microfluidic approach for the scalable formation of 3d soft ......we have developed a...

146
Microfluidic Approach for the Scalable Formation of 3D Soft Materials with Tailored Biomechanical Properties for Tissue Engineering Applications by Lian Leng A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Mechanical and Industrial Engineering University of Toronto © Copyright by Lian Leng 2015

Upload: others

Post on 26-Feb-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

Microfluidic Approach for the Scalable Formation of 3D Soft Materials with Tailored Biomechanical Properties for

Tissue Engineering Applications

by

Lian Leng

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Department of Mechanical and Industrial Engineering University of Toronto

© Copyright by Lian Leng 2015

Page 2: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

ii

Microfluidic Approach for the Scalable Formation of 3D Soft

Materials with Tailored Biomechanical Properties for Tissue

Engineering Applications

Lian Leng

Doctor of Philosophy

Mechanical and Industrial Engineering

University of Toronto

2015

Abstract

Through evolution, nature has defined materials in a highly organized and hierarchical fashion,

with a complex organization on the nano and microscale that is closely linked to the different

tissues and organs function. A variety of macroscale parameters such as mechanical strength,

elasticity, and permeability to various molecules are tissue-specific and are a direct result of their

intricate structural organization. The field of tissue engineering has seen the development of a

broad range of technologies with the common goal of enabling the generation of complex tissues

that mimic the intricate architecture and resulting properties of natural tissues. The methods

developed so far are promising but are faced with short-comings in terms of scalability,

throughput, and dynamic control over the local material composition and properties. These

strategies are often limited in size (<1cm), rely on sequential assembly processes, and the

resulting mechanical properties of these engineered tissues are often not comparable to native

tissues.

We have developed a microfluidic-based printing platform for the continuous and scalable

formation of planar soft materials with high degree of control over the material composition,

Page 3: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

iii

cellular composition, nanoscale to macroscale assembly, and mechanical properties. The spatial

control over the different compositions can be precisely defined, with a resolution of 75µm. The

incorporation of a variety of payloads such as biomolecules, fluorescent particles and viable cells

including cardiomyocytes, keratinocytes, and fibroblasts is demonstrated. The platform was used

with a variety of matrix materials including alginate and collagen type I from rat tail. The

mechanical properties of the printed soft material sheets produced could be tailored by either a

planar assembly of various materials possessing distinct mechanical properties, or by controlling

the degree of alignment of collagen fibers. Demonstrated applications include the formation of

aligned collagen sheets with mechanical properties comparable to native tissues, and up to 6cm

wide cell-populated skin substitute with physiological properties similar to the native skin. The

approach presented promises wide applications in the field of regenerative medicine where the

ability to tailor the material and cellular composition, as well as the mechanical properties of the

replacement tissue have become a key requirement.

Page 4: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

iv

Acknowledgments

This dissertation is dedicated to my family. To my parents, for being supportive and encouraging

through ups and downs. To my sister, for always being there for me and for being such a good

listener. To my family for teaching me the value of hard work and sharing their passion for what

they do.

To my dissertation committee, for their guidance and valuable inputs. To Dr. Axel Guenther, for

teaching me how to be a thorough and meticulous experimentalist, to be inquisitive and take any

failure with excitement for it can only mean new discoveries, and for sharing his passion for

science.

To our collaborators from Dr. Elliot Chaikof’s laboratory. Particularly Dr. Stephanie Grainger

for all the mechanical testing, and TEM-SEM imaging of the collagen samples. To our

collaborators from Dr. Marc Jeschke’s laboratory at the Sunnybrook Research Institute,

particularly Shermineh Minai for being such a great help in providing us with large cell numbers,

and Cassandra Belo for the long hours spent on in vitro and in vivo sample staining and imaging.

I would also like to give my appreciation to Dr. Saeid Amini-Nik and Cassandra Belo for

performing the surgeries for our in vivo experiments. To Dr. Milica Radisic and Boyang Zhang

for our collaborative work on fibroblasts and cardiomyocyte printing. I would like to extend my

thanks to Boyang Zhang for our extensive discussions and for being such a pleasure to work

with. To Prof. Craig Simmons for access to the tensile testing apparatus in his laboratory.

To all members of the Guenther lab, particularly Sanjesh Yasotharan, Arianna McAllister, and

Geoff Vishloff for insightful discussions and help.

Page 5: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

v

Table of Contents

Acknowledgments .......................................................................................................................... iv

Table of Contents ............................................................................................................................ v

List of Figures ............................................................................................................................... vii

List of Tables .................................................................................................................................. x

Contributions .................................................................................................................................. xi

1 Chapter 1 - Introduction ............................................................................................................. 1

1.1 Strategies for the Formation of Organized Materials ......................................................... 2

1.1.1 Photo and Laser Polymerization ............................................................................. 2

1.1.2 Molding and Templating ......................................................................................... 3

1.1.3 Self-Assembly of Living Cells ................................................................................ 5

1.1.4 Bioprinting .............................................................................................................. 6

2 Chapter 2 - Formation of Aligned Soft Material Sheets .......................................................... 10

2.1 Fibrillar Structure of Materials in Nature ......................................................................... 10

2.2 Collagen Organization in Tissues ..................................................................................... 11

2.3 Strategies for Controlling Collagen Self-Assembly In Vitro ............................................ 13

2.4 Strategy for the Continuous Formation of Collagen Sheets with Aligned

Nanostructure .................................................................................................................... 16

2.4.1 Materials and Methods .......................................................................................... 17

2.4.2 Results ................................................................................................................... 21

3 Chapter 3 - Continuous Formation of Patterned Soft Material Sheets Populated with

Molecular and Cellular Payloads ............................................................................................. 34

3.1 Alginate in Tissue Engineering ......................................................................................... 34

3.2 Strategy for the Continuous Formation of Soft Material Sheets of Heterogeneous

Composition ...................................................................................................................... 36

3.2.1 Material and Methods ........................................................................................... 36

3.2.2 Mosaic Hydrogel Formation ................................................................................. 41

Page 6: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

vi

3.2.3 Spatiotemporal Control and Payload Incorporation ............................................. 43

3.2.4 Information Encoding ........................................................................................... 45

3.2.5 Geometric Control over Mosaic Hydrogel Properties .......................................... 46

3.2.6 Scalable Formation of 3D Hydrogel Assemblies .................................................. 47

3.2.7 Planar Co-localization of Single and Multiple Cell Types ................................... 47

4 Chapter 4 - Skin Printer: Continuous and Scalable Organization of Layered Tissue .............. 58

4.1 Skin Structure .................................................................................................................... 58

4.2 Skin grafts and Artificial Skin Substitutes ........................................................................ 59

4.2.1 Skin Grafts ............................................................................................................ 60

4.2.2 Artificial Skin Substitutes ..................................................................................... 61

4.2.3 Microfluidic approaches ....................................................................................... 62

4.3 3D Printer Strategy for the Continuous Formation of Skin Substitute for Wound

Healing .............................................................................................................................. 63

4.3.1 Materials and Methods .......................................................................................... 63

4.3.2 Formation of Cell-Populated Skin Grafts ............................................................. 67

4.3.3 Pattern Formation – No Cells ............................................................................... 68

4.3.4 Characterization of Printed Tissue Substitutes ..................................................... 68

4.3.5 Cell-Populated Skin Grafts in Vitro ...................................................................... 71

4.3.6 Scalable Formation of Tissue Substitutes ............................................................. 72

4.3.7 Cell-Populated Skin Grafts In Vivo ....................................................................... 72

Summary and Future Work ........................................................................................................... 81

References ..................................................................................................................................... 86

Appendix ..................................................................................................................................... 107

Page 7: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

vii

List of Figures

Figure 1. Schematic illustration of the flowable conversation of a collagne solution to a

cross-linked collagen sheet with a high degree of molecular alignment…………...

27

Figure 2. Schematic illustrations of experimental setup………………………...………….... 28

Figure 3. Detailed study of flow-focusing and strain-inducing pulling using a vertical

constriction unit……………………………………………………………………..

29

Figure 4. Measured collagen sheet width and thickness as a function of V* ranging from 0.1

to 10…………………………………………………………………………………

30

Figure 5. Characterization of collagen fibril compaction, alignment, and resulting

mechanical properties…………………………………………………………….....

32

Figure 6. One-step formation of mosaic hydrogels……………………………...………….... 50

Figure 7. Analytical model of the time-dependent concentration of free cross-linker, free

alginate, and cross-linked gel...……………………………………………………..

51

Figure 8. Continuous formation of hydrogel sheets: experimental setup and device designs..

52

Figure 9. Dynamically encoded information in planar hydrogels…………………………....

54

Figure 10. Mosaic hydrogels…………………………..………………………...………….... 56

Figure 11. Intact human skin and bioprinted skin grafts..…………………………………….

74

Figure 12. Skin printer…………………..……………………………………………………

75

Figure 13. Patterned single-layered sheets…………………………………………………....

76

Figure 14. In vitro characterization of printed skin grafts.……………………...………….... 77

Figure 15. Material optimization for keratinocyte printing…..……………………………….

78

Figure 16. Scalable formation of skin grafts………………………………….....……………

79

Figure 17. In vivo characterization of printed skin grafts…………………….....……………

80

Figure 18. Benefit of cell clustering and characterization of skin microtissues...……………

84

Figure 19. In vivo characterization of bilayered cell-populated skin grafts....………………..

85

Page 8: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

viii

Appendix

Figure A1. Rendered device designs…………………..…………………….....……………

108

Figure A2. Characterization of velocity profiles within the constriction………...…………

110

Figure A3. Numerical simulation of the flow behavior of the focusing solution within

½ of the constriction (½ HC)...…………………………………………………..

112

Figure A4. Characterization of collagen sheet width and thickness produced at varying

V* and QF, with constant QM..…………………………………………………...

112

Figure A5. Mechanical properties of collagen sheets subjected to FIB incubation and

drying post-extrusion……………...……………………………………………..

113

Figure A6. TEM images of collagen sheets produced at various V*……………...…………

113

Figure A7. Characterization of fibril spacing and compaction by autocorrelation of TEM

and SEM images………...……………………………………………………….

114

Figure A8. Photograph of microfluidic device with constriction unit…..………...…………

115

Figure A9. Constriction manifold assembly………………………...…..………...…………

116

Figure A10. Constriction manifold – Top piece……………………...…..………...……..…

116

Figure A11. Constriction manifold – Bottom piece……………………...…..………...……

117

Figure A12. XZ constriction manifold assembly….……………………...…..………...……

117

Figure A13. XZ constriction manifold – Frame...….……………………...…..………...…...

118

Figure A14. XZ constriction manifold – Left constriction bracket..……...…..………...…...

118

Figure A15. XZ constriction manifold – Right constriction bracket……...…..………...…...

