. 4.3 Magnetic flux density obtained with coil current of 𝐼𝑐𝑜𝑖𝑙 = 0.183 A: a) With

coarse surrounding medium mesh (B= 28.6 mT); b) With optimised mesh of the

surrounding (B= 88.2 mT).

B= 28.565mT

B= 88.175mT

Finely meshed surroundingmedium (Air)

Electromagnet core

Surrounding region

Magnetic field

Magnetic field

Z

YX

Page 84: Design and Development of Cell Stretching Platforms for ...

83

paper 26

, where we obtained the coil current 𝐼𝑐𝑜𝑖𝑙 and the corresponding magnetic flux density

B of the EM over the voltage range of 0-15 V. Furthermore, utilizing the parametric sweep

option in COMSOL, the magnetic flux density at the center of the EM along actuation axis

was obtained over the above voltage range.

The simulation results showed that the magnetic field generated by the EM is heavily affected

by the meshing of surrounding medium (Fig. 4.3). To achieve the optimal modelling

conditions, an average of ten magnetic flux density readings (i.e. B=86.2±0.6 mT obtained

upon actuation of EM with 𝐼coil= 0.183 A) was considered as a reference. We observed that

the general approach of coarse mesh reduces the effective magnetic field strength (Fig. 4.3a),

yielding a magnetic flux density of only 28.6 mT. Thus, a finer surrounding mesh is important

for accurate propagation of the magnetic field through the surrounding. Moreover,

singularities at the edges were required to minimise the errors and obtain realistic simulation.

However, a finer surrounding mesh was computationally expensive and adds to the

singularities. To mitigate this obstacle, we considered additional smaller medium entities

around the EM with finer mesh. Using this optimisation, we observed that finer mesh along

Fig. 4.4 Magnetic flux density versus applied current at the surface of the electromagnet

(inset: experimental setup and COMSOL Simulation result at 15V and 0.0911A.)

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1

Current (A)

0

5

10

15

20

25

30

35

40

45

50

Ma

gn

etic flu

x d

en

sity (

mT

)

Experimental results

Simulation results

Core

Magnetic filed

(mT)

0

10

20

30

40

50

Electromagnet

Gau

ss m

etr

e

Point of Measurement

(POM)

Current

Actuation axis

Page 85: Design and Development of Cell Stretching Platforms for ...

84

the edges and in the immediate environment of the EM improved the wave propagation. This

approach greatly reduced the processing time of the COMSOL solver as well as the errors

related to wave propagation, yielding a magnetic flux density of 88.2 mT (Fig.4.3b).

Finally, according to the refined geometry domains with the model, a flux density was

obtained at the point of measure (POM) along the actuation axis. The simulation results

agreed well with the experimental data with an error variance of 5%, which is also the error of

the Gauss meter (Fig. 4.4.) The agreement of the data verifies the numerical model of the EM.

4.3.2 Modelling of the magnetic force

To determine the magnetic force, the PM was placed into the previously modeled EM

environment (section 3.1). The magnetic fields of the two magnets were superimposed,

and the force exerted on the surface of the PM was determined.

4.3.2.1 Modelling of the permanent magnet

The same axial symmetric approach as described in Section 3.1 was utilized to build

the PM geometry with a thickness of 2 mm and a diameter of 7.5 mm. Axially aligned

Region of interest (ROI) for measurement

(a)

(b)

Permanent Magnet

Finely meshed surrounding medium for

wave propagation

7.5

mm

Fig. 4.5 Meshed model of the two magnets for the simulation of the magnetic force acting

on the permanent magnet: a) 2D work plane for coupled geometry; b) Optimised meshing

conditions for simulation.

Page 86: Design and Development of Cell Stretching Platforms for ...

85

PM was introduced into the system at a distance of 6 mm away from the EM face

along the actuation axis to model the position of the PM and the EM in the actual cell-

stretching device. Neodymium (NdFeB) was selected as the material of the permanent

magnet.

The Maxwell-Amperes law was utilized to describe the PM:

∇ × 𝐇 = 𝐉 , 𝐁 = ∇ × 𝐀 , and 𝐉 = σ 𝐄 𝐁 (5)

where H is magnetic field strength (A/m), J is current density (A/m2)

, A is magnetic

vector potential (Wb/m), E is electric filed (V/m) and B is magnetic flux density (T).

Further, to incorporate an appropriate input we calculated the magnetic remanence

(𝐵𝑟) of the PM. The magnetic flux at the centre surface of the permanent magnet

along the actuation was measured with a Gauss meter (Hirst Magnetic Instrument

Ltd.). Ten measurements were taken, and the average magnetic flux density was

B=206.2 mT. This experimental value was used to calculate the magnetic remanence

(𝐵𝑟) of the model according to the relationship:

𝐵 =𝐵𝑟

2(

𝐷+𝑧

√𝑅2+(𝐷+𝑧)2−

𝑧

√𝑅2+(𝑧)2) (6)

where D=2mm and R=7.5 mm are the thickness and the radius of the PM, respectively.

The axial distance from the surface was taken as z=0 mm.

The obtained magnetic remanence (𝐵𝑟=1.6 T) served as an input for the model of the

PM. Considering the spring constant of the PDMS wall in the actual cell-stretching

device, the back surface of the PM was fixed to obtain the force reaction onto the PM

upon the field of the EM.

4.3.2.2 Modelling of the superimposed system with electromagnet and permanent

magnet

For the superimposed condition, the magnetic flux density at the centre and on the

surface of the PM was considered. For the measurement, both magnets were mounted

6 mm apart onto the linear stage of a syringe pump along the actuation axis. The Gauss

Page 87: Design and Development of Cell Stretching Platforms for ...

86

meter was mounted perpendicular to the actuation axis and the magnetic flux density

reading where taken at POM over the operational range of 0-15V.

To obtain optimal modelling conditions, the magnetic flux density was evaluated with

the maximum actuation voltage of 15V or a corresponding coil current of 𝐼coil=

0.0911A. Ten measurements resulted in an average magnetic flux density of B= 209

±0.52 mT which was considered as a benchmark.

To achieve the optimal modelling benchmark, we included an additional finely meshed

medium entity in between EM and PM to avoid propagation error in the coupled

system (Fig. 4.5). Finally, the system was meshed according to the geometry domain

and solved to obtain superimposed magnetic flux density B at the centre of the PM

surface over the range of coil current 𝐼coil corresponding to the voltage range of 0-15

V. Figure 4.6 illustrates the simulation and experimental results of the magnetic field

versus the actuation current.

4.3.2.3 Magnetic force calculations

Fig.4.6 Superimposed magnetic flux density on the surface of the permanent magnet

versus the actuation current (Inset: COMSOL simulation result at 15V with 9,11 mA

yielding 203.36 mT.)

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1Current (A)

190

195

200

205

210

215

Ma

gne

tic f

lux d

ensity (

mT

)

0

2.5

2

1.5

1

x102

Experimental results

Simulation results

Magnetic Field

PermanentMagnet

Electromagnet

(mT)

Page 88: Design and Development of Cell Stretching Platforms for ...

87

Maintaining the optimized modelling conditions for the superimposed case, the

magnetic force was calculated at the surface of the PM along the actuation axis. Force

calculation in the magnetic field study was utilised to obtain the force exerted onto the

PM assuming following equation:

F = [∫ 𝑛𝑇𝑑𝑠−𝜹;𝛺

] (7)

𝜏 = [∫ (r − r0) X (nT)𝑑𝑠−𝛅;Ω

] (8)

τax =rax

|rax|. τ (9)

where F is total force (in N), 𝜏 is torque (in Nm), T is Maxwell stress tensor (in Pa), 𝑛

is outward normal from the domain and r0 is a point on the axis of rotation.

Figure 4.7 compares the simulated magnetic force with the measured magnetic force.

The magnetic force was measured in our previous work by fitting the actual deflection

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1

Current (A)

0

0.2

0.4

0.6

0.8

1

1.2

Mag

netic f

orc

e (

N)

Experimental results

Simulation results

0

1.2

0.8

0.4

0.2

Permanentmagnet

6mmElectromagnet

Magnetic force (N)

15 m

m

Fig. 4.7 Magnetic force versus applied current (Inset: COMSOL Simulation result at 15V

and 0.973A generating 1.2N force on the surface of the PM).

Page 89: Design and Development of Cell Stretching Platforms for ...

88

of the cell-stretching device with a FEA model [26]. The statistical analysis of the data

provides an error variance of 15%, which is acceptable considering the simplicity of

the model and the practical requirement.

4.3.3 Modelling of the magnetically actuated cell-stretching device

As the magnetic force acting on PM was modelled and verified in section 3.2, the next step

was to couple the magnetic force with the structural model of the cell-stretching device. In our

previous work 26

, the acceptable working distance between PM and EM was chosen as 5 mm,

confirming no visible actuation. The force requirement varies depending on the cells under

investigation. The FEA model described here can be utilized to optimize the acceptable

distance for better force stability. We utilised the FEA model and calculated the force exerted

on the PM surface at 15V or 𝑰𝐜𝐨𝐢𝐥= 91.1 mA by keeping the PM fixed and varying the EM

position over the range of 0.3 to 8 mm along the actuation axis. Parametric sweep option in

COMSOL was chosen to predict the system behavior.

1 2 3 4 5 6

1

2

3

4

5

6

7

8

Displacement (mm)

Fo

rce

(N

)

Active attraction Phase

Plateau phase

Electromagnet

Cell stretching device

Con

str

ain

ed

ba

ck w

all

Permeant magnet

Actu

atio

n a

xis

Deformable membrane

8mm

Fig. 4.8 Simulated actuation force on the permanent magnet at different distances between

the two magnet (actuation voltage of 15V with the corresponding actuation current of 91.1

mA).

Page 90: Design and Development of Cell Stretching Platforms for ...

89

Figure 4.8 depicts the simulation results, indicating the two distinct phases: 1) active

attraction phase where the force was observed to decrease exponentially from 8 N at 0.3 mm

to 2.3 N at 3.7 mm, and 2) the stable phase, where an average force of 1.45 N was obtained at

4 mm distance, with a 6% variation over the range of 4 to 5.7mm. Thus, 4.2 to 5.7 mm are

considered as a stable acceptable distance range between the PM and EM, that yielded 1.6 N

and 1.3 N respectively. These results agree well with the empirical distance of 5 mm used in

our previous cell-stretching experiments 26

.

4.4 Conclusion

A device was designed, developed and characterised to apply known strain to OEC culture

and investigate mechanotransduction behavior. Primary experimentations of OEC culture

shows that the length OECs decreases with increase in strain experienced because of the fact

that either they are unable to maintain their focal adhesion or morphology change has

distrusted OECs ability to distinguished direction. The observations obtained from the

experimental results illustrate the effectiveness of the developed platform and provides

guidelines for the future exploration of the OECs mechanotransduction studies.

Page 91: Design and Development of Cell Stretching Platforms for ...

90

4.5 References

1. B. D. Hoffman and J. C. Crocker, Annu. Rev. Biomed. Eng., 2009, 11, 259-288.

2. C. L. Happe and A. J. Engler, Circ. Res., 2016, 118, 296-310.

3. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

4. R. P. Brandes, N. Weissmann and K. Schroder, Antioxid. Redox Signal., 2014, 20,

887-898.

5. D. J. Leong, J. A. Hardin, N. J. Cobelli and H. B. Sun, in Skeletal Biology and

Medicine Ii: Bone and Cartilage Homeostasis and Bone Disease, ed. M. Zaidi,

Blackwell Science Publ, Oxford, 2011, vol. 1240, pp. 32-37.

6. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6, 6333.

7. J. H. Wang and B. P. Thampatty, Biomech Model Mechanobiol, 2006, 5, 1-16.

8. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

9. D. E. Ingber, Ann Med, 2003, 35, 564-577.

10. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

11. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

12. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

13. Y. L. Sun, Z. P. Luo, A. Fertala and K. N. An, J. Biomech., 2004, 37, 1665-1669.

14. O. Chaudhuri, S. H. Parekh, W. A. Lam and D. A. Fletcher, Nat. Methods, 2009, 6,

383-387.

15. H. Kufaishi, M. Alarab, H. Drutz, S. Lye and O. Shynlova, Reprod. Sci., 2016, 23,

978-992.

16. P. S. Nayak, Y. Wang, T. Najrana, L. M. Priolo, M. Rios, S. K. Shaw and J. Sanchez-

Esteban, Respir. Res., 2015, 16, 60.

17. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog Biophys Mol Biol, 2014, 115, 93-102.

18. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

19. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

Page 92: Design and Development of Cell Stretching Platforms for ...

91

20. F. H. Bieler, C. E. Ott, M. S. Thompson, R. Seidel, S. Ahrens, D. R. Epari, U.

Wilkening, K. D. Schaser, S. Mundlos and G. N. Duda, J. Biomech., 2009, 42, 1692-

1696.

21. A. M. Throm Quinlan, L. N. Sierad, A. K. Capulli, L. E. Firstenberg and K. L. Billiar,

PLoS One, 2011, 6, e23272.

22. Y. Huang and N. T. Nguyen, Biomed. Microdevices, 2013, 15, 1043-1054.

23. K. B. Roth, K. B. Neeves, J. Squier and D. W. M. Marr, Biomed. Opt. Express, 2015,

6, 807.

24. R. Fior, S. Maggiolino, M. Lazzarino and O. Sbaizero, Microsys. Technol., 2011, 17,

1581-1587.

25. C. P. Ursekar, S. K. Teo, H. Hirata, I. Harada, K. H. Chiam and Y. Sawada, PLoS

One, 2014, 9, e90665.

26. K. Harshad, M. Jun, S. Park, M. J. Barton, R. K. Vadivelu, J. St John and N.-T.

Nguyen, Biomed. Microdevices, 2016, 18, 1-10.

27. G. Nava, F. Bragheri, T. Yang, P. Minzioni, R. Osellame, I. Cristiani and K. Berg-

Sørensen, Microfluid. Nanofluid., 2015, 19, 837-844.

28. Y. Shao, X. Tan, R. Novitski, M. Muqaddam, P. List, L. Williamson, J. Fu and A. P.

Liu, Rev. Sci. Instrum., 2013, 84, 114304.

29. D. Tremblay, S. Chagnon-Lessard, M. Mirzaei, A. E. Pelling and M. Godin,

Biotechnol. Lett, 2014, 36, 657-665.

30. S. Deguchi, S. Kudo, T. S. Matsui, W. Huang and M. Sato, AIP adv., 2015, 5, 067110.

31. H. Kamble, M. J. Barton, M. Jun, S. Park and N.-T. Nguyen, LChip, 2016, DOI:

10.1039/C6LC00607H.

32. Y. Iwadate and S. Yumura, Biotechniques, 2009, 47, 757-767.

33. I. Guido, C. Xiong, M. S. Jaeger and C. Duschl, Microelectron. Eng., 2012, 97, 379-

382.

34. N. Scuor, P. Gallina, H. V. Panchawagh, R. L. Mahajan, O. Sbaizero and V. Sergo,

Biomed. Microdevices, 2006, 8, 239-246.

35. C. Multiphysics, 2012.

Page 93: Design and Development of Cell Stretching Platforms for ...