119

Figure A16. Control over soft material thickness as a function of QF………………………

121

Figure A17. Characterization of pressure in on-chip reservoirs……..………………………

122

Figure A18. Shear stress profile within a microfluidic channel………..……………………

123

Figure A19. Viability and distribution of printed cells……………...………………………

124

Figure A20. Modulus of elasticity for a homogeneous soft material composed

of 2%w.t. alginate produced in free and pulled extrusion modes…..…………

125

Page 9: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

ix

Figure A21. Line camera intensity measurements of the UN Charter………………………

126

Figure A22. Full-thickness burns and current treatments…………………….…..…………

128

Figure A23. Characterization of spot volume as a function of valve actuation….…………

128

Figure A24. Degradation studies of skin grafts……………………………….…..…………

129

Figure A25. Mechanical properties of skin grafts as a function of culture time...………….

130

Figure A26. Stress-strain curves of skin grafts materials as a function of culture time…….

131

Figure A27. Stress-Strain curves of skin graft printed with parallel spots………………….

132

Figure A28. Stress-Strain curves of skin graft printed with parallel stripes………………...

133

Figure A29. Stress-Strain curves of skin graft printed with alternating voids..……………..

134

Figure A30. Stress-Strain curves of skin graft printed with alternating spots……………….

135

Page 10: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

x

List of Tables

Table A1. Natural tissues and their mechanical properties.………………….....………….

109

Table A2. Mechanical and structural properties of natural and synthetic collagen gels

created using a variety of strategies……………….....…………………………..

110

Table A3. Parameters used in numerical model, treating the focusing solution as a

Newtonian fluid…………………...……………….....…………………………..

111

Table A4. Mechanical properties of alginate gels………….………………….....………….

120

Table A5. Literature data of elastic moduli of human skin measured in vivo using

various strategies..………………...……………….....…………………………..

127

Page 11: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

xi

Contributions

Publications

Leng, L., McAllister, A., Zhang, B., Radisic, M., and Guenther, A., "Mosaic Hydrogels:

One-Step Formation of Multiscale Soft Materials", Advanced Materials, 24, July 2012,

3650-3658

Singh, R., Genov, R., Leng, L., Guenther, A., "A Hybrid CMOS-Microfluidic

Luminescence Contact Imaging Microsystem," accepted at 2009 SPIE Optics+Photonics

(Invited), Aug 2009. SPIE Paper Number 7397-39

Manuscript in progress

Leng, L., Grainger, S., Chaikof, E., Guenther, A., “In-Flow Preparation of Collagen

Sheets with Tunable Molecular Alignment and their Influence on Cell Behavior”.

Leng, L., Ba, Q., Amini-Nik, S., Jeschke, M., Guenther, A., “Skin Printer: Continuous

and Scalable Organization of Layered Tissue”.

Patent

Leng, L., Zhang, B., McAllister, A., Wollard, A., Radisic, M., Guenther, A. "Devices and

methods for producing planar polymeric materials using microfluidics, WO2013075248A1.

Others

Chapter 2: TEM/SEM images and tensile data were obtained by Dr. Stephanie Grainger.

Chapter 4: In vitro and in vivo sample staining and imaging was performed by Cassandra Belo.

Animal surgeries and histology data were performed and obtained by Dr. Saeid Amini-Nik and

Cassandra Belo.

Page 12: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

1

1 Chapter 1 - Introduction

Natural selection and evolution through time has resulted in the formation of a wide variety of

materials in living organisms, shaping them to adapt to environmental changes and optimize their

performance. From microbes to plants and mammals, the materials shape and microstructure are

closely related, and are formed in a simultaneous process, that is through growth of the organ 1.

As these materials form at multiple length scales, from nano to macroscale, the resulting

structure of these biological materials becomes highly hierarchical 2. Through evolution, this

complex organization has been optimized to efficiently serve tissue-specific functions and define

the characteristic properties of the materials.

In this thesis, we present a novel platform for the continuous and high throughput formation of

complex soft materials with tunable composition and structure that can be adapted to specific tissues.

In the first chapter, various strategies for the assembly of organized materials are reviewed. In the

second chapter, we explore our ability to define directionally dependent microstructures with the

formation of collagen sheets of highly anisotropic fibril alignment. The anisotropic microstructures

and resulting mechanical properties are investigated. In the third chapter, we introduce an adapted

platform for the formation of materials with heterogeneous composition, possible through the

addition of on-chip wells that can be loaded with a variety of payloads ranging from biomaterials and

fluorescence particles to living cells. The resulting mechanical properties such as elastic modulus and

diffusivity, as well as the viability and attachment of living cells were investigated. In the fourth

chapter, we focus on the application of our platform to the formation of skin grafts composed of

islands of microtissues mimicking the epidermal and dermal layer of skin. In vitro and in vivo

characterization of the grafts mechanical properties, cell proliferation, and wound healing on murine

models were performed.

Page 13: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

2

1.1 Strategies for the Formation of Organized Materials

Over the last few decades, research groups have developed a broad range of strategies to

generate synthetic organized materials in an attempt to recreate the multicellular composition and

structural organization of natural materials. The ultimate goal of these technology developments

is to facilitate the directed assembly of biologically relevant materials, with prescribed three-

dimensional hierarchical organization and physiologically relevant composition. Some of these

approaches include, but are not limited to the formation of microscale building blocks that are

either randomly or orderly assembled into higher-order structures. These building blocks may be

in the form of spherical or spheroidal bubbles and droplets, or polymer particles and disks 3. The

formation of building blocks can be achieved through a variety of approaches such as

photopolymerization 4-17, molding 18-31, self-assembly of living cells 32-40, and bioprinting 26,27,41-

70.

1.1.1 Photo and Laser Polymerization

Photoinitiated polymerization techniques have been adapted to generate such building blocks

with the ability to control the geometry and sizes of these particles within the xy-plane 10-13 or

even in three dimensions 14. These particles can be subsequently assembled either manually 13,

through surface tension at a liquid-liquid interface 11, through shape-affinity and fluid flow 15,16,

or microfluidic templating 17. Similarly, these planar assembly approaches can be scaled-up to

three-dimensional assemblies of droplets 4, polymer particles 5, and microbubbles 6. Cheung et

al. used confocal scanning as well as conventional fluorescence microscopy to polymerize three-

dimensional polymer particles with various geometries, sizes up to 1mm2, and spatial resolution

down to 3µm 7. Select regions could be sequentially photopolymerized, with each step requiring

a wash before flowing the following photocurable reagent. A variety of polymers were tested,

including PEG-DA and other acrylate-based polymers.

Larger-scale photopolymerized scaffolds of heterogeneous cellular and material compositions

were also created by through a bottom-up assembly using stereolithography 8. A reservoir filled

with PEG-DA and RGDS peptide solution containing various cell types was photopolymerized

using a laser beam. Multiple layers were built up in a sequential process where a second layer of

Page 14: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

3

uncrosslinked cell-seeded polymer solution was coated onto the cross-linked layer, exposed, and

the process repeated for subsequent layers. Complex structures 2mm in thickness were produced

with a total of 20 layers. This approach remains sequential, results in wasteful discarding of

excess uncrosslinked cell-seeded polymer, and is limited in thickness due to the lack of a

vascular network. Cell death was noticeable in the centre of the scaffold, with an increase in cell

number at the periphery of the scaffolds after 14 days of culture.

Laser microablation techniques have also been applied to create three-dimensional scaffolds with

geometrically defined micropores 9. Multilayered poly (glycerol-sebacate) (PGS) scaffolds with

controlled pore microarchitectures were fabricated and assembled through a layer-by-layer

oxygen plasma treatment. The final assembly required storage under compression for 18hrs to

fully bond the different layers, resulting in a final construct 6mm in diameter and 500µm thick.

The scaffolds were subsequently immersed in a suspension of cardiomyocytes for seeding and

showed contractile behavior of the entire construct.

1.1.2 Molding and Templating

Hierarchical biological structures can also be produced through sequential templating, molding,

and layer-by-layer assembly approaches.

Template-assisted self-assembly of spherical colloids and hydrogel beads was achieved by

capillary forces and geometric confinement 18,19. Matsunaga et al. produced three-dimensional

tissue architectures by packing cell-populated beads into silicon molds 19. The collagen type I

beads were produced at a rate of 104 beads/min using a flow-focusing device and had diameters

ranging from 100-300µm. Cells were seeded and allowed to attach and penetrate the collagen

beads. The resulting cell-populated beads were stacked into polydimethylsiloxane (PDMS)

molds and observed to attach to one another and contract as a bulk structure from an initial

1.5mm thickness down to 1.2mm. However, the lack of a vascularized network within such thick

substrate presents a challenge in diffuse transport of nutrients throughout the scaffold, a

phenomenon also observed in the formation of large alginate molds for cartilage regeneration 20.

Cell-seeded alginate implants were injection-molded to anatomical shapes, with average molded

construct size of 25.3 × 17.4 × 6.6mm 20. Although articular chondrocytes in cartilage tissue

Page 15: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

4

engineering require less vascularization and do not undergo necrosis in thick (>0.3mm)

scaffolds, it is commonly observed that low oxygen concentration in the chore of thick hydrogel

implants result in an enhanced extracellular matrix (ECM) formation and cell necrosis 21.

Molding and layering approaches have been used separately or in combination to generate cell-

populated multilayered biopolymers constructs. Nishiguchi et al. employed a layer-by-layer

assembly technique to create tissue models such as blood vessels, skeletal muscle, and

connective tissue 22. Less than 10nm thick fibronectin-gelatin films were prepared and deposited

on cell surfaces, promoting cell-cell adhesion between layers. Their approach resulted in the

formation of 1cm2 and 50µm thick cell-populated substrates. Although promising due to the

control over the cell type and location in the z-direction, this sequential approach is lengthy and

therefore limited to submillimeter-thick constructs. In another approach, layer-by-layer assembly

processes were combined with microfluidic geometries to create three-dimensional layered

biopolymer matrices containing cell suspensions at a concentration of 3 × 105 cells/mL 23.

Matrices consisting of collagen, collagen-chitosan, matrigel, and fibrin were sequentially flown

through microfluidic channels 350µm [W] × 300µm [δ] and cross-linked by incubation at 37 ̊C

for each layer introduced, creating multilayered composite biopolymer constructs. The thickness

was however limited by soft-lithography fabrication of the microfluidic channels.

Biomaterials layers were also molded with perfusable channels to mimic in a simplified fashion

the vascular network of natural tissues, thereby promoting the transport and diffusion of nutrients

throughout the entire scaffold 24-30. Choi et al. molded a solution of primary chondrocytes

suspended in 4%w.t. calcium alginate into an aluminum jig 24. A silicon master patterned with a

connected array of 100µm wide microchannels was placed within the jig and serve to create a

perfusable network of microchannels, ensuring cell viability throughout the 2mm thick scaffold.

Perfusable networks were also produced using sacrificial molds 30,31. A few examples of

sacrificial templating include the incorporation of phosphate glass fibers within a 50µm thin

planar collagen layer 31. The structure was subsequently built-up by rolling onto itself and the

fibers degraded, producing a perfusable tubular structure 2mm in diameter. Yoshida et al. used a

similar templating approach to generate biodegradable multi-layered construct mimicking the

bilayered structure of blood capillaries . A parallel array of microchannels within a

biodegradable hydrogel was created by the molding and later removing silica tubes 720µm in

Page 16: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

5

diameter. The hollow microchannels created within the disulfide-crosslinked γ-PGA gels were

then sequentially coated with smooth muscle cells followed by endothelial cells. Through this

approach, they were able to construct multilayered blood capillaries possessing barrier functions

similar to the native blood capillaries. Permeability assay was performed by flowing FITC-

Dextran molecules through the tubular network and comparing the diffusion over 24hrs between

the cell-coated and uncoated constructs. They also attempted to collect these cell-populated

capillaries by decomposing the casting hydrogel by the addition of cysteine, a biocompatible

reductant. The collected capillaries were 1cm in length and maintained their hollow structure,

promising potential use as implantable artificial blood capillary.

1.1.3 Self-Assembly of Living Cells

Another strategy for the assembly of organized three-dimensional tissues relies on the template-

free self-assembly of living cells into sheets 32-37. Yang et al. successfully harvested intact cell

sheets from thermally responsive dish, preserving cell-cell junction protein and ECM that would

be degraded by conventional trypsinization for cell recovery 36. The temperature-responsive

culture dishes were created by covalent grafting of a temperature-responsive polymer poly(N-

isopropylacrylamide) (PIPAAm) to ordinary tissue culture dishes. Under normal culture

conditions at 37ºC, the dish surfaces are hydrophobic and cells are able to attach, spread, and

proliferate. Upon temperature reduction below the polymer’s lower critical solution temperature

of 32ºC, the polymer surface becomes hydrophilic and swells, forming a hydration layer between

the dish surface and the cultured cells. This layer allows the spontaneous detachment of cell

sheets without the need for enzymatic treatments such as trypsinization. By avoiding proteolytic

treatment, critical cell surface proteins such as ion channels, growth factor receptors and cell-to-

cell junction proteins remain intact, and cells can be noninvasively harvested as intact sheets

along with their deposited ECM. These sheets are subsequently able to spontaneously attach to

the region of implant due to ECM present. Relying on the cell sheets spontaneous attachment

capability, up to five cell sheets were layered to engineer transplantable corneal epithelial

sheets 38 as well as cardiac sheets 39. This technique was later on adapted to produce cell sheets

for co-culture of different cell types 35. In this approach, two thermoresponsive polymers

exhibiting different transition temperatures were sequentially patterned onto culture dishes,

Page 17: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

6

namely hydrophilic poly(N-isopropylacrylamide) (PIPAAm) and hydrophobic monomer n-butyl

methacrylate (BMA). Sequential seeding of viable cells enabled the patterned co-culture of rat

primary hepatocytes and bovine carotid endothelial cells. Chen et al. produced in a similar

approach cell sheets by seeding onto thermoreversible methylcellulose hydrogels 32. Human

foreskin fibroblasts were seeded at a density of 4 × 104 cells/cm2 and their attachment was

improved by evenly coating the hydrogel with a layer of neutral aqueous collagen. Cell sheets

were completely detached within 20min, and multilayered sheets were produced by directly

seeding a second layer of cells onto a non-detached cell sheet.

However, due to the reduced diffusion of oxygen and nutrient through the non-vascularized

construct, the maximum number of layers is limited to five layers (~200-300µm total thickness).

In addition, cell sheet strategies require large cell numbers to produce centimeter scale sheets.

Currently, seeding of 2×104 cells/cm2 produces 5.98 ± 0.12cm2 sheets before detachment from

the thermally-reversible dish. Once detached, the cell sheet area is reduced to 1.16 ± 0.08cm2

with a thickness of 45 ± 8µm due to cytoskeletal tensile reorganization 39. The steps involved in

the preparation of cell sheets are also timely as they required approximately 2-3 weeks before the

sheets can be manipulated, about 1h for sheet detachment, and about 30min for sheet

reattachment to create multilayer constructs.

More recently, natural ECM sheets were decellularized and subsequently repopulated with viable

cells to generate scaffold sheets possessing natural structure of tissues. Kim et al. extracted

ECM from human adipose tissue and explored the potential application of this natural substrate

as a scaffold for other types of tissues 40. Decellularized ECM sheets (26mm [L] × 10 mm [W] ×

0.15mm [δ]) were seeded with a variety of human cells including dermal fibroblasts, aortic

smooth muscle cells, chondrocytes, umbilical vein endothelial cells, and adipose-derived stem

cells, and the respective cells attachment and proliferation were assessed.

1.1.4 Bioprinting

Ink-jet printing technologies has gained increasing attention over the last decade for applications

in regenerative medicine 26,27,41-51, with the first international workshop on bioprinting and

biopatterning taking place at the University of Manchester in September 2004. Bioprinting

Page 18: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

7

technologies have emerged from the adaptation of rapid prototyping strategies used to construct

structures made generally of hard plastic for rapid casting 52, ceramics for dental prostheses 53,54,

or the low-cost production of laboratory components 55 and microwells 56 that would have

normally required dedicated infrastructures. These have been adapted to manufacture, through a

layer-by-layer bottom-up approach, scaffolds onto which cells could attach and grow until they

were mature enough for implantation into the body. However through further development, a

more appealing approach is to simultaneously print living cells and scaffolding materials. Early

brioprinters were adapted from existing desktop printers 45,57, and used to create either droplets

of biomaterials and cell suspensions that are dispensed by heating and vaporizing the bioink 58,

mechanical actuation using a piezoelectric element 51,59, or microextrusion of a filament 60.

These bioprinters can be used to create heterogeneous architectures either by controlling the

micropositioning of specific cell types within a plane 43,44,57,61 , or to generate a biological

template that recreates the cellular environment 51,60,62. Direct cell printing can be achieved by

printing cell-adhesion proteins and monoclonal antibodies onto a substrate material and allowing

the cells to specifically attach to patterned regions 57,59,61,64. Roth et al. adapted ink-jet printing

technology to pattern collagen films with a resolution of 350µm onto glass slides dip-coated in

agarose 44. Regular printer cartridges were rinsed thoroughly with ethanol and sterile water prior

to use. Smooth muscles cells and neural cells were seeded onto the protein-patterned glass slides

and their culture monitored over time, showing preferential attachment to the protein patterns.

Boland et al. extended the cell patterning approach from seeding onto a planar surface to

embedding within a gel layer, creating a three-dimensional microenvironment for the cells. Their

approach relied on the sequential deposition of collagen gel and cell suspension 41. Specifically, a

layer of collagen type I was first coated onto a hosting substrate, with thickness ranging from

200-500µm. Bovine aortal endothelial cells were suspended in culture media and subsequently

printed onto the collagen layer, followed by coating with a final layer of collagen. This process

relies on the cells to migrate into the collagen gel, form clusters and aggregate within the

construct. They found the collagen layers to be essential in promoting fusion of adjacent cell

aggregates. They attempted to reproduce this result by replacing the collagen layers with a

thermo-reversible gel and showed successful cell fusion, although not as effective as the collagen

case. One of the limitations of direct cell printing lies in the need for the cell suspensions to

remain stable and homogeneously distributed over several minutes (beyond 20minutes). Cell

Page 19: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

8

agglomeration or sedimentation affecting the printing performance is a common occurrence with

this approach 59. In addition, studies have shown that the dispensing pressure through a

microscale nozzle during printing has more significant effect on the cell viability than the nozzle

diameter, with 38.75% reduction in cell viability when printed at 40 psi as opposed to 5 psi 63.

Complex hierarchical structures mimicking anatomical features of the body were first created by

Vacanti et al. through a molding-casting strategy utilizing alginate gels, creating various shapes

such as human ear 65, trachea 66, nasal tip 67, and nasal septum 68. Cohen et al. utilized printing

strategies to form alginate-based geometries with a surface roughness of 160 ± 20µm in the z-

direction for cartilage tissue regeneration 62. Alginate solutions at a concentration of 2%w.t.

containing chondrocytes at a concentration of 33 × 106 cells/mL were ejected through a 0.84mm

diameter tip at a rate of 0.6mL/min. Gelation of the cell-suspended alginate solution was initiated

by mixing with 10mg/mL of CaSO4 as a crosslinker prior to introducing into the printer. Due to

the time-dependent gelation, this lead to a limited time frame for optimal printing of only 15min.

More recently, Hockaday et al. successfully printed heterogeneous aortic valve scaffolds a

combination of 3D-printing and photocrosslinking strategy 60. Native anatomic and axisymmetric

aortic valve geometries were printed with poly-ethylene glycol-diacrylate (PEG-DA)

supplemented with 10-15%w.t. alginate, which was found to increase sufficiently the bio-ink

viscosity for efficient extrusion. A range of PEG-DA with alginate blends were considered and

showed a wide range in elastic modulus from 5.3 ± 0.9 to 74.6 ± 1.5kPa. Heterogeneous material

printing was achieved by interchanging the bio-ink solutions, enabling a stiffer material to be

printed for the aortic root wall, and a more compliant and extensible material for the leaflets. The

print time ranged from 14 to 45min, depending on the size of the aortic valves, with improved

shape fidelity from 66.6% for small valves (12mm inner diameter) to 93.3% for larger valves

(22mm inner diameter). Porcine aortic valve interstitial cells were subsequently seeded onto the

scaffolds and cultured for up to 21 days.

Other complex structures were also created by printing cell-free and cell-loaded alginate droplets

and building-up into branching constructs 600µm wide and approximately 500µm tall, with

roughly 90µm wide channels throughout the structure 51. Alginate droplets 50µm in diameter

were deposited onto a hydrated gelatin substrate acting as a Ca2+ reservoir (50mM Ca2+

concentration). The Ca2+ diffused upwards into the printed droplets, inducing gelation and

simultaneous fusion of the alginate beads to one another. Due to the coalescence of partially-

Page 20: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

9

gelled droplets, the printed features are often loss when droplets are deposited in a row-by-row

process. They adapted their printing sequence by splitting the planar matrix into four sequential

but alternating patterns, such that each printed droplet would be allowed to gel for tens of

seconds before a neighboring droplet was added.

Printed bio-ink droplets containing viable cells were patterned similarly using a layer-by-layer

approach onto bio-paper consisting of a layer of collagen type I from rat tail 39,55. The technique

takes advantage of the printed cells generating their own extracellular matrix and fusing to one

another into three-dimensional constructs 43,69. The bio-ink particles containing cell suspensions

are made of cell pellets that are incubated to form what they call a cellular “sausage” 69. These

were cut into cylinders which were left overnight to round into spheres 500µm in diameter. The

cell spheroids were printed one-by-one, and the collagen bio-paper was essential between each

spheroid layers to promote fusion of the individual cell beads. Since the particle fusion relies on

cell movement, too rapid cell motion resulted in fragile structures, with distortion resulting from

uneven gelation of successive collagen sheets. This resulted in reduced precision beyond a few

layers, with the need to remove the collagen sheets after printing proving to be challenging.