92

Chapter 5: An electromagnetically actuated double-sided cell-stretching

device for mechanobiology research

The developed electromagnetic uniaxial single sided cell stretching platform detailed

in the preceding Chapter 3 was suitable to obtain strain requirement for the cell type

under investigation. However, further in-depth investigation of the homogenous strain is

critical for a robust cellular response analysis. Thus next logical step was to optimise the

developed cell-stretching platform using existing numerical models of the uniaxial single-

sided cell stretching platform detailed in Chapter 4. Hence, this Chapter aims to optimise the

developed electromagnetic uniaxial single sided cell stretching platform to incorporate

different stretching modes with homogeneous strain pattern. Building on the gained

knowledge, the effect of axially aligned two permanent magnets were explored in this Chapter

to achieve the homogeneous strain condition. This study resulted in the successful

development of an electromagnetically actuated double-sided cell stretching platform to

introduce homogenous strain onto the membrane using two axially aligned electromagnetic

actuators. This chapter reports the working principle, design, simulation, and characterisation

of a novel electromagnetic cell stretching platform based on the double-sided axial stretching

approach. The device is capable of introducing a cyclic and static strain pattern on a cell

culture. The platform was tested with fibroblasts cells.

Page 94: Design and Development of Cell Stretching Platforms for ...

93

Statement of Contribution to Co-authored Published Paper

Page 95: Design and Development of Cell Stretching Platforms for ...

94

5.1 Introduction

The cells in a functioning multicellular system are continuously exposed to various

mechanical forces. The ability of the cells to recognise these mechanical stimuli and

transform them into chemical responses is known as a mechanotransduction 1-3

. It is well

known that abnormalities in mechanotransduction signalling pathways affect the cell

behaviour and tissue homeostasis, consequently leading to pathogenesis 4, 5

.

Mechanical forces related to the cyclic deformation of soft tissues are essential for the

maintenance of various physiological conditions of organs such as the heart, blood vessels,

and lungs 6. The connective tissues of these organs contain a significant amount of fibroblastic

cell types. The mechanotransduction mechanisms in fibroblasts are crucial to modulate tissue

homeostasis 7. Fibroblasts continually perceive external mechanical stimuli, which

subsequently lead to the production and remodelling of extracellular matrix (ECM)

components. For instance, numerous studies have reported that, under cyclic or static strain,

fibroblast cells synthesise the ECM protein 8. Fibroblasts anchor their actin cytoskeleton with

ECM by linkage proteins called integrins. Integrins act as mechanosensors that sense the

physical forces applied to the cell surface. The ECM-integrin-cytoskeleton integration plays a

vital role in the functional and structural adaptation of cell in response to mechanical cues 9.

These mechanical signals are then transmitted to the cytoskeleton by the formation of an

ECM-actin cytoskeleton linkage. This complex system promotes the assembly of focal

adhesion and thereby induces the reconstruction of the actin cytoskeleton that is necessary for

cell stiffening, gripping, and adherence.

The cytoskeleton reorganisation exerts adaptive responses in the cell under mechanical stimuli

10. This process enables cells to regulate the formation of new lamellaipodia and to adjust

their adhesion to resist physical deformation 11

. Thus, cell displacement has been observed for

different amplitudes, types, and directions of mechanical strain. For instance, cyclic stretching

at high frequency aligns the cells perpendicularly to the stretching direction 12

. In contrast,

static stretching induces cells to align parallel to the stretching direction 13

. Hence, better

understanding of the cellular response of the fibroblast culture, mechanical stimuli will

provide an insight into examining how mechanical factors alter the physiology and behaviour

of the fibroblasts. Thus, various techniques have been developed to introduce mechanical

stimuli to the cellular microenvironment. Considering the complex in vivo microenvironment

Page 96: Design and Development of Cell Stretching Platforms for ...

95

of the cells, the majority of the cell stretching approaches has been developed as in vitro

platforms 14-17

.

Most cell stretching approaches include the use of tweezers or micropipettes to induce the

mechanical stimuli 18-21

. However, commercial cell stretching platforms such as Flexcell

(Flexcell International Corporation, Burlington, NC, USA), Strex Systems for cell Stretching

(STREX Inc., Osaka, Japan), and ElectroForce have been recently available 22-24

. Moreover,

various customised stretching platforms have been reported in the last decade. The majority of

these platforms utilise electromagnetic, pneumatic, piezoelectric, or optical actuators to

deform an elastic membrane with cells cultured on it 14, 25-27

. For instance, Shimizu et al. 28

developed a microfluidic device with serially connected balloons. Authors utilised a positive

pressure to inflate the balloons and directly induce strain onto the cells. Similarly, Kamotani

et al. 29

designed a microfluidic device with a deformable membrane at the bottom and an

array of piezo electrically actuated pins, which were placed below the microwell plate. Each

microwell was independently actuated by the pin, which deforms the micro-well membrane

with cells cultured on it. Furthermore, Huang et al. 30

designed a cell stretching platform with

an indenter and utilised a servomotor to introduce strain onto the deformable membrane. In

another study, Sraj et al. 31

designed a microfluidic channel and utilised a single-mode laser

(830 nm, 200 mW) to trap and deform the cells.

In addition to the above-mentioned actuation approaches, electrothermal, electrostatic, and

dielectrophoretic actuations have recently been adapted to introduce mechanical force onto

the cells 32-34

. All cell stretching approaches reported in the literature have their specific

advantages. However, very few platforms provide the main features of a robust cell stretching

tool such as a standardised strain pattern, a wide range of imaging options, and high-

throughput capability. This chapter presents a novel cell stretching platform with a double-

sided uniaxial magnetically actuated stretching approach to introduce both homogeneous

cyclic and static strain onto the cell culture. The cell stretching platform is homogenous and

provides a wide range of strain values, cyclic and static stretching modes, compatibility with

general clinical tools, and imaging options. Thus, our system is suitable for long-term cell

stretching studies. The present chapter discusses in detail the modelling, fabrication, and

characterisation of the cell stretching platform and provides preliminary observations of the

Page 97: Design and Development of Cell Stretching Platforms for ...

96

cellular response for fibroblasts under cyclic stretching, static stretching and non-stretching

conditions.

5.2 Materials and methods

5.2.1 Device design and working principle

Figure 5.1 illustrates the schematic of the developed cell stretching platform. The platform

consists of the polydimethylsiloxane (PDMS) device, with embedded permanent magnets, a

holding clip for the static strain condition, and a mounting stage with electromagnets for

cyclic strain conditions. The PDMS device has two NdFeB disc magnets (15 mm diameter

and 2 mm thickness) embedded in the front and back wall, which are 4 mm thick and placed 8

mm apart.

The permanent magnets are positioned along the actuation axis such that the north poles of

both magnets are facing each other to induce repulsive force and to deform the front and back

wall of the PDMS device. This active magnetic repulsion was utilised to introduce the static

strain onto the deformable membrane. The custom made 3D printed holding clip was used to

Fig. 5.1 Schematic illustration of the working principle: (a) Steps to obtain static

strain; (b) ON state for cyclic stretching condition; (c) Actual cyclic cell stretching

platform and the PDMS device.

Page 98: Design and Development of Cell Stretching Platforms for ...

97

overcome the active magnetic repulsive force and to maintain the PDMS device in a

predefined position. The 200-µm thick deformable membrane was then bonded at the bottom

of the device after oxygen plasma treatment. Removing the holding clip results in the desired

static strain on cells cultured on the deformable membrane, Figure 5.1a. This simple but

effective strategy allows a wide range of predefined strain on the deformable membrane to be

achieved, simply by controlling the initial and final position of the holding clip during the

bonding process of the deformable membrane and the PDMS body.

For homogenous strain, two axially aligned electromagnets (JL Magnet Co., Ltd, Seoul,

Korea.) controlled by a DC power supply (MK Power, M-K Power Products Corporation.,

Mississauga, ON, Canada) were used simultaneously to externally actuate the magnets

embedded in the PDMS device. The mounting platform was 3D printed to introduce the

necessary constraint onto the PDMS device and to provide axial alignment for the permanent

magnets (PMs) and electromagnets (EMs). Upon actuation, the magnetic forces generated by

the EMs and the PMs deform the front and the back wall of the PDMS device. This force is

further transferred to the deformable membrane, which in turn induces a well-defined strain

onto the cells cultured there, as in Figure 5.1b. Figure 5.1c shows the actual developed cell

stretching platform, which is capable of inducing homogeneous cyclic strain onto cells

cultured on a deformable membrane, for studying the behaviour of the cells.

5.2.2 Modelling and fabrication

A finite element analysis (FEA) model of the PDMS device was formulated in COMSOL

Multiphysics 5.2 (COMSOL, Inc., Burlington, MA, USA) to understand and to optimise the

stretching device. We modified and updated the previously reported FEA model to achieve

the necessary parametric optimisation of the device 35. The grade of the permanent magnets

(PMs) was the key optimisation parameter, which was taken into consideration in order to

manipulate the active magnetic repulsive force to obtain optimised static strain conditions for

the PDMS device without physically damaging the device.

To start with, a reference FEA model was formulated in COMSOL to understand the

magnetic field requirement. The two PMs were defined at the axial distance of 8 mm, and the

diameter and thickness of the PMs were defined as 15 mm and 2 mm, respectively. The

Page 99: Design and Development of Cell Stretching Platforms for ...

98

NdFeB material was assigned for the adapted geometry, and the PMs were modelled using

Maxwell-Amperes law.

𝛁 × 𝐇 = 𝐉 , B = 𝛁 × 𝐀, 𝐉 = 𝛔 𝐄 𝐁 (1)

where B is the magnetic flux density in T, E is an electric field in V/m, A is magnetic vector,

and H is the magnetic field strength in A/m.

Furthermore, the optimisation was carried out by varying one parameter while maintaining the

other parameters constant. Maintaining the geometrical parameters constant, the magnetic

remanence (𝐁𝐫) of the PMs was varied over the range of 0.05 T to 2 T, with an incremental

step of 0.05 T, using a parametric sweep function to model the variation of the NdFeB magnet

grade. The inset in Figure 5.2 shows the results for the FEA model with an input magnetic

remanence (Br) of 1.2 T. In the next step, to measure the corresponding deformation, the FEA

model geometry of the PDMS device was built with a front and back wall thickness of 4 mm,

a side wall thickness of 2 mm, deformable membrane thickness of 0.2 mm, and axially

aligned PMs with 15 mm diameters and 2 mm thickness embedded into the front and back

wall. The total volume of the geometry formulated in COMSOL was 30 mm × 25 mm × 12

mm.

Next, the study of the structural mechanics and the magnetic force was coupled in COMSOL

to estimate the outward force acting on the front and back wall for various values of magnetic

remanence. The appropriate material properties (750 kPa Young’s modulus and 0.49

Poisson’s ratio for 10:1 PDMS-cross linker mixture) were selected to match the real device 36-

38. All four corners of the device were fixed to maintain the necessary boundary conditions.

Figure 5.2 shows the average displacement versus the magnetic remanence of the PMs along

the actuation axis. The parametric optimisation suggested that 1.2 T magnetic remanence

(grade N35) yields a magnetic flux of 103 mT at the surface of the PM and generated an

average displacement of 0.788 mm between the two embedded PMs along the actuation axis.

Thus, for the static condition, the magnetic repulsion facilitated a maximum static strain of

9.85% onto the membrane. Considering the experimental requirement, a NdFeB magnet of

grade N35 was selected for the cell stretching system.

Page 100: Design and Development of Cell Stretching Platforms for ...

99

The next step was the fabrication of the optimised design. The fabrication process involved

the fabrication of a master mould, a mounting platform, and the PDMS device. To replicate

the PDMS device with the optimised geometry, all parts of the master mould were designed

with the necessary fabrication tolerance using SolidWorks 2013 (Dassault Systemes

Solidworks Corp., Waltham, MA, USA). Each part was then carefully cut from the non-

magnetic aluminium material and milled as per the design specifications. Finally the

necessary screwing holes were drilled, and the parts were assembled to obtain the master

mould.

The mounting platform and holding clip were also designed in SolidWorks 2013 to achieve

the necessary constraint and axial alignments. The optimised mounting platform and holding

clip designs were 3D printed using an Eden 260V printer (Stratasys Ltd., Eden Prairie, MN,

USA). In the final step, the PDMS device and the deformable 200 μm membrane were

fabricated.

For the fabrication of the PDMS device, 20 g of degassed mixture of PDMS and cross linker

Fig. 5.2 Simulated axial displacement versus magnetic remanence. (Inset: FEA results with

Br=1.2T, selected NdFeB grade N35 magnet, PDMS device with static strain condition).

device.

Page 101: Design and Development of Cell Stretching Platforms for ...

100

(Sylgard 184 elastomer kit, Dow Corning, Midland, MI, USA) was prepared with a volume

ratio of 10:1. The mixture was poured into the master mould and was again degassed for 15

min to remove any remaining air bubbles. The mould was then carefully closed to cure the

PDMS mixture for 2 h at 80 °C in a vacuum oven. Once cured, the mould was carefully

opened, and the PMs were placed into the created cavities such that the north poles of both

magnets faced each other in order to achieve active magnetic repulsion condition. To fix the

magnets in position, a small amount of PDMS-cross linker mixture (10:1 volume ratio) was

coated onto the PMs. The mould was then carefully closed and cured for another 30 min at 80

°C in a vacuum oven to ensure the proper placement of the PMs. The cured PDMS device was

then inspected and carefully removed from the mould. Finally, the PDMS device was cleaned

with isopropanol and de-ionized (DI) water. In the next step, a degassed PDMS-cross linker

mixture with a volume ratio of 10:1 was spin coated at 400 rpm for two min and cured at 80

°C for two hours to achieve the 0.2-mm thick deformable membrane 39, 40

. The cured

membrane was inspected to confirm its uniform thickness and then cleaned with isopropanol

and DI water. In the last step, the PDMS device was plasma bonded with the deformable

membrane and cured for one hour at 80 °C.

5.2.3 Cell culture

Mouse 3T3 fibroblast cells were cultured in DMEM/F12 (Gibco, Thermo Fisher Scientific,

Waltham, MA, USA) medium with 10% fetal bovine serum (FBS) and 1% penicillin at 37 °C

and 5% CO2 in a standard incubator. To sterilise the device, the device was treated with 80%

ethanol and washed three times with 1× phosphate buffered saline (PBS) followed by

ultraviolet (UV) exposure for 20 min. Before seeding the cells, the device was treated with

fresh media and incubated for one hour under standard conditions (37 °C and 5% CO2) to

further ensure its biocompatibility. For seeding the device, 80% confluence was reached in

T75 flask and the cells were harvested at the optimised density of 75 × 103 Cells/250 μL. In

order to achieve the optimal adherence and growth of the cells on the membrane, the device

was incubated at 37 °C and 5% CO2 for 24 h before the mechanical strain was applied.

5.2.4. Application of strain on Fibroblasts

Two stretching modes were tested, namely, cyclic and static axial stretching, and compared to

the non-stretching (control) condition for this study. Cyclic stretching was applied to the cells

Page 102: Design and Development of Cell Stretching Platforms for ...