Norotte et al. created vascularized constructs using the same cellular spheroids printing

approach 46. Improved fusion time between the cellular building-blocks was achieved by

preparing cellular cylinders rather than spheres (2-4 days fusion versus 7 days). These were

prepared by printing cell-suspensions of a bio-ink into non-adhesive Teflon or agarose molds,

and incubation overnight. Template agarose rods were 300-500µm in diameter were prepared

and simultaneously printed with the cell-populated rods, requiring their manual removal to

produce vascularized networks. Final constructs resulted in tubes with 900µm inner diameter and

300µm wall thickness. The need for agarose rod removal post printing limits the complexity of

the vascular networks that can be printed. In addition, the additional time required to prepare a

large number of cellular spheroids and rods before printing, and a wait time of up to a week for

the fusion of spheroids makes this printing strategy timely. In addition, the large diameter of the

cell spheroids and cylinders lead to apoptotic cells throughout the vascular construct after 3 days

of fusion.

Page 21: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

10

2 Chapter 2 - Formation of Aligned Soft Material Sheets

2.1 Fibrillar Structure of Materials in Nature

Nature possesses the unique ability to organize tissues with respect to their cellular and material

composition. In plants, animals and humans, biological tissues possess a hierarchical

organization of their extracellular matrix with characteristic length scales that often extend across

six orders of magnitude: from macromolecular length scales of tens of nanometers to tissue

length scales of tens of millimeters. In several tissues, the multiscale organization of the

extracellular matrix relies on a high degree of molecular alignment to satisfy critical functional

requirements 1. The molecular composition and mechanical properties of selective biological

tissues are summarized in Table A1. For instance, palm trees have been able to survive tornados

due to their fiber-reinforced composite structure 71-73. Many mammalian tissues are characterized

by fibrillar alignment, e.g., blood vessels 74,75, the cornea 76-78, skin79-81, and the tendon 1,82,83.

A key contributor to achieving the tensile properties associated with connective, epithelial, and

muscular tissues found in blood vessel, skin, and cardiac tissues, is associated with the multiscale

organization of collagen. Collagen accounts for 25-35% of the total protein mass in mammals

and is one of the main components of the extracellular matrix (ECM) 75,84. The collagen family

consists of 28 different proteins, with type I represents over 90% of the weight of all collagen in

humans 75,84. The mechanical properties of tissues, i.e., elastic modulus, elasticity, and strength,

are highly tissue-specific and strongly influenced by the collagen fibrils 85. Fibrils are composed

of three polypeptide strands (alpha peptides), each in the form of a left-handed helix. The three

helices self-assemble into a right-handed triple helix through a process dictated by the

distribution of polar charged and hydrophobic amino acid residues in each strands, forming the

collagen molecule that is approximately 300nm in length and 1.5nm in diameter 86,87. These

molecules self-assemble through an entropy-driven process known as fibrillogenesis, forming

fibrils with diameters ranging from 20nm to 70nm 75. The collagen fibrils in series are separated

by a gap zone of approximately 40nm, and the adjacent collagen fibrils within these fibers are

staggered from one another with a periodicity known as the D-period that is approximately 54-

67nm long 84. Collagen fibers have diameters between 10nm and 300nm and constitute of

microfibril functional units 86.

Page 22: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

11

2.2 Collagen Organization in Tissues

The higher order organization of collagen fibrils varies between tissues, and its structure, size,

and orientation are closely linked to functional tissue characteristics. The different types of

collagen organization can be divided into two general categories: those subjected to

unidirectional tensile stress, and those subjected to multidirectional stresses75. The first group

consists mostly of large, heterogeneous fibrils that are tightly packed in parallel and subjected to

tensile stress in the direction of their axis. These include highly tensile structures such as

tendons, ligaments and bone. The second group consists of small, homogeneous fibrils arranged

in helical wavy bundles forming a three-dimensional network, enabling them to be tough yet

highly compliant to multidirectional stresses. Tissues falling within this category include the

cornea, blood vessel walls, skin and nerve sheaths.75

Collagen fibers in the tendon follow a straight and highly aligned organization along the

longitudinal axis of the tendon. These collagen bundles however possess a microscopic crimped

morphology travelling parallel to the axis 82,88. Most of the tendon structure is made up of

collagen type I (>95%) 88, which molecules are arranged and polymerized as fibrils, fibers, and

fiber bundles 83. The organization of these bundles in terms of packing density, diameter, and net

orientation is a determinant factor of the tendon function 82. In the rat tail tendon, collagen fibers

are aligned with the long axis of the tissue and grouped into bundles approximately 61.9-

123.8µm in diameter 82,88,89. These bundles form thick, straight, parallel fascicles 80-320µm

diameter, each possessing a planar-zig-zag structure with a wavelength of approximately

175µm 75,90. The crimp angle was found to be around ±20 ̊ to the longitudinal plane of the

fascicles, with the angle decreasing to ±12 ̊ as it approaches the center of the fascicles 89. At

small strains (2-13% depending on the age of the mice), a very small stress (up to 2MPa) is

sufficient to elongate the tendon by straightening the crimped architecture 90. Under larger

strains, occurring closer to the center of the fascicle, the stiffness of the tendon increases

considerably with extension 1,90,91.

Collagen fibers found in the cornea follow a highly aligned morphology 73,76,78. The cornea is

composed of an epithelium and endothelium separated by a tough collagenous stroma about

500µm in total thickness 77. This dense and organized collagen structure (mainly composed of

Page 23: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

12

type I and V) makes up 90% of the corneal volume and 70% of its dry weight, and is arranged in

orthogonal lamellae 76,78,92. Each lamella contains a highly oriented array of collagen fibrils of

homogeneous diameter. In the bovine cornea, these fibrils are around 36nm, and composed of

approximately 4nm diameter microfibrils that are tilted by about 15 ̊ to the fibril long axis in a

right-handed helix 73. The human cornea is composed of smaller fibrils that are 15.4 ± 0.5nm in

persons younger than 65 years of age, and 16.1 ± 0.5nm in persons older than 65 93. In humans,

lamellar size varies considerably as a function of depth within the stroma. Anterior lamellae are

0.5-30µm wide and 0.2-1.2µm thick, whereas those situated in the posterior stroma are 100-

200µm wide and 1.0-2.5µm thick 92. As a result of this dense collagen fibril packing, the stromal

layer of the cornea possesses an architecture that effectively resists tensile loads and internal

swelling pressures of up to 60mm Hg 77, giving the human cornea an elastic modulus of

15.9 ± 2.0MPa and UTS of 3.3 ± 0.2MPa 94,95.

In other tissues such as skin and the intestine wall 75,86,96-98, the wavy collagen fibers are not

aligned in a preferential direction and instead form an almost random woven network. Skin

consists of an upper layer, the epidermis, and a lower layer, the dermis. In the latter, 85% of the

dermal content consists of collagen type I, which plays an essential role in maintaining the

strength and elasticity associated with intact skin 80. Unlike the tendon, the collagen fiber

organization in skin is more random with a complex network of interlaced fibrils that provides

structural support to the epidermis and gives skin its firmness 84,97. The ability of the collagen

fibers to reorient allows large extensions of the tissues 99. As a result, the tissue also becomes

progressively stiffer and its orientation of the collagen fibers more aligned with the direction of

stretching.

Collagen present in the vasculature is crucial for the determination of the tensile strength and

stiffness of blood vessels. The vascular wall is mainly composed of collagen type I and III, and

their removal was found to reduce up to 50 times the local stiffness in the aorta, as measured by

atomic force microscopy (AFM) 100. In the rat aorta, collagen fibrils were measured to be

81.7 ± 7.6nm in diameter 101. These collagen fibrils form a “helical” arrangement with a winding

angle of approximately 18 ̊ 75. In addition, these fibrils are organized into wavy lamellae where

the collagen fibers are generally aligned in the circumferential direction. This circumferential

organization is characterized by a crimped structure that, at low wall tension, is elongated and

straightened with increasing wall tension 96,98,102. At low transmural pressures, this capacity for

Page 24: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

13

distension without a significant increase in stiffness 102,103, provides physiological levels of

compliance that serve to dampen the pulsatility of blood output from the heart74,75. At high

transmural pressures the vanishing pitch and crimping significantly increase the stiffness and

allow large blood vessels to withstand high burst pressures 103. Large veins display a drastic

increase in stiffness when subjected to pressures beyond 30-50mm Hg as opposed to pressures

beyond 200mm Hg for arteries 102.

2.3 Strategies for Controlling Collagen Self-Assembly In Vitro

Collagen fibrillogenesis depends on electrostatic and hydrophobic interactions that are strongly

dependent on temperature ionic strength, pH and temperature 104. The controlled multiscale

assembly of collagen fibrils in vitro remains a major challenge. Particularly the difficulty in

consistently promoting high degrees of fibrillar alignment and compactness result in low

mechanical properties (e.g. elastic modulus and ultimate tensile strength) compared to native

tissues.

Early protocols for collagen gel preparation favored the inclusion of viable cells. This required

the formation of these gels in culture media, at neutral pH, and cultured in an incubator, thereby

limiting the range of ionic strength, pH, and temperature. This resulted in weak collagen gels

(UTS of 0.01MPa and E of 0.1MPa) that required from 3 weeks up to 3 months of culture until

cellular remodeling increased the elastic modulus of the construct to levels that allowed

manipulation 105-107. Achilli and Mantovani investigated the effect of varying the ionic strength,

pH and temperature and optimized these parameters to 174 mM salt solution, pH 10, and 4 ̊C to

achieve improved mechanical properties of 10kPa UTS and E of 100kPa 104.

In addition to the magnitude of tensile properties, their directional dependence plays a key role in

controlling cell behavior 108,109. For instance, the presence of aligned collagen fibers in the

vascular wall acts as a signaling factor for platelet activation 75,110. Collagen alignment was also

shown to play an important role in directing cell proliferation and migration in vivo after

injury 111. In order to mimic the natural and tissue-specific structure of the ECM such that both

the mechanical properties and bioinductive aspects of the synthetic scaffold can be tailored, it is

vital to be able to manipulate the spatial organization of the cellular microenvironment.

Page 25: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

14

Several reports have focused on achieving various degrees of anisotropic collagen fibril

alignment through: application of anisotropic mechanical forces through fluid flow-induced

shear stress 112-116 or tensional forces 112,117,118, geometric confinement 119, electric currents 120,

magnetic fields 121-126, and electrospinning 74,127-129. The structural and mechanical properties of

the resulting collagen gels produced are summarized in Table A2.

Caves et al. 112 demonstrated the continuous fluid-extrusion of collagen fibers at a rate of

60m/hr. Fibers were collected on a spinning mandrel that exerted a tensional force in the

direction of extrusion and were manually transferred to a collection frame that maintained

tension during incubation and drying. The fibers had a high degree of fibrillar alignment, a cross

section of 53 ± 14µm 21 ± 3µm, UTS up to 94 ± 19 MPa and E up to 775 ± 173 MPa. Lai et al.

combined hydrodynamic shear and translation-induced shear to generated aligned collagen slabs

that were approximately 1mm 1.2mm 30mm. Acidic solutions of collagen with

concentrations ranging from 0.2mg/ml to 0.8mg/ml were extruded from a syringe onto a

translating glass slide immersed in a salt buffer solution at pH 7.4. The hydrodynamic shear

induced alignment of the collagen monomers and fibrillogenesis was triggered from the change

in pH from acidic to neutral. As a result, the collagen formed were highly aligned and resulted in

elastic modulus of 3.6 ± 1.9MPa 116.