101

with 1.4% strain at 0.01 Hz and 50% duty cycle. For a comparative study, the same 1.4%

strain was applied to the fibroblast cell culture under the static condition. The strain was

introduced onto the cells over five different time instances (0.5 h, 1 h, 2 h, 3 h, and 4 h) for a

maximum of 4 h. The cells’ characteristics such as area, aspect ratio, and orientation were

observed after stretching and compared to the control. For the analysis, the central region of

the membrane was imaged with an inverted microscope (Nikon Eclipse Ts2, Nikon

Corporation, Tokyo, Japan). Biological triplicates were performed.

5.2.5 Cell fixing, immunofluorescence staining and imaging

The stretched cells were fixed with 4% paraformaldehyde (PFA) for 15 min followed by the

three 10-min PBS washes. The deformable membranes with fixed cells were stored in the

PBS solution at 4 °C. For actin and nuclei observation, the cells were incubated with

ActinGreenTM 488 and NucBlueTM ReadyProbeTM Reagent (Thermo Fisher Scientific) for

30 min, followed by three post-staining washes with PBS. Images of the cell nuclei and actin

fibres were finally obtained with a fluorescent microscope (Olympus BX50, Olympus

Corporation, Tokyo, Japan) using 10× and 20× magnification.

5.2.6 Image analysis

For the image analysis, three separate locations within the central region, i.e., the region of

interest (ROI) with homogenous strain, were captured. Each image was enhanced using post

processing with Image J (National Institutes of Health, Bethesda, MD, USA), which mainly

included fast fourier transform (FFT) bandpass filtering, sharpening, enhancing the image

contrast, and thresholding. For quantification, the captured cells were analysed to estimate the

averaged area, aspect ratio, and orientation of the cells in ROI at each time instance.

5.3 Results and discussion

5.3.1 Force calculation

For estimating the experimental magnetic force, we utilised a similar approach as that

reported in our previous work 41. The optimised FEA model of the PDMS device was utilised

to obtain the spring constant for the experimental conditions. Considering the stretching

condition, the magnetic actuation was modelled in COMSOL by introducing an outward force

Page 103: Design and Development of Cell Stretching Platforms for ...

102

onto the PMs to deform the front and back walls of the PDMS device. Further, simulated

input force within a FEA model was varied from 0.5 Newton to 1 Newton with an incremental

of 0.5 mNewton for each step to observe the corresponding displacement of the PMs.

Considering the material properties, we assumed the Hooks law to obtain the spring constant

k for this study:

𝑭 = −𝒌 ∙ 𝒙, (2)

where, k is the spring constant in N/mm, F is the force in N, and x in the displacement in mm.

The next step was to experimentally obtain the displacement of the PMs over the applied

voltage. Both EMs were simultaneously actuated by supplying voltage ranging from 1 V to 30

V. The corresponding displacement of the marked points on the PDMS device wall along the

actuation axis was recorded for each step using a digital camera (EO Edmund Optics, Edmund

Optics, Barrington, NJ, USA). Furthermore, our particle tracking algorithm based on digital

image correlation and the Matlab image processing toolbox was utilised to detect and measure

Voltage (V)

Forc

e (

N)

0 5 10 15 20 25 300

0.5

1

1.5

2

stageMotorised XY

Camera

XY

Electromagnet

Power Supply

Permanent Magnet

Co

nstr

ain

ed

Co

rners

Deformable wall

Deformable wallMagnetic

force

Constrained corners

Permanentmagnet

Actuation axis

Deformable wall

Magnetic force

Deformable Membrane

FEA model

Spring constant K= 2.41(N/mm)

Experimental resultsSimulation results

8mm

Fig. 5.3 Magnetic force over the voltage range of 1- 30V (Inset: Experimental setup

and FEA model for PDMS device).

Page 104: Design and Development of Cell Stretching Platforms for ...

103

the displacement of the randomly marked points 41. Finally, the obtained average

displacement was used to calculate the force using a spring constant of 2.41 N/mm,

determined by the FEA simulation.

In the next step, we modified and updated our previously reported FEA model to calculate the

magnetic force between the PM and the EM 35 and to validate the experimental data. We

considered the symmetric nature of the system and obtained the magnetic force at the PM

surface along the actuation axis over the voltage range of 1 V to 30 V.35 The simulation

results were verified with the experimental data in Figure 5.3. As expected, a linear force-

voltage relationship can be clearly observed from Figure 5.3. The simulation agrees well with

the experimental data. The results provide an acceptable error variance of 9.42% over the

range of 9 V to 30 V between the experimental and simulation data.

5.3.2 Strain calculation

The characterisation of the strain applied to the deformable membrane was observed using

both experiments and simulation. For measuring the strain experimentally, the membrane

deformation was recorded with a digital camera (EO Edmund) over the voltage range of 1 V

to 30 V. The particle detection and displacement measurement algorithm based on digital

image correlation and the Matlab image processing toolbox was further utilised to calculate

the offset displacement of the marked points. For reliable experimental data, the membrane of

each recorded image was divided into 2 × 5 regions. A minimum of three marked samples

from the central region (M1,2, M1,3, M2,2, M2,3) was observed. Finally, to warrant the

repeatability of the results, three experimentally obtained results were averaged to represent

the displacement of the region. The inset of Figure 5.4 depicts the experimental setup and an

example of the particle detection and tracking algorithm result.

For cross validating the experimental data, we utilised a reference FEA model. The magnetic

force obtained from the force calculation (Section 3.1) over the voltage range of 1V to 30 V

was used as the input for the FEA model. The central region of the membrane was considered

the region of interest (ROI).

An average strain across the membrane was obtained for the operating voltage range, i.e., 1 V

to 30 V. Figure 5.4 compares the average strain over the ROI from both the simulation and the

experiments. The experimental and simulation results agree well. An average error variance

Page 105: Design and Development of Cell Stretching Platforms for ...

104

of 7.89% was observed over the voltage range of 9 V to 30 V. Based on the strain

characterisation, we selected an input voltage of 27 V for both actuators, which provided an

average homogeneous cyclic strain of 1.38 ± 0.021% over the central region of the membrane.

For an understanding of the membrane deformation and strain pattern with the selected input

voltage of 27 V, we utilised the same experimental platform and obtained the image sequence

for the membrane deformation. The images were analysed using the existing particle detection

and tracking algorithm to obtain the strain pattern over the 2 × 5 region matrix. A minimum of

three marked points from each subregion was evaluated to obtain reliable results. Finally, the

average value was utilised to represent the strain over each predefined subregion.

Furthermore, three experiments were conducted for each set of data. Figure 5.5 shows the

obtained strain deformation pattern from (a) the experiment and (b) the simulation. The

expected, homogenous strain pattern over the central region of the membrane is evident from

the results. The experiment and the simulation agree well and provide an average strain of

1.38 ± 0.021% and 1.49%, respectively.

0 5 10 15 20 25 30

Voltage (V)

0

0.5%

1%

1.5%

2%

Str

ain

(1

)

Strain

ON StateOFF State

Camera

ElectromagnetPower supply

Permanent Magnet

Constrained Corners

Deformable wall

stageMotorised XY

X Y Imaging Access

OFF State

Particle tracingpixel density

Particledetection

ON State

Region Of Interest (ROI)

selection

Experimental resultsSimulation results

8mm

Fig. 5.4 Strain on the deformable membrane over the voltage range of 1-30V. (Inset:

Experimental arrangement, the membrane in ON and OFF state, an example for the

particle detection and tracking.)

Page 106: Design and Development of Cell Stretching Platforms for ...

105

The results provide a better understanding of the membrane deformation and confirm the

homogenous strain pattern for the subsequent cell stretching experiments. Figure 5.5 clearly

shows that we could expect a homogenous strain pattern in the central region of the

membrane. Thus it can be assumed that cells located in this region will experience an equal

amount of strain.

5.3.3 Cell area and aspect ratio

Figure 5.6 shows the obtained averaged cell area and aspect ratio for each predefined time

point. The results show the gradual decrease in the cell area under cyclic stretching as a

result of cell displacement and aggregation. The cell aspect ratio over the stretching duration

for a maximum of 4 h also decreased, which was in line with the previously reported

observations 42-45

. The morphological observation suggests that both cyclic and static

stretching led to significant changes in fibroblast adherence. However, we focused more on

cyclic stretching, as native tissues within the body are more exposed to cyclic strain rather

than static strain. Under cyclic strain, mechanotransduction and intercellular physiology

Fig. 5.5 Strain pattern on the membrane with a selected input of 27 V: (a)

Experimental results; (b) Simulation results.

Page 107: Design and Development of Cell Stretching Platforms for ...

106

involving cytoskeletal elements and adhesion molecules drive the morphogenetic process of

cells 46

. The cells respond adaptively against external stress transmission and reorganize their

cytoskeleton integrity by reconstructing the actin stress fibres 47, 48

. We also observed that

cyclic stretching of the fibroblast cell culture triggered the formation stress fibres, and over

time cells enhanced their cell-cell connection and formed cell clusters.

The reorganization of actin stress fibres seems to promote a cell adhesion dynamic and

changes in cell morphology 49

. Furthermore, it was interesting to note that, after two hours of

cyclic stretching, significant cell cluster formation was observed. The stronger cell

connections and the formation of cell clusters explain how individual cells sense and transmit

physical forces to and from neighbouring cells, that is, by the binding of adhesion molecules

to expand cell-cell cohesion 50

.

Furthermore, it is interesting to address the question as to why stretching induces cell

rearrangement and the adhesion of the cells into clusters. Generally, cells on a substrate

Fig. 5.6 Analysis results of cell area and aspect ratio for cyclic, static and no stretch

conditions over the predefined time points (Inset: fluorescent images of the fibroblast

cells with 20x objective for corresponding time points with 50-µm scale bar).

Page 108: Design and Development of Cell Stretching Platforms for ...

107

surface are dynamic in nature and are constantly reorganizing their actin filament network to

retract and extend protrusions (the formation of lamellipodia and filopodia). At the cellular

level, changes in the surface topology lead to the underlying processes involved in the

maintenance of mechanical homeostasis 51

. The stress fibres are projecting tension-bearing

bundles of actin filaments, which act as non-muscle sarcomeres 52

. Cyclic stretching seems to

trigger actin fibres to realign and to establish stronger adhesive cell-cell connections to resist

deformation 53

. Thus we hypothesise that actin organization under stress is critical for

promoting cell adhesion dynamics to form cell aggregates and withstand the applied strain.

In contrast, the cell area and aspect ratio with static stretching did not change significantly.

The cells did not spread and promoted the formation of actin stress fibres under static strain

as compared with cyclic strain, as in Figure 5.6. Additionally, no significant cell clusters

were observed in the static strain mode. This discrepancy further suggests that prolonged

static stretching imposes different effects on the regulation of ECM and adhesion proteins.

Previous investigations into the effect of static strains alone on ECM synthesis showed the

degradation of ECM and adhesion molecules 54

. In a recent study, Cui et al. 55

showed that

cytoskeleton organization differs for cyclic and static and indicated that cyclic stretching

promotes actin fibre formation and cell spreading.

Moreover, as expected for non-stretching (control) conditions, the actin stress fibres were

randomly distributed and no significant changes in the cell area and aspect ratio were

observed.

5.3.4 Cell orientation

The orientation range of 0° to 180° was considered for the cell orientation analysis, which

was further divided into six equal angular regions in 30° increments. For the quantitative

analysis of the cell orientation, the 0° to 180° range was set along the stretching direction in

an anti-clockwise direction, and a total of 900 samples from the three images of the ROI

were analysed using Image J for each time instance. Figure 5.7 shows the distribution of the

cell orientation for each time instance. It was interesting to observe distinct cell orientation

trends for the cyclic and static stretching modes. Under the cyclic stretching mode, the cell

orientation showed two distinct peaks over the 30° to 60° and the 120° to 150° ranges.

Prominent cell orientation was observed after 1 h of cyclic stretching, whereas,, under the

static stretching conditions, two distinct peaks were observed over the 0° to 30° and the 150°

Page 109: Design and Development of Cell Stretching Platforms for ...

108

to 180° ranges. Furthermore, it was interesting to observe that the prominence of the cell

orientation under the static stretching conditions increased over the time. From this

observation, we can conclude that most of the cells are oriented approximately either at 45°

or 135° under cyclic stretching and at 15° or 165° under static stretching conditions (Figure

5.7 inset depicted by arrows).

Fig. 5.7 Cell orientation analysis results for cyclic stretching (1.4%.0.01Hz,50% Duty

Cycle) , static stretching (1.4%) and no stretching condition at 0 hr , 0.5 hrs ,1 hrs, 2

hrs , 3 hrs and 4 hrs . (Inset: fluorescent images of the fibroblast cells with 10x

objective for 0 hr (a-c), 0.5 hrs (d-f), 1 hrs (g-i), 2 hrs (j-l), 3 hrs (m-n) and 4 hrs (p-r)

with 100-µm scale bar).

Page 110: Design and Development of Cell Stretching Platforms for ...

109

Recently, Ugolini et al. 56

reported the ability of human primary cardiac fibroblasts to

differentiate between cyclic mechanical stimuli and controlled oxygen tension. Moreover, in

another study, Ugolini et al. 57

subjected human primary cardiac fibroblasts (CFs) to 2% and

8% cyclic strain for 24 and 72 h and interestingly observed the different cellular responses of

the CFs based on the duration and amplitude of strain. In the present study, we observed a

similar trend. Under a constant strain of 1.4% and duration of 4 h with the same culture

conditions, we observed different cellular responses for the static and cyclic modes.

Furthermore, we also observed that, under static strain, fibroblasts aligned parallel to the

stretch direction, while fibroblast cells reoriented approximately ±45° to the stretching

direction under 1.4% cyclic strain for 4 h, which was in line with the above observations. We

hypothesise that, with further parametric optimisation of the amplitude and duration of the

applied strain, the widely accepted perpendicular alignment could be achieved for the cyclic

stretching mode. Overall, our observation further strengthens the argument that cells are

capable of differentiating, not only between different types, but also between different

magnitudes of stimuli.

Furthermore, as expected for the non-stretching conditions, a random distribution of the cell

orientation was observed. Two distinct and almost symmetric peaks were observed for both

the cyclic and static stretching modes, which provide important evidence on the ability of the

fibroblast cells to recognize and respond not only to the applied strain but also to the strain

direction 58-60

. Moreover, the arrangement of actin stress fibres is clearly involved in the

realignment of the cells under strain. Our observation agrees well with previous studies,

which reported that cells reorient and align themselves due to external strain 58, 61-65

.

The adhesion dynamics of cells are guided by cytoskeleton rearrangement and were believed

to be responsible for cell alignment 53, 59

. The cell–ECM connections are established by

association between actin stress fibres and the focal adhesion, which endows cells to form

stronger ECM connections and cell-cell adhesion 66, 67

. The focal adhesion is responsible for

reorienting cells in a direction in which cells can maintain their stability 59

. In line with this

hypothesis, we also observed cytoskeleton organisation and a distinct cell orientation under

both cyclic and static stretching conditions. In contrast, the non-stretching (control)

conditions led to a random distribution of the cell orientation.

Page 111: Design and Development of Cell Stretching Platforms for ...