Cell-laden collagen gels have been subjected to anisotropic mechanical forces for the formation

of aligned collagen gels with anisotropic mechanical properties. Thomopoulous et al. constrained

fibroblast-laden collagen sheets (4cm × 4cm) uniaxially or biaxially for 72h and reported gel

compaction in the unconstrained axis, with an obvious structural and mechanical anisotropy 118.

The authors found collagen gels to develop anisotropic molecular alignment under uniaxial

tension even in the absence of cells.

Geometric confinements have also been utilized to define collagen fiber alignment during

fibrillogenesis. Lee et al. used this approach to generate three-dimensional collagen gels with

aligned fibers for the investigation of cell signaling in vitro 119. Collagen gels were cast and

geometrically confined within 1cm long, 40µm deep, and 10-100µm wide microfluidic channels.

Under static flow conditions at room temperature, the first fibers appeared 10min after casting

and fibrillogenesis was completed after 30min. These narrow microfluidic channels resulted in

Page 26: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

15

the formation of 20µm long collagen fibers that were 20-40% aligned within 5 degrees from one

another.

Electrochemical cues have been similarly employed to create aligned and densely packed

elongated collagen bundles. Cheng et al. used parallel wire electrodes to apply electrical currents

(current density 0.55A/m2, nominal electric field strength 2.5kV/m) to collagen solutions spread

onto a glass slide. The currents created a pH gradient within the footprint of the glass slide,

causing the collagen molecules to migrate and assemble within a plane 120. This approach

enabled the formation of single collagen bundles 50-400µm in diameter and 3-7cm in length

depending on the length of the electrodes, achieving half the mechanical strength of the native

tendon (UTS of 24-88MPa and E of 277-671MPa).

Due to the small negative diamagnetic susceptibility of collagen molecule, the exposure of

collagen solution to a strong static magnetic field on the order of a few teslas (T) during gelation

aligns collagen fibrils perpendicular to the direction of the field 122,130,131. Torbet et al. subjected

collagen solutions to a 7T magnetic field for a duration of 30min in an incubator at 37°C 126. The

authors repeated the process by rotating the collagen samples along the z-axis and adding an

ungelled layer of collagen solution onto the already gelled and aligned collagen films. This

resulted in the formation of a laminated structure of varying fibril orientations that mimicked the

organization of the stroma in the cornea. Others have demonstrated the effect of interstitial fluid

flow on collagen gels, where initially entangled fibers were successfully disentangled and

aligned perpendicular to the flow field 132,133. Guo et al. combined magnetic field and fluid-flow

effects to create collagen films with aligned microfibrils 122. They incorporated 2.5µm diameter

magnetic beads within collagen gels, utilizing the movement of the beads attracted to an external

magnetic field and fluid-flow to pull the collagen fibers during gelation.

Electrospinning is a well-established technique used by many researchers to prepare nanofibrous

matrices for tissue engineering applications 74,128,129. Although the resulting fibers are generally

random in orientation, it is possible to control the degree of collagen fiber alignment by

collecting the spun fibers onto a rotating mandrel 74,128. Oryan et al. created highly aligned

collagen nanofibers by combining electrospinning with magnetic fields to reproduce the collagen

fibrils in the tendon 127. The film of electrospun collagen fibers formed was subsequently

embedded into collagen solution and polymerized at 4 ̊C for 48hr under 12 T magnetic fields to

Page 27: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

16

induce fibril alignment. Final collagen gel composite were cut to the size of rabbit’s Achilles

(L = 2cm, H = 3.5mm, W = 3mm) before implantation.

These approaches promote collagen alignment in small dimensions between 0.05-3mm, but with

a limited degree of fibril alignment and packing density the gelled collagen has a low mechanical

strength in comparison to native tissue (E and UTS up to 775MPa and 94MPa respectively 112,

Table A2), and often require hours to complete gelation 134,135. There is therefore a currently

unaddressed need for the controlled alignment of anisotropic matrices in centimeter scale in

several tissue engineered applications.

2.4 Strategy for the Continuous Formation of Collagen Sheets with Aligned Nanostructure

Here, we present a method for the continuous formation of nanofibrous collagen sheets with

precise control over the sheet thickness, width, and the degree of fibril alignment. Our strategy is

illustrated in Figure 1, and consists of the continuous in-flow formation of collagen sheets using

a multilayer microfluidic device with a flow constriction that attached downstream. A layered

fluid consisting of a central sheath of collagen solution sheathed by streams of PEG solutions on

the top and bottom, simultaneously exited the microfluidic device. Downstream of the device

exit, the focusing and collagen solutions were hydrodynamically focused in the sheath-normal

direction, collagen fibrils were formed and aligned in the flow direction and cross-linking

progressed starting from the top and bottom boundaries where the collagen solution was in direct

contact with the focusing PEG solution. A flowable thin sheet of cross-linked collagen was

formed and an initial compaction of the collagen solution achieved. A strain was applied on the

partially cross-linked sheet when collecting it onto a drum at a location downstream of the

confinement. We report the obtained changes in the tensile properties of the collagen sheet as a

function of the degree of fibril alignment and compaction, and demonstrate the scalability of this

approach to form meter-long highly aligned collagen sheets with very large aspect ratios (sheet

width divided by thickness) of up to 400.

Page 28: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

17

2.4.1 Materials and Methods

2.4.1.1 Experimental Setup

Figure 2a shows a schematic illustration of the experimental setup, which consists of a multilayer

microfluidic device, a constriction manifold and a rotating drum. Two syringe pumps

continuously supply the collagen solution and the focusing solution at their respective flow rates,

QM and QF. The multilayered microfluidic device distributes through microchannel networks in

separate device feature layers (Fig. 2b) a focusing solution (top), a collagen solution (center) and

a second focusing solution (bottom). These solutions meet at the device exit, form a layered or

sheath flow, and are hydrodynamically focused in the sheath-normal direction while passing

through a downstream flow constriction with a gap height of HC =1mm (Fig. 2c). Prior to the

constriction, an opening gap HG = 4mm was introduced to allow some degree of freedom in the

alignment of the device exit region with the constriction manifold. The continuously produced

collagen sheet was pulled onto a rotating drum, decreasing the sheet thickness and enabling the

degree of collagen fibril alignment and their packing density to be precisely tailored. The flow-

focusing constriction plays three roles in our system by (1) inducing through hydrodynamic

focusing the alignment of collagen in its monomeric state, by (2) preventing through

hydrodynamic focusing any unwanted deformation of the collagen liquid sheet that would be

expected in the case of an elastomeric substrate material 136, thereby maintaining thickness

uniformity of the collagen sheet, and by (3) stabilizing the co-flow of the focusing and collagen

solutions by isolating the co-flowing system from any flow recirculation within the liquid filled-

reservoir created by the rotating drum. The microfluidic device and constriction manifold are

immersed in a reservoir filled with the same liquid as the focusing solution. As the collagen and

focusing solutions meet at their common interface, the specific composition of the focusing fluid

(10% w/v PEG at pH 8) triggers the gelation of collagen, thereby fixing the structural change

imposed on the collagen sheet. The composition of the focusing fluid was chosen after numerous

optimization studies by our collaborators from Dr. Chaikof’s lab for optimal and rapid collagen

fibril formation 112,137,138. Such manipulation of the material structure results in the formation of

collagen sheets with a wide range of mechanical properties directly linked to the degree of their

fibril alignment and packing density.

Page 29: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

18

2.4.1.2 Fabrication of Microfluidic Device and Constriction Manifold

The microfluidic device was fabricated using standard soft-lithography techniques 139 and

consists of three polydimethylsiloxane (PDMS) layers that were individually fabricated and

subsequently bonded to form the final multilayered device. The top and bottom layers distribute

a flow-focusing solution, while the middle layer distributes an acid solubilized collagen solution

(Fig. 2b). On the top and bottom sides, the collagen solution is confined by a layer of the

focusing solution as it exits the device and enters a liquid-immersed constriction unit.

Hydrodynamic focusing takes place at a location downstream of the microfluidic device within a

constriction unit that is 12mm wide, has a LG=2mm long section with gap height HG=4mm and a

LC=6mm long flow constriction. The horizontal distance between the end of the constriction unit

and the edge of the rotating drum (LP) (Fig. 2c). We machined a device-external constriction unit

in aluminum in order to retain a uniform constriction height, HC, across the 1:12 aspect ratio slit,

and thereby avoid any unwanted deformation that would be expected in the case of an

elastomeric substrate material 136. The constriction unit also prevented any flow instabilities

within the co-flowing system caused by flow recirculation induced by the rotation of the drum

within the liquid-filled reservoir. The constriction gap was horizontally aligned and tightly sealed

against the exit section of the microfluidic device. The value of HG exceeded slightly the height

of the device exit section by approximately 2.5mm to ensure fluids from all three layers are

consistently guided through the constriction.

2.4.1.3 Collagen Gelation through Molecular Crowding

In this work, we rely on a phenomenon known as molecular crowding to initiate the rapid

gelation of collagen at room temperature, enabling physical handling of the collagen sheets

immediately after they exit the microfluidic device. In past studies, the assembly and

disassembly of biomolecules of cytoskeletal filaments have been investigated 140, and revealed

that the self-association of molecules into organized bundles of actin filaments is highly

influenced by volume exclusion and confinement provided by the crowded cell environment 141-

143. Cuneo et al. utilized 6000kDA polyethylene glycol (PEG) molecules to mimic the

macromolecules of the cell and found that concentrations of PEG between 6 to 7% w/v resulted

in massive conversion of actin filaments into bundles 140. Self-assembly and organization of

Page 30: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

19

collagen fibrils can be controlled in vitro through either geometric confinement that concentrates

pure collagen monomers, or in open spaces using fibril forming buffers. These fibril forming

buffers essentially consists of a solution containing macromolecules that occupy a significant

fraction of the volume of the medium, thereby commonly referred to as ‘crowded’ solutions.

These can be produced by adding high concentrations of a synthetic or biomolecular co-

nonsolvent such as polyethylene glycol or hyaluronic acid 143,144. Due to the large size of the

crowding molecules, little free space is left in the solution, causing volume exclusion and

confinement, thereby forcing the collagen molecules to align themselves in order to maximize

entropy.In 1994, Cavallaro et al. adapted a collagen fibril extrusion developed by Kato et al.

where collagen was extruded into a fibril forming buffer 145. Cavallaro have used polyethylene

glycol (PEG) as the fibril forming buffer and demonstrated that molecular crowding of the

collagen monomers was induced by this hypertonic environment (i.e. higher osmotic

pressure) 144. The osmotic pressure difference between the PEG and collagen solution generates

an osmotic pressure gradient, leading to the dehydration of the collagen sheath through depletion

forces, which results in compaction and coagulation of the collagen. As a result, the collagen

fibers produced were denser and could therefore be continuously extruded without rupture.