110

5.4 Conclusions

In summary, we developed a simple yet effective cell stretching platform capable of

introducing homogeneous cyclic and static strain onto the cell culture. The

characterisation of the developed platform provides a clear understanding of the device

function. Furthermore, the effectiveness of the developed platform was tested with

stretching assays of fibroblasts. Our preliminary analysis suggested that the orientation

of the cells is highly influenced by external mechanical cues. The cell aggregation

observed in the cyclic stretching mode suggests that the cells reorganise their

cytoskeleton to avoid external strain and to maintain intact their extracellular matrix

arrangements. The developed cell stretching platform may serve as a tool to investigate

the cell behaviour under a wide range of strain with cyclic or static stretching modes.

Moreover, considering these preliminary results, the platform may facilitate the active

manipulation of fibroblasts to achieve the desired cell arrangements. This alternative

cell culture method will have a broad range of applications in the field of tissue

remodelling and regenerative medicine. Furthermore, it is also important to note that

cells can be harvested after stretching using general clinical tools to perform standard

biology analysis, which could be critical for clinical diagnosis and subsequent

therapeutic screening.

Page 112: Design and Development of Cell Stretching Platforms for ...

111

5.5 References

1. C. S. Chen, J. Cell Sci., 2008, 121, 3285.

2. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

3. A. W. Orr, B. P. Helmke, B. R. Blackman and M. A. Schwartz, Dev. Cell, 2006, 10,

11-20.

4. D. E. Ingber, Annals of medicine, 2003, 35, 564-577.

5. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

6. T. Mammoto and D. E. Ingber, Development, 2010, 137, 1407-1420.

7. F. Klingberg, B. Hinz and E. S. White, The Journal of pathology, 2013, 229, 298-309.

8. M. Chiquet, A. S. Renedo, F. Huber and M. Flück, Matrix biology, 2003, 22, 73-80.

9. M. A. Schwartz, Cold Spring Harb Perspect Biol, 2010, 2, a005066.

10. D. A. Fletcher and R. D. Mullins, Nature, 2010, 463, 485-492.

11. J. L. Hoon, M. H. Tan and C. G. Koh, Cells, 2016, 5.

12. J. H.-C. Wang, G. Yang, Z. Li and W. Shen, Journal of Biomechanics, 2004, 37, 573-

576.

13. R. De, A. Zemel and S. A. Safran, Nature Physics, 2007, 3, 655-659.

14. S. Deguchi, S. Kudo, T. S. Matsui, W. Huang and M. Sato, AIP Advances, 2015, 5,

067110.

15. D. Tremblay, S. Chagnon-Lessard, M. Mirzaei, A. E. Pelling and M. Godin,

Biotechnol. Lett, 2014, 36, 657-665.

16. H. Kamble, M. J. Barton, M. Jun, S. Park and N.-T. Nguyen, LChip, 2016, 16, 3193-

3203.

17. E. W. Esch, A. Bahinski and D. Huh, Nat. Rev. Drug Discov., 2015, 14, 248-260.

18. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

19. O. Chaudhuri, S. H. Parekh, W. A. Lam and D. A. Fletcher, Nat. Methods, 2009, 6,

383-387.

20. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

21. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

Page 113: Design and Development of Cell Stretching Platforms for ...

112

22. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog. Biophys. Mol. Biol., 2014, 115, 93-102.

23. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

24. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

25. G. Nava, F. Bragheri, T. Yang, P. Minzioni, R. Osellame, I. Cristiani and K. Berg-

Sørensen, Microfluid. Nanofluid., 2015, 19, 837-844.

26. Y. Huang and N. T. Nguyen, Biomed. Microdevices, 2013, 15, 1043-1054.

27. Y. Shao, X. Tan, R. Novitski, M. Muqaddam, P. List, L. Williamson, J. Fu and A. P.

Liu, Rev. Sci. Instrum., 2013, 84, 114304.

28. K. Shimizu, A. Shunori, K. Morimoto, M. Hashida and S. Konishi, Sensors Actuators

B: Chem., 2011, 156, 486-493.

29. Y. Kamotani, T. Bersano-Begey, N. Kato, Y. C. Tung, D. Huh, J. W. Song and S.

Takayama, Biomaterials, 2008, 29, 2646-2655.

30. L. Huang, P. S. Mathieu and B. P. Helmke, Ann. Biomed. Eng., 2010, 38, 1728-1740.

31. I. Sraj, C. D. Eggleton, R. Jimenez, E. Hoover, J. Squier, J. Chichester and D. W. M.

Marr, JBO, 2010, 15, 047010.

32. N. Scuor, P. Gallina, H. V. Panchawagh, R. L. Mahajan, O. Sbaizero and V. Sergo,

Biomed. Microdevices, 2006, 8, 239-246.

33. Y. Iwadate and S. Yumura, Biotechniques, 2009, 47, 757-767.

34. I. Guido, C. Xiong, M. S. Jaeger and C. Duschl, Microelectron. Eng., 2012, 97, 379-

382.

35. H. Kamble, M. J. Barton and N.-T. Nguyen, Sensors and Actuators A: Physical, 2016,

252, 174-179.

36. D. Armani, C. Liu and N. Aluru, 1999.

37. S. Choi and J. K. Park, Lab Chip, 2009, 9, 1962-1965.

38. C. Roh, J. Lee and C. Kang, Molecules, 2016, 21.

39. H.-L. Yin, Y.-C. Huang, W. Fang and J. Hsieh, Sensors and Actuators A: Physical,

2007, 139, 194-202.

40. J. H. Koschwanez, R. H. Carlson and D. R. Meldrum, PLoS One, 2009, 4, e4572.

41. K. Harshad, M. Jun, S. Park, M. J. Barton, R. K. Vadivelu, J. St John and N. T.

Nguyen, Biomed Microdevices, 2016, 18, 45.

42. J. M. Mann, R. H. Lam, S. Weng, Y. Sun and J. Fu, Lab Chip, 2012, 12, 731-740.

Page 114: Design and Development of Cell Stretching Platforms for ...

113

43. X. Mao, R. Said, H. Louis, J. P. Max, M. Bourhim, P. Challande, D. Wahl, Z. Li, V.

Regnault and P. Lacolley, Cardiovasc. Res., 2012, 96, 513-523.

44. K. Sethi, E. J. Cram and R. Zaidel-Bar, Semin. Cell Dev. Biol., 2017.

45. B. Geiger and K. M. Yamada, Cold Spring Harb. Perspect. Biol., 2011, 3, a005033.

46. F. Boccafoschi, M. Bosetti, P. M. Sandra, M. Leigheb and M. Cannas, Cell Adh Migr,

2010, 4, 19-25.

47. E. Kassianidou and S. Kumar, Biochim Biophys Acta, 2015, 1853, 3065-3074.

48. R. Kaunas, 2010.

49. H. Ngu, Y. Feng, L. Lu, S. J. Oswald, G. D. Longmore and F. C. Yin, Ann Biomed

Eng, 2010, 38, 208-222.

50. E. Bazellieres, V. Conte, A. Elosegui-Artola, X. Serra-Picamal, M. Bintanel-Morcillo,

P. Roca-Cusachs, J. J. Munoz, M. Sales-Pardo, R. Guimera and X. Trepat, Nat Cell

Biol, 2015, 17, 409-420.

51. J. T. Parsons, A. R. Horwitz and M. A. Schwartz, Nat Rev Mol Cell Biol, 2010, 11,

633-643.

52. A. B. Verkhovsky and G. G. Borisy, J Cell Biol, 1993, 123, 637-652.

53. A. Tondon and R. Kaunas, PLoS One, 2014, 9, e89592.

54. V. Gupta and K. J. Grande-Allen, Cardiovasc Res, 2006, 72, 375-383.

55. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6, 6333.

56. G. S. Ugolini, A. Pavesi, M. Rasponi, G. B. Fiore, R. Kamm and M. Soncini, eLife,

2017, 6.

57. G. S. Ugolini, M. Rasponi, A. Pavesi, R. Santoro, R. Kamm, G. B. Fiore, M. Pesce

and M. Soncini, Biotechnol. Bioeng., 2016, 113, 859-869.

58. A. Pavesi, G. Adriani, M. Rasponi, I. K. Zervantonakis, G. B. Fiore and R. D. Kamm,

Sci. Rep., 2015, 5, 11800.

59. L. Wang, Y. Li, B. Chen, S. Liu, M. Li, L. Zheng, P. Wang, T. J. Lu and F. Xu, ACS

applied materials & interfaces, 2015, 7, 15088-15097.

60. S. Chagnon-Lessard, H. Jean-Ruel, M. Godin and A. E. Pelling, Integr Biol (Camb),

2017, DOI: 10.1039/c7ib00019g.

61. J. H. Wang, G. Yang and Z. Li, Ann. Biomed. Eng., 2005, 33, 337-342.

62. Q. Wang, H. Huang, K. Wei and Y. Zhao, Biotechnol. Bioeng., 2016, 113, 2191-2201.

Page 115: Design and Development of Cell Stretching Platforms for ...

114

63. N. K. Weidenhamer and R. T. Tranquillo, Tissue engineering. Part C, Methods, 2013,

19, 386-395.

64. H. Abiko, S. Fujiwara, K. Ohashi, R. Hiatari, T. Mashiko, N. Sakamoto, M. Sato and

K. Mizuno, J Cell Sci, 2015, 128, 1683-1695.

65. A. Papadopoulou, A. Iliadi, T. Eliades and D. Kletsas, Eur. J. Orthod., 2016, DOI:

10.1093/ejo/cjw075, 258-263.

66. L. Blanchoin, R. Boujemaa-Paterski, C. Sykes and J. Plastino, Physiol Rev, 2014, 94,

235-263.

67. K. Burridge and C. Guilluy, Exp Cell Res, 2016, 343, 14-20.

Page 116: Design and Development of Cell Stretching Platforms for ...

115

Chapter 6: Pneumatically actuated cell-stretching array platform to

engineer cell pattern in vitro

Different cells within a human body are known to have different morphological and

behavioural patterns in vivo, depending on their anatomical location. The developed

cell stretching platforms detailed in preceding Chapters provide a systematic

standardised approach for analysing the strain requirement of the cells under

investigation and further in-depth mechanobiological investigation. However, to

further improve mechanobiological understanding, more accurate tools are needed to

accurately mimic the complex in vivo biological environment. Although many

approaches have been developed in recent years to study the effect of external

mechanical stimuli on the cellular behaviour, most of them have not explored the

ability of the mechanical stimuli to engineer the cell orientation to obtain predefined

patterned cell culture. Hence, this Chapter introduces a simple, yet effective

pneumatically actuated 4x2 cell stretching array for concurrently inducing a range of

cyclic strain onto cell cultures. The developed cell stretching platform can be used to

achieve pattered cell culture with predefined cellular alignment. Moreover, for

improving experimental flexibility an array based pneumatic actuation approach was

designed for this study. The study also aimed to improve the compatibility of the

developed platform by maintaining the dimension of the cell stretching platform the

same as the F bottom 96 well plates which was the simple and effective approach. This

Chapter report principle design, simulation and characterisation of the developed cell-

stretching platform with the preliminary test with fibroblast cells. The preliminary

observation of the circular alignment for the fibroblast cells confirms the suitability of

the developed platform to obtain predefined patterned cell culture.

Page 117: Design and Development of Cell Stretching Platforms for ...

116

Statement of Contribution to Co-authored Paper Submitted for Publication

Page 118: Design and Development of Cell Stretching Platforms for ...

117

6.1 Introduction

Cells within the human body are responsive to the surrounding mechanical stimuli.1-4

The

ability of cells to sense and respond to mechanical stimuli is known as a

mechanotransduction.5-7

At the cellular level, mechanical stimulus acts as a key driver in

regulating cellular functions. It is well known that such mechanical stimuli induce

biochemical responses, which are important to maintain normal cellular activities such as

cytoskeleton organisation, cellular orientation and pattern formation during tissue maturation

and regeneration.8-12

A mechanical stimulus triggers the formation of the stress fibres and the

reorganisation of cytoskeleton, which is directly coupled with extracellular matrix

(ECM) interface through focal adhesion assembly.13-15

This process allows cell to resist

mechanical deformation by creating new actin filament networks at the cell leading edge to

retract and extend protrusions.16, 17

This response determines the direction of the cell motion

and changes in the adhesive sites to resist the strain perturbation.18

Thus, the complex

mechanotransduction signalling process is critical for cells to maintain homeostasis and can

be used to achieve a desired cellular pattern in vivo.

For example, cardiac fibres and myocytes in the in heart align due to tensile strain caused by

heart beat and form the sheet of myocardium.19

Another example is the circumferential

alignment of the muscle fibres due to the shear caused by the changes in vascular wall sizes.20,

21 Moreover, many previous studies reported that abnormal mechanotransduction function

within the cells such as maladaptive remodelling of cytoskeleton may disturb the cellular

activities and lead to diseases.22, 23

Furthermore, some studies also claimed that external forces

acting on the cells also can influence the cell functionality by regulating the mechanical

properties of ECM. 24-26

Hence, the ability to achieve patterned cell culture with predefined

cellular alignment in vitro could serve as an important tool to mimic in vivo cellular

conditions. This technology will facilitate better understanding of the mechanotransduction

signalling process and will provide an insight on the role of mechanical stimuli in guiding

cytoskeleton reorganization and cell reorientation which could be critical in the field of tissue

engineering and regenerative medicines.

Indeed, various methods have been developed to apply external mechanical stimuli in-vitro to

understand and mimic the complex pathological in-vivo microenvironment of the cell.

Common techniques to induce mechanical force onto the cell culture use existing tools such

Page 119: Design and Development of Cell Stretching Platforms for ...

118

as atomic force microscope (AFM), micropipette or tweezes.27-30

In recent years, more

sophisticated cell-stretching designs have been introduced into the commercial market.

Flexcell (Flexcell International Corporation), ElectroForce (Bose Corporation) and Strex

Systems for cell Stretching (STREX Inc.) are some of the widely accepted commercial

platforms. Many cell-stretching studies included these commercial platforms due to its well

characterised parameters and ease of operation.31-33

Apart from these commercial approaches,

many custom made cell stretching platforms also have been reported utilizing elastic

membranes by the means of electromagnetic, pneumatic, optical or other micro actuation

techniques.34-38

For instance, Huang and Nguyen (2013) developed a multilayered PDMS

device with pneumatic actuators to deform the elastic membrane and introduce strain onto

cells cultured on the membrane. Kamotani, et al. (2008) reported an array of PDMS

microcells with deformable membrane, placed over piezoelectrically actuated pins.

Furthermore, Chang, et al. (2013) 41

utilised a motor actuated translation stage to stretch the

deformable substrate in one direction and to introduce strain onto the cells cultured on the

substrate. Electro-thermal, electrostatic and di-electrophoresis actuations have also been

explored for cell stretching.42-44

Although, all these existing cell stretching approaches have

particular advantage for their specific applications. Most of these studies were focused on

investigating the effect of the strain on cell behaviour. However, very few studies explored

the possibility of active cell manipulation using external mechanical stimuli for achieving

patterned cell culture with predefined cellular alignment, which closely mimics the in vivo

cellular microenvironment. Thus, a simple and effective method compatible with existing pre-

clinical tools and protocols is urgently needed for engineering patterned cell culture to better

mimic the in vivo cellular microenvironment.