Following this observation, other groups have utilized the same approach to generate organized

collagen fibers 112,143,146-148. In our approach, in addition to PEG as a molecular crowding agent,

the focusing solution used is also pH and salt-balanced (see materials and methods section),

ensuring the simultaneous self-assembly of tropocollagen molecules (consisting of a triple helix

composed of three alpha-peptide strands) into aligned fibers through the process of

fibrillogenesis. As the collagen sheath is focused into the constriction, hydrodynamic flow

focusing and strain-induced pulling align the collagen in its monomeric state. The process of

fibrillogenesis occurs through a longer time period through the pH and salt concentration change

as they diffuse from the focusing solution into the collagen sheath. In recent works, salt and pH

induced fibrillogenesis was observed to form after 10min at room temperature 119.

2.4.1.4 Isolation and Purification of Monomeric Collagen

Acid-soluble, monomeric rat-tail tendon collagen (MRTC) was obtained from Sprague-Dawley

rat tails following Silver and Trelstad 149. Frozen rat tails (Pel-Freez Biologicals, Rogers, AK)

were thawed at room temperature and tendon was extracted with a wire stripper, immersed in 10

mM HCl (pH 2.0; 150 mL per tail) and stirred for 4 hr at room temperature. Soluble collagen

Page 31: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

20

was separated by centrifugation at 30,000g and 48 ̊C for 30 minutes followed by sequential

filtration through P8, 0.45 µm, and 0.2 µm membranes. Addition of concentrated NaCl in 10 mM

HCl to a net salt concentration of 0.7 M, followed by 1 hr stirring and 1 hr centrifugation at

30,000g and 48 ̊C, precipitated the collagen. After overnight redissolution in 10 mM HCl the

material was dialyzed against 20 mM phosphate buffer for at least 8 hr at room temperature.

Subsequent dialysis was performed against 20 mM phosphate buffer at 48 ̊C for at least 8 hr and

against 10 mM HCl at 48 ̊C overnight. The resulting MRTC solution was stored at 48 ̊C for the

short-term or frozen and lyophilized.

2.4.1.5 Preparation of Collagen Neutralization Buffer

The rapid gelation of collagen sheets during extrusion was induced by the addition of

polyethylene glycol (PEG) in the focusing solution. PEG triggers collagen molecular crowding

and gelation150, a phenomenon only seen in a much smaller scale during collagen wet spinning

(~10s of microns), but yet demonstrated on a macro-scale 112,137. The focusing solution consisted

of a neutralization buffer, which that contained 10%w.t. PEG (MW 35kDa), 4.14mg/mL

monobasic sodium phosphate, 12.1mg/mL dibasic sodium phosphate, 6.86mg/mL TES, and

7.89mg/mL sodium chloride.

2.4.1.6 Collagen Sheet Incubation and Drying

After collagen extrusion and pulling onto the rotating drum, the sheets were collected and

immersed in flow focusing for 1 hr, after which they were washed three times with ddH2O.

Sheets were subsequently incubated in a fiber incubation buffer (FIB) (7.89 mg/mL sodium

chloride, 4.26 mg/mL dibasic sodium phosphate, 10 mM Tris, pH 7.4) at 37 ̊C for 48 hr.

Following incubation, the collagen sheets were rinsed in ddH2O for 1 hr and dried on a glass

slide under constant forced air flow.

Page 32: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

21

2.4.2 Results

The presented approach enables the continuous formation of collagen sheets with a controlled

width, w, thickness, , and angle of fibrillar alignment, . The sheet width was determined at the

drum from measurements performed with three microfluidic that had the exit widths: w0=5mm,

10mm, and 25mm. The thickness and fibril alignment of the collagen sheets depended on the

following experimental parameters: the collagen flow rate, QM, the flow rate of the focusing

fluid, QF, and the pulling velocity, VP. In the following, we will assess the roles of hydrodynamic

focusing and stretching on the formed sheets.

2.4.2.1 Flow Confinement

We hypothesize that the initial thickness reduction is dependent on the re-organization of the

focusing and collagen streams while they pass through the confinement. In order to

experimentally characterize how the local sheet thickness locally varies at different locations

downstream of the exit section of the microfluidic device while the collagen sheath flows

through the constriction and pulling-induced strain is being applied, a separate constriction unit

was fabricated. The experimental setup consisted of a vertically arranged microfluidic device,

provided for visual access within the constriction and allowed the collagen sheet thickness

variation to be imaged in the (x, z)-plane (Fig. 3a), using an inverted microscope. Figure 3b

shows bright-field images of the exit region of a microfluidic device placed within the vertically

positioned constriction unit (top), and of a collagen sheet being formed within the constriction

(bottom). The bottom image was captured for a device with w0=10mm at conditions QM =

100µl/min, QF = 1ml/min, V* = 4.5. Here, we introduce a non-dimensionalized velocity

parameter V* obtained by relating the pulling velocity with the total velocity of the working

fluids. Specifically V* = (VP – VTotal)/VTotal, where VTotal = (QF + QM)/AConst, and the cross-

sectional area at the constriction AConst = W × HC. The flow profile of the focusing solution

within the confinement was further visualized by incorporating fluorescent microbeads (Nile red

carboxylate microbeads 1µm in diameter) at a concentration of 0.08% v/v. Long-term exposure

images (exposure time 400ms) captured the streamlines within the two regions of interests that

are indicated in Figure 3c. Specifically, streamlines within the entrance region of the sheath flow

entering the constriction (window 1) and the upper wall of the chamber before the constriction

Page 33: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

22

(window 2) were investigated. The images shown in Figure 3c-1 were obtained at

QM = 100µl/min, QF = 1ml/min, V* = 10 and illustrate the streamlines of the focusing fluid travel

parallel to the moving collagen sheath in its proximity. In Figure 3c-2, no collagen was flown

through the microfluidic device and QF = 1ml/min. The presence of recirculating flows can be

observed in the upper wall of the open region before the constriction. The size of the

recirculation zone decreased when increasing QF from 1ml/min to 6ml/min (see Appendix Fig.

A2). However, the recirculating vortices do not interact with the collagen sheath, and follow a

laminar flow-profile with linear streamline parallel to the flow direction as they pass through the

constriction region, suggesting that the formation of collagen with consistent control over the

width and thickness is unaffected by their presence. A numerical model of the velocity profile

within the constriction region was developed using a multiphysics solver based on the finite

element method (COMSOL). The collagen sheet was treated as a moving wall with velocity

equal to the pulling velocity and no-slip boundary condition. The focusing solution (10% w/v

PEG) was treated as a Newtonian fluid with a density of 131mg/mL and constant dynamic

viscosity of 54.4cP confirmed from viscosity measurements under increasing shear rates using a

rheometer (DV-III Rheometer, Brookfield, Massachusetts, US) (see Appendix Fig. A3a). The

velocity profile of the focusing solution was investigated, with QF = 1-16ml/min and

VP = 2mm/s, with Reynolds number ranging from 0.02-0.32 confirming that the experimental

conditions remain within the laminar region (fig. A3b,c). In agreement with the experimental

data, the numerical data showed no backflow or recirculation within the constriction region HC.

2.4.2.2 Inducing Alignment

We experimentally investigated the increase of fibril alignment and compaction within the

collagen sheet with increasing strain being applied by the pulling drum. The vertically arranged

microfluidic device and constriction unit were used to produce collagen sheets at varying pulling

velocities V* while was measured from bright-field microscopic images. Figure 3f shows the

results obtained at different values of V* and QM = 100µl/min, QF = 0.5 and 1ml/min.

Measurements were taken at four different streamwise locations within the confinement: (A) the

microfluidic device exit, x = 0, (B) x = 0.5LG, (C) x = LG+0.5LC, (D) x=LG+LC+LP. For all

considered values of V*, as the focusing flow rate QF increases, the overall thickness of the

collagen sheet decreases, suggesting a focusing and stabilizing effect induced by the focusing

solution. For both QF at low V*= 0.1-2, the collagen sheet thickness increases slightly as it exits

Page 34: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

23

the device into the opening region within LG (B), similar to the ‘die swell effect’ observed in

flow extrusion where the melt expands as it exits from the die 151,152. However, at higher pulling

velocities, V* > 2 (VP>5mm/s), the sheath thickness monotonically decreases. We attribute this

effect to the rapid onset of collagen fibril formation and gelation locally within the regions of the

collagen sheath that are in direct proximity to the PEG solution. We assume molecular crowding

to induce the rapid gelation of collagen within the thin outermost regions of the collagen sheath,

resulting in the thickness of the collagen sheath along its entire length to be affected by an

increased pulling velocity exerted at the drum.

2.4.2.3 Sheet Formation

The effect of the constriction unit on sheet thickness and width was investigated experimentally

by comparing results of collagen sheets formed with and without the use of the flow-focusing

unit. The experiments were conducted using three devices with w0=5mm, 10mm, and 25mm. For

the device with w0=5mm conditions QM=50µl/min, QF at 1ml/min, and VP = 1-20mm/s were

applied. In the case of the two other devices, the same range of VP was considered, and QM and

QF were adjusted proportionally with the increase in device width (i.e., the flow rates were

twofold higher in case of w0=10mm, and fivefold higher in case of w0=25mm), and the

corresponding V* were calculated accordingly The use of a constriction unit produced wider and

thinner collagen sheets. Collagen sheets formed without constriction were between

0.65±0.21mm and 3.3± 0.17mm wide. With a constriction they were at the same flow rates

between 3.3± 0.09mm and 17.3 ± 0.1mm wide (Fig. 4a). The constriction unit reduced by up to

88% the thickness of the produced sheets, from =260±8µm to 1140±10µm without and from

30±3µm to 213±15µm with constriction (Fig. 4b). The measurements of the external sheet

dimensions w and for all three devices were non-dimensionalized by w0 and HC, respectively.

The self-similarity of the results demonstrates the utility of the approach for the predictive

formation of a large aspect ratio collagen sheet with a certain target width, by selecting a

microfluidic device with an appropriate width w0. Sheet dimensions w and , were studied for

w0=10mm, V*=0.1-10, QM=100µl/min, and QF=1-6ml/min (Fig. A4). The obtained data suggest

a decrease in both width and thickness for an increasing flow rate of the PEG solution, QF, with

w/w0= 0.32-0.8, and /HC=0.025-0.3.