This chapter presents a pneumatically actuated cell stretching array platform (Fig. 6.1a),

which can apply a range of strains onto the cell cultures to achieve predefined in vitro

circumferential cellular alignment, which is similar to the in vivo muscle fibre alignment in

vascular wall. The dimensions of the cell-stretching platform were maintained same as the

standard 96 well plates. Thus, the developed cell-stretching platform is compatible with the

general pre-clinical tools and protocols. The compatibility could be critical for designing

high-throughput cell-stretching assays. To prove the concept, the presently developed cell-

stretching platform incorporates an array of 4x2 wells. The wells in the each column was

utilised for introducing different predefined strain onto the cultured cells. As a first

Page 120: Design and Development of Cell Stretching Platforms for ...

119

Fig.6.1. The pneumatically actuated cell-stretching platform: (a) Schematic illustration of

the cell stretching platform in the OFF state. (b) Schematic representation of working

principle of the device with OP1 and OP3 in the ON state. (c) FEA model results

illustrating the OP1 and OP3 behaviour during ON state. (d) Lower PDMS layer (4x2)

with deformable output patterns and control pattern. (e) Actual multilayer PDMS cell-

stretching platform.

demonstration, we describe the design, simulation, fabrication and characterisation of the

developed cell stretching platform in detail. Further, for confirming the suitability of the

developed platform and to obtain patterned cell culture with predefined circumferential

alignment, we considered fibroblast cells for this study. We observed cellular response of the

fibroblast cells under different cyclic strain. Fibroblast is a unique cell type which shares the

plastic behaviour of mesenchymal phenotype and can be found in abundance within a human

body. Hence, it is considered as a prime candidate to understand the role of

mechanotransduction in altering the cell behaviour. Many studies have shown the capability

of the fibroblast cells to recognise and respond to the external cyclic strain and reorient

perpendicular to the strain direction45-49

. Moreover, some studies had claimed that fibroblast

cells also able to recognised strain gradient which is also an important guiding factor for cell

reorientation in order to attain mechanical stability50, 51

. Thus, respecting existing literature we

hypothesised that the cyclic stretching of the fibroblast cells with a ring shaped output

stretching pattern may facilitate patterned fibroblast cell culture with circumferential cellular

Page 121: Design and Development of Cell Stretching Platforms for ...

120

alignment similar to that of in vivo muscle fiber alignment in vascular walls under shear

stress.

Our preliminary observation of the cytoskeleton reorganisation and reorientation for fibroblast

cells are consistent with the previous studies and confirm the use of the developed cell

stretching platform to actively manipulate cell orientation and obtain patterned cell culture

with predefined cellular alignment. Furthermore, the developed cell stretching platform is

compatible with generic laboratory tools and protocols and ensures its adaption in pre-clinical

applications. The output cell pattern can be further optimised for different shapes and cellular

alignments. The developed platform can be scaled up to the standard 96 wells for screening of

cellular response in tissue engineering and regenerative medicine. Thus, the cell-stretching

platform presented here could be critical to obtain patterned cell culture for high-throughput

in-vitro assays.

6.2 Materials and methods

6.2.1 Device design and working principle

The pneumatically actuated cell-stretching array platform was designed and characterised to

achieve patterned cell culture with circumferential alignment. Furthermore, the platform was

particularly designed to induce different pre-defined strain onto the cells cultured in each well

to establish the relationship between strain amplitude and cellular alignment, Figure 6.1b,c.

The developed platform provides advantages such as ease of fabrication, ease of imaging,

compatibility with generic pre-clinical tools and predefined strain pattern. The array design

makes it suitable for the high throughput cellular response assays, Figure 6.1d.

Figure 6.1e depicts the actual image of the cell-stretching platform. The platform includes

three PDMS layers: (i) the lower layer consisting of the microfluidic channels and the ring

shaped pneumatic chamber, (ii) the middle layer consisting of a 200-µm deformable PDMS

membrane bonded with lower layer and finally, (iii) the upper layer consisting of the cell

culture wells which are aligned with the output patterns to facilitate cell seeding. For the

current study, output stretching pattern was varied for wells in each column to generate a

range of strain magnitudes with a single device. Each strain pattern consists of a toroidal

structure with a fixed outer diameter of 6.39±0.01 mm and variable inner radius (Ri). The

Page 122: Design and Development of Cell Stretching Platforms for ...

121

variable nature of the inner radius allows achieving variable strain amplitude onto the

deformable membrane upon actuation.

Figure 6.1d shows the deformable ring-shaped output pattern 1 (OP1) with the inner radius

Ri1=2 mm, deformable output pattern 2 (OP2) with the inner radius Ri2=1.5 mm and the

deformable output pattern 3 (OP3) with the inner radius Ri3=1 mm. The dimension of the

output patterns was maintained according to the F-bottom standard 96 well polystyrene

microplates dimension (Greiner Bio-One International GmbH) to maintain the compatibility

with generic tools. Similar to a standard 96 well plate, the centre to centre distance between

each pattern was maintained at 9.02±0.01mm with a fixed outer diameter of 6.39 mm for all

output patterns (OPs). Further, the depth of the cell culture wells was maintained at

5±0.5mm.

We used a pressure controller (Elveflow®) to actuate the OPs by providing positive pressure

of a given magnitude and frequency to the inlets of the cell stretching platform. Upon

actuation, the membranes over the output patterns inflate and deform into semi toroid shapes

which subsequently induce strain onto the cells cultured on the deformable OPs.

6.2.2 Design and optimisation

To assess the feasibility of the developed cell stretching platform and predict the system

behaviour, a finite element analysis (FEA) model of the platform was formulated in

COMSOL Multiphysics 5.2. In the optimisation phase, we optimised the geometry of the

platform by varying one geometric parameter and maintaining the other parameters constant.

The variable inner radius for the output pattern was one of the key parameter for the

optimisation, as the possible strain amplitudes can be determined. Assuming a constant

pressure, the reference geometry was adapted in COMSOL with the total volume of

9x9x10.2 mm3.

The outer diameter was fixed at 6.39 mm, and the inner diameter and height were maintained

at 3 mm and 0.1 mm, respectively. Furthermore, the height of the channel and the membrane

thickness were kept at 0.1 mm and 0.2 mm, respectively. The material was selected as

polydimethyl-siloxane (PDMS) for the entire geometry with the Young’s modulus as 750 Kpa

and Poisson’s ratio as 0.49 of PDMS mixed with crosslinker in 10 to 1 volume ratio. In the

Page 123: Design and Development of Cell Stretching Platforms for ...

122

0.5 1 1.5 2 2.5Radius of the post (mm) R

i

0.3

0

0.1

0.2

0.4

0.5

0.6

Ma

x.

dis

pla

ce

men

t of

mem

bra

ne

(

mm

) D

m

Pressure (P)= 80 mbar

Deformable OP 3

Deformable OP 2

Deformable OP1

200ummembrane

Fixed PatternedLower layer

Fixed inner post

StrainRi = 2mm

Dm= 0.158mm

Op1

Inflated membraneRi = 1mm

Dm= 0.413mm

Z

YX

Op3

Ri = 1.5mm

Dm= 0.218mm

Op2

Fig.6.2 Simulation results of the maximum planar displacement versus the inner post

radius (Ri) versus (Inset: optimised geometry for the deformable OP1, OP2 and OP3, and

the membrane deformation results for OPs with applied pressure of 80 mbar).

next step, necessary boundary conditions were implemented for the reference FEA model by

allowing free deformation of the output pattern and fixing all other parts. A positive pressure

of 80 mbar was applied to the membrane to represent the pneumatic actuation. Finally,

utilising parametric sweep function in COMSOL the inner radius of the output pattern was

varied from 0.25 mm to 5.5 mm with a step of 0.25 mm to obtain the corresponding

maximum planar displacement of the membrane.

Figure 6.2 shows the membrane displacement as function of the inner radius. The simulation

provided a wide range of strain amplitudes that could be achieved with the current cell

stretching platform. Such a wide range allows for establishing the relationship between

cellular alignment and applied strain. In the current study, we kept the inner diameter for the

OP1, OP2 and OP3 at 3 mm, 2 mm and 1 mm, respectively. The insets in Figure 6.2 show the

simulation results for the selected OPs.

Page 124: Design and Development of Cell Stretching Platforms for ...

123

Cover slip

(b) 200um spin coated PDMS

Plasma Bonding

(a)

SU-8 PDMS

Silicon wafer

Lower layer mould foroutput pattern.

e)(

Actuation cavity

Punched well

200um Membrane

Peeling offc)(

PDMS post

PDMS

(d) Upper layer mould for culture chamber

5m

m

(f)

Culture chamber

Multilayer PDMS device

Fig.6.3 Fabrication process for cell stretching platform.

6.2.3 Fabrication process

The software CleWin 3.0 was utilised to design the cell-stretching platform. The optimised

design was plotted on the 28 mmx22 mm plastic mask for pattern replication. The fabrication

of the multilayered stretching platform mainly involved the soft lithography technique. Figure

6.3 illustrates the basic fabrication steps. The process comprised of three major steps.

In the first step, a clean and pre-baked (80o for 30 min.) silicon wafer was spin coated with a

100-m layer of photoresist (SU-8 50) at 3,300 rpm for 30 sec. A series of pre-baking steps

(65 °C for 10 min, 95 °C for 30 min, 50 °C for 30 mins) were performed to avoid SU-8

cracking. Subsequent UV exposure (9.8 mW/cm2 for 35 to 40 seconds) allowed transferring

the device pattern to the wafer. Finally, the SU-8 was post baked (80 °C for 30 mins) and the

mask was developed in 1-methoxy-2-propanol acetate for 15 minutes. The same procedure

was followed to prepare the master mould for the upper PDMS layer with the cell culture

wells.

In step two, a clear degassed PDMS and cross linker (Sylgard 184 elastomer kit, Dow

Corning) mixture (10:1 volume ratio) was poured onto the master moulds, so that a thickness

of 2-2.5mm and 5mm was achieved for the lower and upper layer masks, respectively. The

mixture was then cured at 80oC for 2 hour in a vacuum oven. The cured PDMS layers were

Page 125: Design and Development of Cell Stretching Platforms for ...

124

inspected and carefully peeled off from the master mould. Furthermore, the upper PDMS

layer was carefully punched to form the cell culture wells.

In the final step, the middle layer, i.e. the 200-µm deformable PDMS membrane, was

fabricated and all three layers were aligned and plasma bonded to from the final cell-

stretching platform. A clear degassed PDMS-crosslinker mixture (10:1 volume ratio) was spin

coated at 400 rpm for 2 mins to achieve the 200-µm thickness. Prior to plasma bonding, the

deformable PDMS membrane and the lower PDMS layer with output pattern was thoroughly

cleaned with isopropanol and DI water to achieve a dust-free surface. The surface of both

membrane and PDMS layer was then activated using oxygen plasma (Harrick Plasma) for 45

seconds before bonding. The bonded device was then cured in a vacuum oven at 80oC for 1

hour. The same plasma bonding procedure was repeated again to bond the prepared device

with the upper PDSM layer.

6.2.4 Cell culture

The cell-stretching platform was sterilised with ethanol (80%) and washed with PBS for three

times followed by the 20 minutes UV exposure. Furthermore, the cell-stretching platform was

treated with the media and incubated for 1 hour at 37oC and 5% CO2 before seeding the cells

to enhance biocompatibility.

Fibroblast cells were cultured in a T75 cell culture flask with DMEM/F12 (Gibco, Thermo

Fisher Scientific) medium enriched with 10% FBS and 1% penicillin. The cells were

maintained in a standard incubator at 37oC and 5% CO2 to ensure proper cell growth. For

seeding the device, 70% confluent T75 flask was trypsinised with TrypLE Express (Thermo

Fisher Scientific) and resuspended into the fresh media. Aliquots of optimised seeding density

of 10x103 Cells/150 µl were used to seed each well of the device (i.e. total 80x10

3

Cells/1200ul). To ensure the cell adhesion and adequate cell growth onto the membrane, the

device was incubated at 37oC with 5% CO2 for 24 hours. Finally, following literature

reporting for the fibroblast cells, predefined cyclic strain of 0.8% (OP1), 1.2% (OP2) and

2.2% (OP3) with 80 mbar, 0.1Hz frequency at 50% duty cycle was introduced onto the cell

culture for 2 hours to obtained circumferential alignment of the fibroblast cells24, 52-54

. Three

biological repeats were carried out with the same strain parameters to confirm the

repeatability of the experimental results.

Page 126: Design and Development of Cell Stretching Platforms for ...

125

Fig.6.4. Developed algorithm to detect the inflated membrane to fit the curve and

calculates the displacement and strain parameters. (a) Input: Reference Image and scale

(pixels per mm). (b) Image enhancement process: Normalised contrast and sharpening. (c)

Thresholding and Edge detection using canny. (d) Dilation operation and noise removal

for curve detection. (e) Fill the holes to enhance mask calculations by eliminating minor

curves. (f) Calculate mask and get minimum Y coordinates along the curve for selection.

(g) Generate symmetric reflected curve along major axis for curve fitting. (h) Apply least

Square criterion for curve fitting. (i) Calculate curve parameters and display results.

6.2.5 Cell Fixing, immunofluorescence staining and imaging

After two hours of stretching, the cells were immediately fixed with 4% PFA and then washed

with PBS for three times. Further, the cells were kept in cold PBS solution at 4oC. For the

quantification, cell orientation was then recorded with Nikon Eclipse Ts2 inverted phase

contrast microscope using the 10x magnification. Actin and nuclei were stained respectively

with ActinGreenTM 488 and NucBlueTM ReadyProbeTM Reagent (ThermoFisher Scientific)

to observe the cytoskeleton reorganisation. The standard protocol was optimised and adapted

to suit the developed cell stretching platform. Finally, we utilised an inverted fluorescent

microscope (Olympus IX73) with 20X magnification to observe the cell nuclei and actin

fibres.

6.3 Results and discussion

6.3.1 Device characterisation

Page 127: Design and Development of Cell Stretching Platforms for ...

126

0

0.1

0.2

0.3

0.4

0.5

0.6

Dis

pla

ce

men

t (

mm

)

Experimentalresults

OP 1OP 2OP 3

Simulationresults

OP 1OP 2

OP 3

00.25

0.5

0.55Op1 Op2 Op3

Pressure = 150mbar

20 40 60 80 100 120 140

Pressure (mbar)

10(a)

0

1

2

3

4

5

6

Cir

cu

mfe

ren

tial S

tra

in (

%)

OP 1OP 2OP 3

Experimental results

Simulationresults

OP 1OP 2OP 3

Control

Op1

Op

2O

p3

Pre

ssure

co

ntr

olle

rP

um

p

PDMS device

XY stageMotorised

Camera

XY

20 40 60 80 100 120 140

Pressure (mbar)

0

(b)

Fig.6.5 Comparison of the simulation and experimental data: (a) Maximum displacement

versus pressure. (Inset: simulation results with pressure of 150 mbar); (b) Circumferential

strain versus pressure. (Inset: experimental setup).