Page 35: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

24

2.4.2.4 Nano and Microscale Properties

At a distance LP downsteam of the confinement unit, the formed collagen sheet is being collected

on a rotating drum. Along with the applied flow rates QF and QM, the speed of drum rotation, V,

or the corresponding dimensionless parameter V*, affect not only sheet dimensions but also the

alignment of collagen fibrils and their packing density. The cross-sectional area of the wet

collagen sheets, w , was calculated, plotted against V*, and compared to the calculated cross-

sectional area. The calculated values were obtained from QM/VP (Fig. 4c). Data were obtained

from experiments were conducted with three device widths, s w0=5mm, 10mm and 25mm; where

QM = 50µl/min (w0/5mm) and QF=1.5ml/min (w0/5mm). For values of V* below a threshold V*th,

the measured cross-sectional area exceeded the one predicted under the assumption of a

conserved volume. For V* > V*th, the opposite case was observed, suggesting compaction of the

collagen sheet. The degree of compaction was estimated by comparing the measured cross-

section of the hydrated sheets with the initial sheet cross-section, equivalent to w0HC. Values of

compaction were found to range from 3.3 % to 95.5 % (Fig. 4d). This can be explained by the

relationship between the flow rates QM and QF with the pulling velocity VP. At an initially low

VP, the average total velocity of the collagen and focusing solutions through the constriction is

larger than the pulling velocity, suggesting that the fibril alignment is solely due to

hydrodynamic focusing and no strain is being exerted by the drum rotation. However, once VP

exceeds the average velocity of the collagen sheath leaving the microfluidic device, a strain is

applied by the pulling drum, that causes the alignment of fibrils along the length of the sheet, a

reduction of the average fibril-to-fibril spacing and a contraction of the sheet.

The degree of compaction and fibril packing density was also characterized based on

transmission electron microscopic (TEM) and scanning electron microscopic (SEM) images of

dried collagen sheets. Collagen samples were produced across a range of V* were examined, and

the TEM and SEM images revealed the degree of fibril alignment and packing density with an

increase in fibril packing density and alignment observed with V* increasing from 0 to 10 (Fig.

5a, b; see Fig. A6 for TEM and SEM of samples obtained at V* = 0.6 and 4.5). In addition, D-

period banding typical of collagen fibers can be observed in both TEM and SEM images of the

highly aligned collagen sheets produced (Fig. 5a-2, b-2). The degree of compaction was

measured by image analysis of the SEM images of collagen sheets formed at V* = 0.1, 0.6, 4.5,

and 10. An autocorrelation function was calculated for the intensity distributions in SEM and

Page 36: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

25

TEM images using the software program Matlab (Mathworks, Econometrics Toolbox, Natick,

MA, USA). Fibril spacing was measured from the resulting plots. As a sample, Fig. 5c shows an

autocorrelation function of the TEM image in Fig. 5a-2 (see Supplement Fig. A7 for all

autocorrelation plots). Fibril spacing for all V* conditions are summarized in Fig. 5d and indicate

a 95.3% decrease, from 139.3 ± 37.3nm for V* = 0.1 to 6.51 ± 1.16nm for V* = 10 (Fig. 5d

insert). It is important to note that the degree of compaction is much larger in the dried samples

as compared to the wet samples. We attribute this difference to the drying process post-extrusion

which further compacts the fibers through water evaporation. The D-period banding of collagen

fibers was calculated by applying an autocorrelation function to line intensity plot obtained in the

x-axis of the SEM image in Fig. 5e (V* = 7). A banding period of 67nm was obtained,

characteristic of collagen fibrils in the body and confirming that triple helical fibrils are formed.

In addition to the degree of compaction and the banding length, fibril alignment of the collagen

sheets was characterized by applying a Fast Fourier Transform (FFT) algorithm to the SEM

images obtained using an image processing software (ImageJ). The percentage of aligned fibrils

was plotted as frequency (%) versus the angle of alignment (Fig. 5f), confirming an increased

degree of alignment with increasing V* from 0 to 10. We have achieved up to 40% alignment of

fibrils within ± 5 ̊ from one another in collagen sheets with aspect ratio w/δ ranging from 15 to

375.

2.4.2.5 Macroscale Properties

The direct impact of fibril alignment on the overall mechanical properties of these collagen

sheets was confirmed through tensile measurements. Samples were prepared and mechanically

tested using an inverted DMTA (Dynamic Mechanical Thermal Analysis). Sample drying prior

to mechanical testing was critical to compact the fibers and prevent gel compaction in the

potential case where cells would be seeded on the surface (thickness drops to 1/10th of the initial

wet state). Sheets were placed on a glass substrate and subjected to a constant air flow (fan) to

accelerate the drying process. Figure A5 demonstrates the effect drying and incubation in fiber

incubation buffer (FIB) have on the mechanical properties of collagen sheets. The dehydration

process is shown to improve the upper tensile strength in both cases with or without incubation

in FIB post-extrusion. On the other hand, the Young’s modulus of samples without FIB

Page 37: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

26

incubation was reduced by 67% when dried, while when incubated in FIB, the drying process

brought the Young’s modulus back to its highest value post-extrusion. Mechanical testing of our

FIB-incubated and dehydrated samples was performed by rehydration with a drop of DI water,

lift-off from the glass slide, and immersion in PBS at 37˚C for 30min. Samples were kept

immersed in PBS at 37 ̊C throughout the experiment. As expected, the structural anisotropy

resulted in mechanical anisotropy with increased, significantly higher ultimate tensile strengths

(UTS) and Young’s moduli in the direction of alignment when compared to unaligned matrices.

Specifically, V* between 0.6-10 lead to the formation of sheets with ultimate tensile strengths

between 1.25-13MPa and Young's modulus between 1.3-130MPa (Fig. 5g). These values are

well within the range of those found for native blood vessels, 1.4-11.1MPa (UTS) and

1.5±0.3MPa (E).153

We have demonstrated a high throughput approach for the continuous formation of wide

collagen sheets (width:thickness ratio up to 400) with tunable alignment and compaction of

collagen fibrils. The combination of flow-focusing and strain-induced pulling of the collagen

sheet results in sustained collagen fibril alignment in the direction of flow, with the degree of

alignment and the density of fibrils consistent throughout the entire sheet width, a result

unprecedented at the macroscale. Large aspect-ratio collagen sheets with dimensions that ranged

from 3-12mm in width and 30-250µm in thickness were continuously produced. The degree of

alignment and compaction of the collagen fibrils was controlled, with up to 40% of fibers aligned

within ± 5° of one another, and up to 95.5% of compaction. As a result, these highly aligned

collagen sheets achieved mechanical properties comparable to native native blood vessel, with

E between 1.3-130MPa, UTS between 1.25-13MPa, and strain to failure from 15-35%. In

addition, the presence of D-banding periods of ~67nm typical of collagen fibrils was consistently

observed. We believe the presented strategy to be promising in developing large collagen

substrates of biologically relevant composition and tunable mechanical properties for

applications in tissue engineering such as vascular grafts.

Page 38: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

27

Figure 1. Schematic illustration of the flowable conversion of a collagen solution (shown in

orange) to a cross-linked collagen sheet with a high degree of molecular alignment (shown in

red). The illustrated approach involves (1) the reduction of the thickness of the collagen sheath

and initiation of fibril formation and cross-linking during hydrodynamic focusing in the sheath-

normal direction, (2) the application of axial strain between the constriction (blue arrow) and

collection on a rotating drum. The combination of these two steps, along with rapid gelation of

the collagen through molecular crowding, result in the formation of collagen sheets with high

aspect ratio of w/δ from 15 to 375, and highly aligned fibers (3).

Page 39: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

28

Figure 2. (a) Schematic illustration of experimental setup consisting of microfluidic device for

preparation of biopolymer sheet, constriction unit, and rotating collection drum to collect and

pull the cross-linked collagen sheet (VP). The drum is located at a distance LP = 20 mm from the

constriction unit. (b) Three-layer multilayer microfluidic device. Layers 1 and 3 distribute

focusing solutions (shown in green) and layer 2 distributes collagen solution (shown in orange)

to the device exit where the three fluid layers are brought in contact and cross-linking is initiated

(symbolically illustrated by the sheet color turning red). (c) A machined confinement unit (LG =

2 mm, LC = 6 mm, HC = 1 mm, HG = 4mm) reduces the thickness of the biopolymer layer, .

Schematic of collagen sheet viewed from the xy plane (top). Schematic of the sheet formation

within the confinement (bottom). Scale bars 10 mm (b) and 2 mm (c).

Page 40: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

29

Figure 3. Detailed study of flow-focusing and strain-inducing pulling using a vertical

constriction unit. (a) Schematic of experimental setup for imaging of collagen sheet formation

in the (xy)-plane. Sheat extruded into flow-focusing solution. (b) Bright-field images of collagen

sheet formation using vertically oriented manifold in (a). Images taken at device exit (top) and

within constriction region (bottom). QM = 100µl/min, QF = 1ml/min, V* = 4.5, w0=10mm. (c)

Schematic of regions (1) and (2) investigated using fluorescence microscopy. (1) Flow profile at

Page 41: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

30

entrance to constriction: QM = 100µl/min, QF = 1ml/min, V* = 10. (2) Recirculation profile

within focusing solution in entrance region to constriction, QF = 1ml/min. Fluorescent

microbeads added to focusing solution at a concentration of 0.08% v/v. (d) Sheet thickness

measured at four different locations within the constriction manifold. Data obtained with QM =

100µl/min, QF = 0.5ml/min and QF = 1ml/min (*), V*= 0.1, 2, 4.5, 10 (light to dark color bars).

Scale bars 250µm (b), 1mm (c), 200µm (c, 1-2), 2mm (d).

Figure 4. Measured collagen sheet width and thickness as a function of V* ranging from 0.1

to 10. (a,b) Collagen sheet width and thickness obtained with (full line) and without (dotted line)

use of constriction manifold. Three devices of varying exit width (5, 10, 25mm) were used with

the manifold. 5mm wide device was considered in the case without manifold. For the 5mm wide

device, QM = 50µl/min, QF = 1ml/min. QM and QF were varied proportionally with device width.

(c) Comparison of experimental and calculated collagen sheet cross-sectional area as a function

of V* ranging from 0.1 to 10. Dotted lines represent calculated value (QM/VP), while full lines

Page 42: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

31

represent experimental values obtained for 5mm (square), 10mm (triangle), and 25mm

(diamond) wide devices. (d) Degree of compaction measured as a percentage change in cross-

sectional area calculated from comparison between experimental cross-sectional areas and cross-

sectional area of the device exit section. Data plotted for all three devices (5mm, 10mm, 25mm)

as a function of V*.

Page 43: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

32

Page 44: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

33

Figure 5. Characterization of collagen fibril compaction, alignment, and resulting

mechanical properties. (a) TEM images of fibrillar alignment in collagen sheet obtained at V* =

0 (1), 10 (2). (b) SEM images of collagen fiber alignment obtained at V* = 0 (1), 10 (2). (c)

Autocorrelation function of a TEM image of collagen sheets produced at V* = 10 showing an

average spacing of ~6.5nm. (d) Degree of compaction quantified by autocorrelation of SEM

images of collagen fibers obtained at V* = 0.1, 0.6, 4.5, and 10. Compaction quantified as

percent change in fibril spacing in reference to the fibril spacing at V* = 0.1. Results plotted in

comparison to percent change in cross-sectional area in Fig. 4c (insert). (e) Autocorrelation

function of SEM image (insert) showing repeated banding pattern (D-period) of ~67nm. (f)

Collagen fibril alignment obtained from SEM image processing of sheets formed at V* = 0, 0.1,

0.6, and 10. Full width half max (FWHM) summarized in table insert. (g) Young’s modulus (E),

ultimate tensile strength (UTS), and strain to failure (%) of collagen sheets formed by passing

through constriction and subsequent alignment induced by different values of V* = 0.6-10

(*p<0.05). All experiments conducted at QM = 100µl/min, QF = 1ml/min. SEM and TEM images

were obtained by Dr. Grainger. Scale bars 200nm (a), 1µm (b, 1 left), 500nm (b, right), 50nm (b,

2-left), 500nm (e, insert).