Well defined strain is an important aspect for achieving cell culture with predefined cellular

alignment. Thus, insight into the strain experienced by the cells under investigation based on

their position is important. Following, we numerically and experimentally investigated strain

on the deformable output patterns.

Experimental approach

For experimental strain characterisation, the deformable output patterns (OP1, OP2 and OP3)

were actuated by positive pressure ranging from 10 mbar to 150 mbar with a step of 10 mbar

using a pressure controller. A digital camera (EO® Edmund Optics) with 1X magnification

lens was placed onto the xy-stage (Zaber Technologies) and the side view along xz- plane was

recorded to observe the inflation of the output patterns. A custom made algorithm was

developed for detecting the membrane inflation, fiting an ellipse into the detected curve and

obtaining the displacement and strain parameters.

The developed algorithm was implemented in Matlab R2016a using the image processing

toolbox and curve fitting techniques. The detailed flow chart with a sample step output for the

implemented algorithm is shown in Figure 6.4. We repeated the experiments three times for

each pressure point, to obtain reliable results from the image analysis. A minimum of three

Page 128: Design and Development of Cell Stretching Platforms for ...

127

Dis

pla

cem

en

t o

f m

em

bra

ne

(m

m)

0

0.1

0.2

0.3

0.4

Experimental results(a)

0

OP 1

OP 2

OP 3 D = 0.382mmm

D = 0.165 mmm

D = 0.197 mmm

OP 1OP 2OP 3

P= 80 mbar

0

0.1

0.2

0.3

0.4

Simulation results(b)

0

0.1

0.2

0.3

0.4

OP 2

OP 3

D = 0.218 mmm

D = 0.413 mmm

D = 0.158 mmm OP 1

OP 1OP 2OP 3

P= 80 mbar

0 1 2 3 4 5 6X coordinate length (mm)

Fig.6.6 Comparison of the membrane deformation for OP1, OP2 and OP3 with pressure input

80 mabr. (a) Experimental results (actual membrane deformation in XZ plane) (b) Simulation

results (FEA model results for membrane deformation in XZ plane) .

image frames from each video was analysed. The averaged result for each pressure point was

obtained to assure the repeatability of the data.

For computational simplicity, the ideal condition of an uniform pressure distribution on a

deformed membrane with negligible thickness variation was assumed for the strain

evaluation. Considering the thin-walled membrane and obtained elliptical fit of the membrane

inflation, circumferential strain εc was calculated utilising the Hoops stress (σc)

relationship:55, 56

σc =Pri

t(1 −

ri3

a2(2ri+t)) (1)

εc = σc

E, (2)

where σc is the stress (in mbar), P is the input pressure (in mbar), ri is the internal dimeter of

the inflated membrane (in mm), 𝑡 =0.2mm is the thickness of the membrane, 𝑎 is the height

Page 129: Design and Development of Cell Stretching Platforms for ...

128

Fig.6.7 Output membrane deformation pattern of OP1: (a) Experimental results. (b)

Simulation results.

along minor axis (in mm), 𝐸 = 7500mbar is the Young’s modulus of the PDMS membrane

and εc is circumferential strain.

Numerical approach

The numerical verification of experimental data utilised a reference 3D FEA model with OP1

(Ri1 = 1mm), OP2 (Ri2 = 1.5mm) and OP3 (Ri3 = 2mm). All material properties and boundary

conditions were fixed, while the applied pressure was varied from 10 mbar to 150 mbar with a

step of 10 mbar to obtain the corresponding maximum planar displacement and

circumferential strain of all three OPs.

Figure 6.5a compares the experimental and simulation data of the maximum displacement of

the membrane. As expected, all three OPs shows a linear relationship between membrane

displacement and the pressure. The simulation and experimental data agree well and provide

an acceptable average error variance of 1.63% for OP1, 2.37% for OP2 and OP3 for 3.84%.

Furthermore, the maximum applied pressure of 150 mbar results in the maximum planar

displacement of 0.2 mm for OP1, 0.32 mm for OP2 and 0.53 mm for OP3. In the next step,

the circumferential strain was compared.

Page 130: Design and Development of Cell Stretching Platforms for ...

129

Fig. 6.8 Averaged cell orientation in degrees over the ROI for OP1 and OP2 with

corresponding 10X cell images of each ROI and the COMSOL strain distribution map.

Scale bar is 100 µm.

Figure 6.5b shows the experimental and simulation data of the circumferential strain obtained

with the same range of applied pressures. The data show that at 80 mbar the circumferential

strains of OP1, OP2 and OP3 are 0.8±0.023%, 1.2±0.07% and 2.2±0.081% respectively.

The obtained deflection and strain data served as a guideline for the selection of the

applied pressure. Considering the fibroblasts in the later experiments, we selected 80 mbar as

the input pressure for the cell-stretching platform. The next step is determining the

deformation pattern of the membrane under 80 mbar, utilising the existing experimental tools

and optimised FEA model. Figure 6.6 depicts the experimental and simulation results of the

shape of the deformed ring-shaped membrane. The semi-elliptical deformation of the different

OPs can be clearly observed.

As the deformation patterns are axial-symmetric, the experimentally obtained 2D curve of the

deformed membrane was virtually rotated along minor axis using Matlab R2016a to obtain

the 3D representation of the membrane. For comparison, the corresponding FEA model was

solved. Figure 6.7 compares the experimental and numerical shape of the membrane of OP1

Page 131: Design and Development of Cell Stretching Platforms for ...

130

Fig. 6.9 Averaged cell orientation in degrees over the ROI for OP3 (a) and Control (b)

with corresponding 10X cell images of each ROI and the COMSOL strain distribution

map. Scale bar is 100 µm.

under 80 mbar. Experimental and simulation results agree well and show the a maximum

planar displacement (Dm) of 0.165 mm and 0.159 mm with an average error variance of 2.5%,

respectively. The 3D data show that the output pattern deforms into a semi-toroidal shape.

The results suggest that over the central circumferential region with maximum planar

displacement, cells would experience approximately the same amount of strain. Thus, we

hypothesised that cells in this circumferential region of homogenous strain will experience

approximately the same mechanical cues. The physiological or morphological changes

observed after stretching would mainly depend on the strain amplitude.

6.3.2 Cell orientation

For the quantification of cell orientation, each cell culture well was virtually divided into 8

equal quadrants of 45o each i.e. region of interests (ROI). Images from each ROI were

considered for the analysis. To enhance the features and correct the background noise in phase

contrast images, basic post processing steps including application of Fast Fourier

Transformation (FFT) bandpass filter, sharpening and enhancing image contrast were carried

Page 132: Design and Development of Cell Stretching Platforms for ...

131

1.2% Strain (Op2)

No Strain (Control)0.8% Strain (Op1)

2.2% Strain (Op3)

0

20

40

60

80

Percentage of strain

0 2.2 1.2 0.8

***

***

*

n.s

##

Perc

enta

ge

of

cells

Fig. 6.10 Percentage of cells in given direction

over the ROI and ANOVA analysis for OP1,

OP2, OP3 and Control. The cell orientation

varies according to strain amplitude (***p<

0.001, and *p< 0.05 versus control; ##p< 0.01

versus the lowest strain amplitude; n.s., not

significant). Results are expressed as means ±

s.d. (n = 3) and data were analysed using one-

way ANOVA followed by post hoc Bonferroni’s

test.

out with ImageJ tools. The obtained binary image was further utilised for the analysis of cell

orientation. In each ROI, orientation of the 100 cells were analysed to estimate an average

orientation of the cells in each predefined quadrant. Figures 6.8 and 6.9 shows the average

cell orientation for each ROI of OP1, OP2, OP3 and control.

The data and the images clearly show that the cells recognise the external mechanical cues

and reorient themselves in an optimum direction to avoid the strain.57-60

As expected, the

fibroblast cell reoriented perpendicular to the strain direction.61, 62

Moreover, the cells

appeared to recognise the homogenous strain gradient and realign themselves along this

region. Although, alignment of the cells with respect to the strain pattern was observed for all

three output patterns, it was interesting to note that cell alignment along the strain profile

increases with the increasing strain amplitude. This behaviour of fibroblasts consistent with

previous reports.24, 50, 63

An average cell orientation for the three

output patterns (OP1, OP2 and OP3)

were 131.3±14.9o for the ROI 1&2, 221

±11.5o for ROI 3&4, 295.89±13.43

o for

ROI 5&6 and 66.82±10.41o for ROI

7&8, Figures 6.8 and 6.9a. In contrast,

cells distributed randomly without strain

and did not show any preferred

orientation or significant alignment,

(Fig. 6.9b).

Furthermore, we calculated the

percentage cells aligned in

circumferential pattern for each ROI to

establish the relationship between strain

amplitude and cell orientation. Figure

6.10 compares the percentage of aligned

cells for the different strain amplitude

with one-way analysis of variance

(ANOVA) results. The ANOVA results

Page 133: Design and Development of Cell Stretching Platforms for ...

132

Fig. 6.11 Cytoskeleton organisation for OP1 (a), OP2 (b) and OP3 (c) and Control (d) with

20x Objective (scale bar: 50um) after 2 hours.

suggested that all three output patterns exhibit significant cell orientation compared to the

control (OP1: p<0.001, OP2: p<0.001 and OP3: p<0.05). However, the cell orientation of

OP2 and OP3 is significantly higher (p<0.01) than that of OP1. Moreover, no significant

difference between cell orientation was observed for the OP2 and OP3 (n.s) and clear

circumferential alignment of the fibroblasts can be seen (Fig.6.10).

In depth molecular or biological assays for the physiological study of the stretched cells is

outside the scope of the current work. Thus, we adapted the observational approach and

detected noticeable cell morphology changes for all OPs to confirm any changes due to

different strain amplitude. Interestingly, we noted that cells enhanced their cell-cell

connections and formed clusters under a high strain amplitude (Fig. 6.11). The gradual

increase in cell-cell connections and rearrangement of stress fibres with increasing strain

amplitude is evident. The images also indicated that reorganization of actin stress fibres is

critical for promoting cell adhesion dynamic and to induce morphological changes.24, 64

We observed that the stress fibre network aligns parallel to cell orientation. This observation

is in line with previous insights, that these stress fibres serve as a major tension-bearing

structures and initiate cell reorientation.65

,66

The interrelationship between actin stress fibres

and ECM is crucial for cell to transmit the physical forces onto its neighbouring cells and to

expand cell-cell connections.50, 67

Moreover, our observation further strengthens the argument

that fibroblasts are be able to sense and distinguish between the strain amplitudes and adapt

their adhesion site according to the strain direction to attain the optimum orientation.68, 69

We

hypothesise that enhanced cell-cell connection, reorientation and realignment of the

fibroblasts can be defined as a mechanosensing process linking actin stress fibres with focal

Page 134: Design and Development of Cell Stretching Platforms for ...

133

adhesion.63, 70

This process allows cells to reorient in an optimum direction to withstand the

tension and to maintain their mechanical stability.71

Considering the above preliminary observations, we could conclude that pattered cell culture

can be obtained with low strain (0.8-1.2%) amplitude without any noticeable physiological

and morphological changes. Our experimental results confirm the capability of our cell-

stretching platform to actively manipulate the cell orientation to achieve patterned cell culture

with circumferential cellular alignment.

6.4 Conclusions

In summary, a novel cell-stretching array platform compatible with the generic tools is

designed, fabricated and characterised to obtain patterned cell culture with predefined

circumferential cellular alignment in vitro. We observed with the platform the widely reported

behaviour of fibroblasts. The cells reorient and realign approximately perpendicular to the

stretching direction. We also demonstrated that the cyclic strain can be utilised to actively

manipulate the cellular alignment and achieve patterned cell culture with predefined cellular

alignment. Moreover, we examined the cell reorientation with different strain amplitudes.

Our observations suggest that the cell orientation increases with increasing strain amplitude.

However, noticeable cell clustering was observed with higher strain amplitude. Although

further optimisation is needed, the preliminary results allow us to hypothesise that low strain

amplitude of 1.2% could pattern fibroblast cell culture without noticeable physiological and

morphological changes.

The developed cell-stretching platform provides a new direction of bioengineering utilising

cells mechanotransduction capability to actively manipulate the cellular alignment and

achieve patterned cell culture similar to that of in vivo cell patterns. The developed platform

can be easily optimised by redesigning the desired OPs to achieve a variety of predefined cell

culture patterns, which could be critical in tissue remodelling and regenerative medicine.

Furthermore, the ease of handling and the compatibility of the developed platform with

existing tools and protocols make it attractive for in-depth molecular and biological analysis

involving mechanotransduction. Finally, the developed cyclic cell stretching approach may be

an attractive alternative strategy to understand how various mechanical cues such as strain and

Page 135: Design and Development of Cell Stretching Platforms for ...

134

strain gradient influence the cell alignment and orientation which provide insights on cellular

level interactions between cells and their immediate microenvironment.

Page 136: Design and Development of Cell Stretching Platforms for ...

135

6.5 References

1. B. D. Hoffman and J. C. Crocker, Annu. Rev. Biomed. Eng., 2009, 11, 259-288.

2. R. P. Brandes, N. Weissmann and K. Schroder, Antioxid. Redox Signal., 2014, 20,

887-898.

3. J. H. Wang and B. P. Thampatty, Biomech Model Mechanobiol, 2006, 5, 1-16.

4. C. L. Happe and A. J. Engler, Circ. Res., 2016, 118, 296-310.

5. F. H. Silver and L. M. Siperko, Crit. Rev. Biomed. Eng., 2003, 31, 255-331.

6. A. W. Orr, B. P. Helmke, B. R. Blackman and M. A. Schwartz, Dev. Cell, 2006, 10,

11-20.

7. C. S. Chen, J. Cell Sci., 2008, 121, 3285.

8. D. J. Leong, J. A. Hardin, N. J. Cobelli and H. B. Sun, in Skeletal Biology and

Medicine Ii: Bone and Cartilage Homeostasis and Bone Disease, ed. M. Zaidi,

Blackwell Science Publ, Oxford, 2011, vol. 1240, pp. 32-37.

9. K. Harshad, M. Jun, S. Park, M. J. Barton, R. K. Vadivelu, J. St John and N.-T.

Nguyen, Biomed. Microdevices, 2016, 18, 45.

10. B. D. Hoffman, C. Grashoff and M. A. Schwartz, Nature, 2011, 475, 316-323.

11. S. Y. Chew, R. Mi, A. Hoke and K. W. Leong, Biomaterials, 2008, 29, 653-661.

12. S. Hoehme, M. Brulport, A. Bauer, E. Bedawy, W. Schormann, M. Hermes, V. Puppe,

R. Gebhardt, S. Zellmer, M. Schwarz, E. Bockamp, T. Timmel, J. G. Hengstler and D.

Drasdo, Proc Natl Acad Sci U S A, 2010, 107, 10371-10376.

13. J. Yan, M. Yao, B. T. Goult and M. P. Sheetz, Cell. Mol. Bioeng., 2015, 8, 151-159.

14. R. O. Hynes, Nat. Rev. Mol. Cell Biol., 2014, 15, 761-763.

15. M. A. Schwartz, Cold Spring Harb Perspect Biol, 2010, 2, a005066.

16. D. A. Fletcher and R. D. Mullins, Nature, 2010, 463, 485-492.

17. J. L. Hoon, M. H. Tan and C. G. Koh, Cells, 2016, 5.

18. F. Klingberg, B. Hinz and E. S. White, The Journal of pathology, 2013, 229, 298-309.

19. P. Helm, M. F. Beg, M. I. Miller and R. L. Winslow, Ann N Y Acad Sci, 2005, 1047,

296-307.

20. M. B. Chan-Park, J. Y. Shen, Y. Cao, Y. Xiong, Y. Liu, S. Rayatpisheh, G. C. Kang

and H. P. Greisler, J. Biomed. Mater. Res. A, 2009, 88, 1104-1121.

21. M. Morioka, H. Parameswaran, K. Naruse, M. Kondo, M. Sokabe, Y. Hasegawa, B.

Suki and S. Ito, PLoS One, 2011, 6, e26384.

Page 137: Design and Development of Cell Stretching Platforms for ...

136

22. C. C. DuFort, M. J. Paszek and V. M. Weaver, Nat. Rev. Mol. Cell Biol., 2011, 12,

308-319.

23. D. E. Ingber, Ann Med, 2003, 35, 564-577.

24. Y. Cui, F. M. Hameed, B. Yang, K. Lee, C. Q. Pan, S. Park and M. Sheetz, Nat.

Commun., 2015, 6, 6333.

25. A. E. Ekpenyong, N. Toepfner, E. R. Chilvers and J. Guck, Biochim. Biophys. Acta,

2015, 1853, 3105-3116.

26. H. Oda, T. Konno and K. Ishihara, Biomaterials, 2013, 34, 5891-5896.

27. Y. H. Tan, D. Sun, J. Z. Wang and W. H. Huang, IEEE Trans. Biomed. Eng., 2010,

57, 1816-1825.

28. S. Nishimura, K. Seo, M. Nagasaki, Y. Hosoya, H. Yamashita, H. Fujita, R. Nagai and

S. Sugiura, Prog. Biophys. Mol. Biol., 2008, 97, 282-297.

29. N. J. Sniadecki, C. M. Lamb, Y. Liu, C. S. Chen and D. H. Reich, Rev. Sci. Instrum.,

2008, 79, 044302.

30. O. Chaudhuri, S. H. Parekh, W. A. Lam and D. A. Fletcher, Nat. Methods, 2009, 6,

383-387.

31. S. Dhein, A. Schreiber, S. Steinbach, D. Apel, A. Salameh, F. Schlegel, M. Kostelka,

P. M. Dohmen and F. W. Mohr, Prog. Biophys. Mol. Biol., 2014, 115, 93-102.

32. Z. He, R. Potter, X. Li and M. Flessner, Adv. Perit. Dial., 2012, 28, 2-9.

33. S. Higgins, J. S. Lee, L. Ha and J. Y. Lim, Biores Open Access., 2013, 2, 212-216.

34. Q. Wang, X. Zhang and Y. Zhao, Sensors Actuators B: Chem., 2014, 194, 484-491.

35. W. W. Ahmed, T. Wolfram, A. M. Goldyn, K. Bruellhoff, B. A. Rioja, M. Moller, J.

P. Spatz, T. A. Saif, J. Groll and R. Kemkemer, Biomaterials, 2010, 31, 250-258.

36. S. P. Arold, J. Y. Wong and B. Suki, Ann. Biomed. Eng., 2007, 35, 1156-1164.

37. K. Sato, S. Kamada and K. Minami, IJMS, 2010, 52, 251-256.

38. H. Kamble, M. J. Barton and N.-T. Nguyen, Sensors and Actuators A: Physical, 2016,

252, 174-179.

39. Y. Huang and N. T. Nguyen, Biomed. Microdevices, 2013, 15, 1043-1054.

40. Y. Kamotani, T. Bersano-Begey, N. Kato, Y. C. Tung, D. Huh, J. W. Song and S.

Takayama, Biomaterials, 2008, 29, 2646-2655.

41. Y. J. Chang, C. J. Tsai, F. G. Tseng, T. J. Chen and T. W. Wang, Nanomedicine, 2013,

9, 345-355.

Page 138: Design and Development of Cell Stretching Platforms for ...

137

42. Y. Iwadate and S. Yumura, Biotechniques, 2009, 47, 757-767.

43. I. Guido, C. Xiong, M. S. Jaeger and C. Duschl, Microelectron. Eng., 2012, 97, 379-

382.

44. Z. L. Shen, M. R. Dodge, H. Kahn, R. Ballarini and S. J. Eppell, Biophys. J., 2008, 95,

3956-3963.

45. P. Feng, X. Li, W. Chen, C. Liu, S. Rong, X. Wang and G. Du, Biomedical

engineering online, 2016, 15, 63.

46. S. Chagnon-Lessard, H. Jean-Ruel, M. Godin and A. E. Pelling, Integr. Biol. (Camb.),

2017, DOI: 10.1039/c7ib00019g.

47. Y. Chen, A. M. Pasapera, A. P. Koretsky and C. M. Waterman, Proc Natl Acad Sci U

S A, 2013, 110, E2352-2361.

48. S. C. Tyagi, K. Lewis, D. Pikes, A. Marcello, V. S. Mujumdar, L. M. Smiley and C.

K. Moore, J Cell Physiol, 1998, 176, 374-382.

49. T. W. Gilbert, A. M. Stewart-Akers, J. Sydeski, T. D. Nguyen, S. F. Badylak and S. L.

Woo, Tissue Eng., 2007, 13, 1313-1323.

50. L. Wang, Y. Li, B. Chen, S. Liu, M. Li, L. Zheng, P. Wang, T. J. Lu and F. Xu, ACS

applied materials & interfaces, 2015, 7, 15088-15097.

51. S. Chagnon-Lessard, H. Jean-Ruel, M. Godin and A. E. Pelling, Integr Biol, 2017, 9,

607-618.

52. H.-S. Yu, J.-J. Kim, H.-W. Kim, M. P. Lewis and I. Wall, J. Tissue Eng., 2015, 6,

2041731415618342.

53. E. H. Kim, N. Oh, M. Jun, K. Ko and S. Park, BioChip Journal, 2015, 9, 306-312.

54. F. Boccafoschi, M. Bosetti, S. Gatti and M. Cannas, Cell adhesion & migration, 2007,

1, 124-128.

55. J. C. Wang, 2005.

56. P. A. McHALE and J. C. Greenfield, Circ. Res., 1973, 33, 303-312.

57. H. Abiko, S. Fujiwara, K. Ohashi, R. Hiatari, T. Mashiko, N. Sakamoto, M. Sato and

K. Mizuno, J Cell Sci, 2015, 128, 1683-1695.

58. U. Faust, N. Hampe, W. Rubner, N. Kirchgessner, S. Safran, B. Hoffmann and R.

Merkel, PLoS One, 2011, 6, e28963.

59. C. Tamiello, C. V. Bouten and F. P. Baaijens, Sci. Rep., 2015, 5, 8752.

60. J. Qian, H. Liu, Y. Lin, W. Chen and H. Gao, PLoS One, 2013, 8, e65864.

Page 139: Design and Development of Cell Stretching Platforms for ...

138

61. Y. Morita, S. Watanabe, Y. Ju and S. Yamamoto, Cell Biochem. Biophys., 2013, 67,

1249-1259.

62. A. Pavesi, G. Adriani, M. Rasponi, I. K. Zervantonakis, G. B. Fiore and R. D. Kamm,

Sci. Rep., 2015, 5, 11800.

63. S. Chagnon-Lessard, H. Jean-Ruel, M. Godin and A. E. Pelling, Integr. Biol. (Camb.),

2017, 9, 607-618.

64. H. Ngu, Y. Feng, L. Lu, S. J. Oswald, G. D. Longmore and F. C.-P. Yin, Annals of

biomedical engineering, 2010, 38, 208-222.

65. N. K. Weidenhamer and R. T. Tranquillo, Tissue Engineering Part C: Methods, 2013,

19, 386-395.

66. Q. Wang, H. Huang, K. Wei and Y. Zhao, Biotechnology and bioengineering, 2016,

113, 2191-2201.

67. A. Tondon and R. Kaunas, PLoS One, 2014, 9, e89592.

68. S. Fujiwara, K. Ohashi, T. Mashiko, H. Kondo and K. Mizuno, Mol. Biol. Cell, 2016,

27, 954-966.

69. D. Tremblay, S. Chagnon-Lessard, M. Mirzaei, A. E. Pelling and M. Godin,

Biotechnol. Lett, 2014, 36, 657-665.

70. K. Burridge and C. Guilluy, Experimental cell research, 2016, 343, 14-20.

71. L. Wang, Y. Li, B. Chen, S. Liu, M. Li, L. Zheng, P. Wang, T. J. Lu and F. Xu, ACS

applied materials & interfaces, 2015, 7, 15088-15097.

Page 140: Design and Development of Cell Stretching Platforms for ...

139

Chapter 7: Conclusions and perspectives

The developed platform, as describe ostentatiously in this thesis, extends the capability of

existing cell stretching platforms and provides a systematic approach for extensive

mechanobiological investigation, which could also be easily customised for specific medical

application.

The overall focus of this thesis was to design and develop novel cell stretching platforms

which are easy in fabrication, operation and compatibility with generic pre-clinical tools and

also facilitates capability for further optimisation to achieve standardised high throughput

cell-stretching assays w ith high biological relevance. Chapter 1 provides a background and

summarises the specific motivation for the research work presented in this thesis. The

extensive literature review presented in Chapter 2 provides an understanding of the existing

cell stretching platforms and facilitates perspective in the development of the new approaches

for further improvements. Chapters 3 to 6 focus on the design and development of the novel

cell stretching platforms to establish a systematic approach for extensive mechanobiological

investigation.

In brief, the thesis describes the design and development of three novel cell stretching

platforms that can induce controlled external strain onto the cell culture to observe cellular

responses. These platforms offer advantages such as ease of fabrication, ease of operation,

compatibility with pre-clinical tools, controlled strain profile, high throughput and high

biological relevance. Furthermore, each cell stretching platform is designed to facilitate a

specific application, which serves as a tool to establish the systematic approach for extensive

mechanobiological investigations. For instance, single sided electromagnetically actuated

cell-stretching platforms described in Chapter 3 provide the capability to observe cellular

response under the same biological conditions for different strain amplitudes within the same

field of view which is important for determining the strain requirement for the cells under

investigation whereas, double-sided electromagnetically actuated cell stretching platform

described in Chapter 5 facilitates static and dynamic stretching modes which further provides

a range of parametric tools for in-depth mechanobiological investigation. Additionally,

pneumatically actuated cell-stretching array platform described in Chapter 6provides the

ability to concurrently induce strain onto the cell cultures (2x4) to engineer cellular pattern in

Page 141: Design and Development of Cell Stretching Platforms for ...

140

vitro similar to the in vivo cellular condition and facilitates high biological relevance. All the

developed cell stretching platforms provide simple cell harvesting process after stretching

which is critical for in depth molecular and biological analysis.

The developed platforms serve as tools to establish a systematic approach for

mechanobiological investigations with standardised cell-stretching assays. Thus, the work

presented in this thesis further advance the state-of-the-art of cell stretching and provides a

promising solution with systematic approach for the detailed investigation of the

mechanobiological responses in standard clinical environment. The investigation enabled by

the developed cell stretching platforms may facilitate better understanding of cascade of

cellular signalling and pathways involved in mechanobiological responses.

Following, some of the future directions that could potentially improve versatility of the cell

stretching platforms are briefly summarised.

Existing cell stretching platforms both custom made or commercial, are restricted to simple

methods to induce mechanical strain on to the cells culture. The existing cell stretching

platforms reported in literature are usually capable to induce constant strain in single

experiments. In Chapter 6, we introduced a simple and effective approach to improve

throughput by adopting an array based design with dimensions same as the 96 well plate.

4

1

2

3

B

A

C

D

C

Flow channel

Cells seeding

Flow channel

MembraneVacuumchamber

Vacuum Vacuum

Stretched cells Stretched membrane

96 well plate with micro-fludic device

Micro-fludic Device at the bottom of the well

Fig. 7.1: 96-well plates with cell stretching platform for high-throughput screening.

Page 142: Design and Development of Cell Stretching Platforms for ...

141

Currently, the developed approach facilitates a 2x4 array with three different strain profiles.

However, the same concept could be scaled up in the future to the actual 96-well or 384-well

microplates, Fig.7.1. This integration of the microfluidics platform with commercial high-

content imaging systems such as Pperetta CLSTM

System (PerkinElmer) will enable

researchers to perform large-scale biological screening with advanced clinical robotic liquid

handlers.

Active cell manipulation is another critical perspective which could be further explored for

better mimicking in vivo cellular conditions. Chapters 5 and 6 provide an insight on the

capability of the cell stretching platforms to influence cell orientation and achieve patterned

cell culture in 2D. In future, a programmable 3D cell stretching platforms could be developed

to engineer 3D patterned tissues which could be critical and have vast applications in the field

of tissue engineering and regenerative medicine

Furthermore, the use of biocompatible and chemically inert material such as Silicon

carbide in the development of cell stretching platforms could be another direction to

explore. For instance, silicon carbide could be used to develop a cell stretching

platform with micro-structures which could induce precise and controlled mechanical

stimuli to single cells for active manipulation. This could be a difficult task with

common PDMS or hydrogel material. Thus, such materials could be critical for

enhancing the precision and overcoming existing limitations to further improve the

versatility of the cell stretching platforms.

Very few recent approaches have actually explored the advancement in microfluidics

and MEMS to design a complex and efficient cell stretching platforms. A complex

cascaded of microfluidic network with automated handling process will enable

research to deliver different reagents or solutions in precise concentration to better

mimic in vivo biological conditions. Such improved cell stretching platform will

provide a more relevant and reliable model for mechanobiological investigations.

Recent mechanobiological studies with cell stretching platform have shown vast

applications in the field of tissue engineering, drug discovery, regenerative medicines

and biological therapeutics. Thus, considering the current trend the development in the

Page 143: Design and Development of Cell Stretching Platforms for ...

142

cell stretching platforms for mechanobiological studies will grow exponentially in the

future.

Page 144: Design and Development of Cell Stretching Platforms for ...

143

Appendix A: MATLAB Image processing

A.1 Custom made MATLAB program for particle detection and tracing

% clear cache

clear all;

close all;

warning off;

clc;

%% input

img_1=imread('Image Path'); % loading first image

per_frame_time=1.3/2726; % as per camera settings

per_pixel_length=17e-6; % as per camera settings

%% selection area

img_2=imread(strcat('Image Path’,num2str(1002726),'.tif')); % loading last frame

img_3=abs(double(img_1) - double(img_2)); % checking the difference between first and last

frame

img_3=img_3>15; % converting difference into logical threshold

se=strel('disk',1);

img_3=imdilate(img_3,se); % dilation operation

img_3=bwareaopen(img_3,100); % removing small noise changes

img_e=edge(uint8(img_3),'canny'); % edges detection

se=strel('disk',1);

img_e=imdilate(img_e,se); % boarding edges

img_grid(:,:,1)=img_2; % converting image into RGB

img_grid(:,:,2)=img_2;

img_grid(:,:,3)=img_2;

[r,c]=find(img_e==1);

for i=1:length(r)

img_grid(r(i),c(i),:)=[0 255 0];% making edges of detected particles with green colour

end

figure(1)

imshow(img_grid); % display detected particle

title('select the ROI area')

%% Grid generation

hold on

M = size(img_grid,1);

Page 145: Design and Development of Cell Stretching Platforms for ...

144

N = size(img_grid,2);

a=100;

b=100;

for k = 1:a:M

x = [1 N];

y = [k k];

plot(x,y,'Color','black','LineStyle','-');

set(findobj('Tag','MyGrid'),'Visible','on')

end

for k = 1:b:N

x = [k k];

y = [1 M];

plot(x,y,'Color','black','LineStyle','-');

set(findobj('Tag','MyGrid'),'Visible','on')

end

hold off

drawnow % display detected particle with predefined grid

%% user selection

[roi_r,roi_c]=ginput(1); % taking input from user for ROI

roi_r=round(roi_r); %X position

roi_c=round(roi_c); %Y position

[lab,num]=bwlabel(img_3); % checking on which object user clicked

pls=zeros(num,2);

for i=1:num

[r,c]=find(lab==i);

pls(i,:)=[mean(r),mean(c)];

[r1,c1]=find(r==roi_r & c==roi_c);

if(length(r1)>0)

break;

end

end

pls=reshape(pls(pls~=0),size(pls(pls~=0),1)/2,2);

min_roi_r=min(r)-10; % calculating ROI square area

max_roi_r=max(r)+10;

min_roi_c=min(c)-10;

max_roi_c=max(c)+10;

rs=max_roi_r-min_roi_r+1;

cs=max_roi_c-min_roi_c+1;

Page 146: Design and Development of Cell Stretching Platforms for ...

145

%% Frame by frame displacement and strain calculations

k=1;

step=50; % step for images

kern_size=50;

for i=1000001:step:1002726

k;

img_2=imread(strcat('Image sequence path',num2str(i),'.tif'));

[disp(k),xoffset(k),yoffset(k),t1,t2]=calc_disp(img_1,img_2,roi_c,roi_r,kern_size,per_pixel_l

ength); %Calculating the displacement

[dt1y,dt1x]=gradient(double(t1)); %Calculating the gradient in x

[dt2y,dt2x]=gradient(double(t2)); %Calculating the gradient in y

[dyyoffset(k),dyxoffset(k)]=findoff(dt1y,dt2y); %Using the gradient to calculate the strain

in y

[dxyoffset(k),dxxoffset(k)]=findoff(dt1x,dt2x); %Using the gradient to calculate the strain

in x

dyyoffset(k)=dyyoffset(k)*per_pixel_length*per_pixel_length;

dxxoffset(k)=dxxoffset(k)*per_pixel_length*per_pixel_length;

dxyoffset(k)=dxyoffset(k)*per_pixel_length*per_pixel_length;

dyxoffset(k)=dyxoffset(k)*per_pixel_length*per_pixel_length;

strain(k)=sqrt(dxxoffset(k).^2+dyyoffset(k).^2);

mat_displacement(k,:)=[k*step*per_frame_time xoffset(k) yoffset(k) i-1000000];

mat_strain(k,:)=[k*step*per_frame_time dxxoffset(k) dyyoffset(k) i-1000000];

k=k+1;

img_d1(1:size(img_3,1),1:size(img_3,2))=0; % checking image density with pixel wise

img_dx(1:size(img_3,1),1:size(img_3,2))=0; % checking image density with pixel wise in

x

img_dy(1:size(img_3,1),1:size(img_3,2))=0; % checking image density with pixel wise in

y

for j=1:length(r)

d=sqrt( (r-r(j)).^2 + (c-c(j)).^2);

dx=abs(r-r(j));

dy=abs(c-c(j));

d=d<10; % pixel density circular area 10

img_d1(r(j),c(j))=sum(d(:));

img_dx(r(j),c(j))=sum(dx(:));

img_dy(r(j),c(j))=sum(dy(:));

end

e=calculate_strain(img_dx,img_dy);% Eulerian Strain for the image pair

Page 147: Design and Development of Cell Stretching Platforms for ...

146

max_e=reshape(max(max(e)),2,2);

tmp1=reshape(e(:,:,1,1),size(e,1),size(e,2));

tmp2=reshape(e(:,:,2,2),size(e,1),size(e,2));

mat_strain_e(k,:)=[k*step*per_frame_time max(max(tmp1)) max(max(tmp2)) i-

1000000];% logging data for this frame

k=k+1;

figure(6);

for o1=1:2

for o2=1:2

subplot(2,2,(o1-1)*2+o2);

tmp=reshape(e(:,:,o1,o2),size(e,1),size(e,2));

imagesc(tmp);

end

end

end

%% Plots

figure(1)

plot(mat_displacement(:,1),mat_displacement(:,2)) % displacement in X axis

title('Displacement in X (mm) Vs Image Number');

xlabel('Image number');

ylabel('Displacement(mm)');

figure(2)

plot(mat_displacement(:,1),mat_displacement(:,3)) % displacement in Y axis

title('Displacement in Y (mm) Vs Image Number');

xlabel('Image number');

ylabel('Displacement(mm)');

figure(3)

plot(mat_strain(:,1),mat_strain(:,2)) % strain in X axis

title('Strain in X (mm) Vs Image Number');

xlabel('Image number');

ylabel('Displacement(mm)');

figure(4)

plot(mat_strain(:,1),mat_strain(:,3)) % strain in Y axis

title('Strain in Y (mm) Vs Image Number');

xlabel('Image number');

ylabel('Displacement(mm)');

A 2 Custom made MATLAB program for curve detection and ellipse fitting

% clear catch

close all;

Page 148: Design and Development of Cell Stretching Platforms for ...

147

clear all;

clc;

%% Input image

pixel_values = imread('120mbar.tif');

pixelpermm=213; % pixels per mm

figure, imshow(pixel_values)

%% Edge detection

edge_image=edge(pixel_values,'canny'); % Apply canny edge

figure, imshow(edge_image)

%% calculate the mask for edge image. If edge value = 255 else 0.

size(edge_image);

mask=zeros(size(pixel_values)); % mask same as original size

for i=1:1:size(pixel_values,2) % scan raw

for j=1:1:size(pixel_values,1) % scan column

if (edge_image(j,i)==1)

mask(j,i)=1;

break;

end

end

end

% Get the minimum y coordinate (here x is y and y is x in image. See the image_axis.jpg)

imin=1;

for i=1:1:size(pixel_values,1) % scan raw

for j=1:1:size(pixel_values,2)% scan column

if (mask(i,j)==1)

if i>imin

imin=i; jmin=j;

end

end

end

end

cropped_image=mask(1:imin,:); % crop the mask

figure, imshow(cropped_image)

Page 149: Design and Development of Cell Stretching Platforms for ...

148

%% Generate symmetric reflected image along major axis and combined to fit the ellipse

reflected_image=[];

for i=1:1:size(cropped_image,1)

for j=1:1:size(cropped_image,2)

reflected_image(i,j)=cropped_image(size(cropped_image,1)-i+1,j); % Calculate the

reflected image to fit the ellipse

end

end

combined_image=[cropped_image; reflected_image]; % combine the original and the

reflected image to form the full ellipse

figure, imshow(combined_image)

%% Find location of ellipse, convert linear indices to 2d and fit the ellipse.

indexes=find(combined_image); % find the locations where the ellipse is located

[xcoord,ycoord]=ind2sub([size(combined_image,1) size(combined_image,2)],indexes); %

convert linear indices to 2d

elipse_fit=fit_ellipse(xcoord, ycoord); % fit the ellipse from the points

%% plotting fitted ellipse onto the original image for visualisation.

figure, imshow(combined_image) % display the original image

hold on

phi = linspace(pi,2*pi,213); % line 43 - 60 will plot the fitted ellipse on the original image

cosphi = cos(phi);

sinphi = sin(phi);

xbar = elipse_fit.Y0;

ybar = elipse_fit.X0;

theta = pi*elipse_fit.phi/180;

R = [ cos(theta) sin(theta) -sin(theta) cos(theta)];

xy = [elipse_fit.b*cosphi; elipse_fit.a*sinphi];

xy = R*xy;

x1 = xy(1,:) + xbar;

y1 = xy(2,:) + ybar;

y2=repmat(max(y1),1,300);

x2=linspace(min(x1),max(x1),300);

x=[x1 x2]; y=[y1 y2];

plot(x,y,'r','LineWidth',2); % plot the ellipse points in the image

%% ideal graph 2D

y3=max(y1)-y1; % the image coordinate starts from top left. We need the plot from bottom so

subtract it from max.

Page 150: Design and Development of Cell Stretching Platforms for ...

149

indx=1:1:size(y3,2);

figure, plot(indx./ pixelpermm,y3./ pixelpermm)

xlabel ('X-cordiante (mm)');

ylabel ('Displacment of membrnae mm)');

%% fitted plot 2D

y1=y1((x1>0)); % find the positive coordinates

y1=max(y1)-y1; % the image coordinate starts from top left. We need the plot from bottom so

subtract it from max.

indx=1:1:size(y1,2);

figure, plot(indx./ pixelpermm,y1./ pixelpermm)

xlabel ('X-cordiante (mm)');

ylabel ('Displacment of membrnae(mm)');

x1=x1(x1>0);

major_axis=max(x1)-min(x1); % subtract max and min to find the length of major axis

major_axis=major_axis/ pixelpermm

minor_axis=max(y1);

minor_axis=minor_axis/pixelpermm ;% find the max to get the maximum value (length of

minor axis)

A 3 Custom made MATLAB program for ElveFlow pressure controller

% clear catch

close all;

clear all;

clc;

%% Initialisation

Elveflow_Load; %% load the DLL

error =0;% int error to zero, if an error occurs in the dll, an error is returned

answer='empty_sring';% store the user answer in this variable

%the instrument name can be found in NI Max

Instrument_Name = libpointer('cstring','OB1'); % ‘OB1’is the name of my instrument

%create a pointer for calibration

CalibSize = 1000;

Calibration = libpointer('doublePtr',zeros(CalibSize,1));

%pointer to store the instrument ID (no array)

Inst_ID=libpointer('int32Ptr',zeros(1,1));

Page 151: Design and Development of Cell Stretching Platforms for ...

150

%Initiate the device and all regulators and sensors types

if error==0

error=OB1_Initialization(Instrument_Name,2,2,4,4,Inst_ID);

end

if error==0

error=OB1_Add_Sens(Inst_ID.Value,3,1,1,0); %add digital flow sensor ;

end;

disp(strcat('Instrument ID = ', num2str(Inst_ID.Value)));%Display the instrument number

%% CALIBRATION

%Chose between new calibration mode: load calibration or default calibration

Calib_Save = libpointer('cstring',' ');

Calib_Load = libpointer('cstring',' ');

% ask user what kind of calibration to use

while (~(strcmp(answer,'new')||strcmp(answer,'load')||strcmp(answer,'default')))

prompt = 'what kind of calibration do you want to use ?(new, load, default)\n';

answer = input(prompt,'s');

end

if strcmp(answer,'new')% New calibration

if error==0

error=OB1_Calib(Inst_ID.Value,Calibration, CalibSize);

end;

%Save the calibration for further use

if error==0

error=Elveflow_Calibration_Save( Calib_Save , Calibration, CalibSize);

end

end

if strcmp(answer,'load')%load previous calibration

if error==0

error = Elveflow_Calibration_Load( Calib_Load , Calibration, CalibSize);

end

end

if strcmp(answer,'default')%use default calibration

if error==0

error = Elveflow_Calibration_Default(Calibration, CalibSize);

end

end

% set all channels to 0

open=0;

if error==0

Page 152: Design and Development of Cell Stretching Platforms for ...

151

OB1_Set_Press(Inst_ID.Value,1,open,Calibration,CalibSize);

OB1_Set_Press(Inst_ID.Value,2,open,Calibration,CalibSize);

OB1_Set_Press(Inst_ID.Value,3,open,Calibration,CalibSize);

OB1_Set_Press(Inst_ID.Value,4,open,Calibration,CalibSize);

end

% Initial variables

Press_Array = libpointer('doublePtr',zeros(4,1)); % to store 4 channel pressure

flow_rate = libpointer ('doublePtr',zeros(1,1)); % pointer to store the flow rate

trigger = libpointer ('int32Ptr',zeros(1,1)); % pointer to trigger

channel_n = -1;

trigger_value=0;

%% Wave input

Amplitude=100; %change the pressure wave amplitude as desired in micro litre

Frequency=10; %change the frequency as desired

Cycles = 5;

Wave_type=2; %use 1 for sin, 2 for square, 3 for sawtooth

Time_total=1/Frequency;% Simulation for 1 cycle.

Time_step=Time_total/100;%initialising the time step.

End_Time= Time_total*Cycles;

Time=0:Time_step:End_Time;

Pressure=Input_function(Amplitude,Frequency,Wave_type,Time);

%% slect desired channel

ChannelToSetPressure=3;% You can change this value to 1,2,3,4 as desired to set the

pressure.

ChannelToGetFlow=1;% You can change this value to 1,2,3,4 as desired to get the flow.

flow=[];

presCh1=[];

presCh2=[];

presCh3=[];

presCh4=[];

Inst_ID.Value

%% Main loop

for pres=Pressure

if Wave_type==2

if pres == Amplitude

OB1_Set_Press(Inst_ID.Value,ChannelToSetPressure,pres,Calibration,CalibSize);

else

OB1_Set_Press(Inst_ID.Value,ChannelToSetPressure,0,Calibration,CalibSize);

end

else

OB1_Set_Press(Inst_ID.Value,ChannelToSetPressure,pres,Calibration,CalibSize);

Page 153: Design and Development of Cell Stretching Platforms for ...

152

end

OB1_Get_Press(Inst_ID.Value,10,Calibration, Press_Array, CalibSize, 4);

temp=Press_Array.Value;

presCh1=[presCh1,temp(1)];

presCh2=[presCh2,temp(2)];

presCh3=[presCh3,temp(3)];

presCh4=[presCh4,temp(4)];

end

figure(2)

plot(Time,presCh1)

title(strcat('Pressure ','for channel 1'));

xlabel('Time');

ylabel('Pressure');

figure(3)

plot(Time,presCh2)

title(strcat('Pressure ','for channel 2'));

xlabel('Time');

ylabel('Pressure');

figure(4)

plot(Time,presCh3)

title(strcat('Pressure ','for channel 3'));

xlabel('Time');

ylabel('Pressure');

figure(5)

plot(Time,presCh4)

title(strcat('Pressure ','for channel 4'));

xlabel('Time');

ylabel('Pressure');

%% EXIT routine

close=0;

if error==0

OB1_Set_Press(Inst_ID.Value,ChannelToSetPressure,close,Calibration,CalibSize);

OB1_Destructor(Inst_ID.Value);%close communication with the instrument

end

Elveflow_Unload;% Unload DLL

clear Instrument_Name;

clear Inst_ID;

clear MyCalibPath;

Page 154: Design and Development of Cell Stretching Platforms for ...

153

clear Calibration;

clear Press_Array;

clear flow;

%clear trigger;

clear presCh1;

clear presCh2;

clear presCh3;

clear presCh4;