Page 45: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

34

3 Chapter 3 - Continuous Formation of Patterned Soft Material Sheets Populated with Molecular and Cellular Payloads

Soft materials with a spatially non-uniform composition that is closely linked to their function

are abundant in nature. Such materials often possess a hierarchical architecture that extends from

cell to tissue scales in several directions. In this report, we demonstrate a one-step, continuous

process for the scalable formation of soft material sheets while controlling their local and global

composition. A ten-layer microfluidic device enabled us to first dynamically define mosaic

hydrogels by incorporating within a flowing biopolymer sheet a secondary biopolymer and to

retain the microstructure in a subsequent cross-linking step. The secondary biopolymer was

either a different hydrogel or it carried a biomolecular, colloidal or cellular payload. We

continuously organized hydrogel sheets to 2D and 3D soft material assemblies with millimeter to

centimeter length scales, stored information within unsuspended hydrogel sheets, incorporated

void regions, created mosaic stiffness and diffusivity patterns and populated tessellations with

different viable primary cells. We envision mosaic hydrogels to become continuous, automatable

and physiologically meaningful formats for engineering cell instructive microenvironments and

3D tissues.

3.1 Alginate in Tissue Engineering

Alginate, a naturally derived hydrogel forming polymer, has been widely used as a biomaterial

for tissue engineering and drug delivery applications 154-156, due to its gentle gelling kinetics and

low toxicity 157-159. Alginate is derived primarily from brown algae and is a linear polysaccharide

co-polymer of (1-4)-linked β-mannuronic acid (M) and α-guluronic acid (G) monomers. Within

this polymer, the M and G monomers are sequentially assembled in repeating MM or GG blocks,

or alternating MG blocks 160,161. Factors such as seaweed species, age, and section define the

amount and distribution of these M and G monomers. An attractive property of alginate is its

ability to rapidly form a hydrogel at room temperature and conditions that are mild to living

Page 46: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

35

cells. The gelation process is triggered by the cross-linking of alginate molecules with divalent

cations such as Ca2+. These calcium cations preferentially interact with G-block monomers in the

polymer chains to form ionic bridges between adjacent polymer chains, resulting in an “egg-box”

structure. This highly cooperative binding process requires more than 20 G-monomers 162. For

this reason, the ratio of M : G blocks is critical in defining the resulting mechanical strength of

the cross-linked alginate hydrogels. The mechanical properties of various types of alginate

(M : G ratio from 0.3-1.56) at concentrations ranging from 1.5-4% w.t. gelled with different

concentrations and types of cross-linkers have been summarized in Table A4.

In tissue engineering applications, the main function of alginate is to provide mechanical

integrity while transmitting initial mechanical signals to the cells and developing tissue. The

ability to tune the alginate mechanical properties by controlling the cross-linking time and M : G

ratio makes it an attractive material for a variety of tissues. Alginate has been widely utilized as a

scaffold material for engineered cartilage 163-165, bone 166,167, and skeletal muscles 168. It has also

been processed in the form of hydrogel beads to be used as a carrier matrix for encapsulating

molecules such as enzymes, drugs, microbial 169, or viable cells 170,171, with applications as

scaffolds for tissue engineering or the controlled release of drugs and biological molecules.

Bioprinting of alginate microdroplets 50µm in diameter was utilized to construct branched

microvasculatures 51. These were printed onto gelatin substrates soaked in 10mM CaCl2, relying

on the upward diffusion of Ca2+ cations into the printed droplets. The gelation of the alginate

beads was completed in approximately 8s, generating a fused bifurcated construct with a 90µm

diameter channel and 300µm thick walls. Due to the soft material properties, the hollow channel

collapsed slightly as the structure was built-up in height, resulting in a slightly deformed and

smaller opening of approximately 40µm in the upper region. Although alginate is a

biocompatible material non-toxic to the cells, it is limited in directing the fates of cells into

organized structures and cannot be remodelled by the cells 158. They incorporated collagen type I

into the alginate solution, relying on the biologically active property of collagen and the fast-

gelling characteristic of alginate. The rapid gelation of alginate prevents the diffusion and loss of

collagen, allowing it enough time to complete gelation at 37ºC. Once completed, the alginate

hydrogel was completely removed using a chelating agent (EDTA), leaving behind a structured

biologically relevant material that can be remodelled by the cells.

Page 47: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

36

3.2 Strategy for the Continuous Formation of Soft Material Sheets of Heterogeneous Composition

We report an approach for the tessellation and coding of planar soft materials that is scalable and

continuous, does not involve substrate support or moving device components, and is compatible

with a range of biopolymers and different cell types. Figure 6 shows a schematic illustration. A

microfluidic device dynamically incorporates one or several secondary biopolymer solutions

within one layer of a base polymer. The secondary biopolymers are either chemically distinct

from the base biopolymer or they carry a molecular, colloidal or cellular payload. Upon exiting

the device, the spatial organization within the fluid layer is retained via diffusion-mediated ionic

cross-linking and a mosaic hydrogel is formed. The presented strategy promises a one-step

process for information to be encoded, concentration gradients of diffusing or binding molecules

to be established, directionally dependent mechanical and transport properties to be realized and

cells to be co-localized and co-cultured within the same soft material substrate. Depending on the

choice of biopolymers, payloads, tessellations and microenvironmental conditions, the mosaic

hydrogel may either display time-constant or dynamically changing characteristics. Three-

dimensional bulk structures of homogeneous or organized heterogeneous composition can be

subsequently produced using the same microfluidic platform, in a single continuous step.

3.2.1 Material and Methods

3.2.1.1 Materials

Alginate (alginic acid sodium salt) and calcium chloride were purchased from Sigma-Aldrich

(St. Louis, MO, US). The alginate sample contained 2%w.t. alginate in a solution of 60% v/v

glycerol in DI water. The pectin-alginate solution was obtained by incorporating 1%w.t. pectin

(Sigma-Aldrich) into an aqueous solution containing 1%w.t. alginate and 65% v/v glycerol. The

crosslinking solutions consisted of 50mM, 100mM, and 150mM CaCl2 in DI water containing

65%, 63%, and 61% v/v of glycerol respectively. The density of all solutions was 1.168g/mL.

Two types of fluorescence microbeads were used either for continuously projecting wide-field

fluorescence images of the formed hydrogels from an upright fluorescence microscopic setup

(Nikon Eclipse E600, Nikon, Japan) onto a line camera (LC1-USB, Thorlabs, Newton, NJ, USA)

or for off-line characterization using laser-scanning confocal microscopy (Olympus IX81

Page 48: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

37

Inverted Microscope with FluoView FV1000, Olympus, Pennsylvania, USA). Specifically,

microspheres with mean diameter of 1μm with excitation/emission of 505/515nm and

535/575nm were purchased (F8852 and F8819, Invitrogen, Canada). Microbeads were added to

the biopolymer solutions at a ratio of 1:200, followed by 20min sonication (B5510-MT, Branson

Ultrasonics, Danbury, Connecticut, USA) to minimize aggregation.

3.2.1.2 Alginate Crosslinking Kinetics

Alginate gelation is triggered by an external crosslinking process during alginate extrusion into a

calcium chloride solution. As the two solutions come into contact at the device exit, the outer

surface of the sheet is instantly crosslinked, followed by further diffusive flux of calcium ions

from the reservoir solution into the gel to complete the gelation throughout the alginate sheet.

The maximum growth rate of the alginate gel crosslink density occurs within the first 15 minutes

of crosslinking 172. The gelation time is directly proportional to the alginate and crosslinker

concentration. Higher concentration of alginate directly translates to an increased number of

binding sites, therefore increasing the time needed for gelation. On the other hand, a higher

concentration of crosslinker results in a faster initial crosslinking rate due to the higher ionic

gradient created. Crosslinking then continues within the gel until either the ion source is depleted

or uncrosslinked positions in the gel are depleted.

Since the binding kinetics of calcium and alginate are so rapid compared to the diffusive

transport of calcium ions, the diffusion of calcium is the rate-limiting step in this gelation

process. In general, the diffusivity of calcium ions in porous gels and alginate is estimated to that

in water and has a value of DC = 1 × 10-9m2/s 173,174. This assumption can be made as no

significant diffusive resistance is present for molecules that are less than 20kDa (calcium

chloride molecular weight = 110.98Da). Mikkelsen and Elgsaeter developed a numerical model

to predict the calcium, alginate, and gel concentration over time 175. Assuming a homogeneous

system, the reaction should be consistent along the flow-direction. The analysis presented by

Mikkelsen could therefore be simplified to a one-dimensional diffusion analysis, with

concentration of the solutions dependent on time t and thickness of the alginate sheet x. For

simplicity, we assumed the diffusion coefficient of alginate to be much smaller than the diffusion

Page 49: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

38

coefficient of Ca2+ ions, therefore negligible. This resulted in the following equations predicting

the time-dependent concentration of Ca2+, alginate, and gel respectively.

(1)

(2)

(3)

(4)

(5)

(6)

Where c is the free Ca2+ concentration, a the free alginate concentration, g the gel concentration,

NC is the average number of Ca2+ ions per alginate-alginate dimer formation (=50), k is the

reaction rate (= 2 × 10-2 M-2s-1) 172.

The initial concentrations are zero for calcium and gel, and 1 for free alginate molecules. The

boundary conditions are summarized in equations (4) and (5) where in the center of the sheet (x

= 0), fluxes are equal to 0. At the external boundary of the sheet (x = δ), it is assumed that only

the free Ca2+ penetrate the alginate sheet and the fluxes of alginate and gel are therefore equal to

zero. Time-dependent concentrations of free Ca2+, alginate, and gel are summarized in figure 7.

The data obtained suggests complete gelation of a 100µm thick alginate sample after 15 minutes

exposure to the cross-linking solution from one side. Since our alginate samples are immersed in

a reservoir filled with cross-linking solution and exposed to diffusive transport of Ca2+ from all

directions, our extruded 200µm thick alginate sheets are predicted to fully gel after 15 minutes.

Page 50: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

39

For this reason, printed samples are kept immersed in the cross-linking solution for an additional

15 minutes before collecting them.

3.2.1.3 Neonatal Rat Heart Isolation

Neonatal Sprague–Dawley rats (1-2 day old) were euthanized according to the procedure

approved by the University of Toronto Committee on Animal Care. The cells from the heart

ventricles were isolated by treating with trypsin overnight (4○C, 6120 U/mL in Hanks’s balanced

salt solution, HBSS) followed by serial collagenase digestion (220 U/mL in HBSS)176. The

supernatant from five collagenase digests of the tissues was centrifuged at 750 rpm (RCF = 94 ×

g) for 4 min, resuspended in culture medium, and pre-plated into T75 flasks (Falcon) for 1 h

intervals to separate the adherent cells (non-myocyte) from the non-adherent cells (enriched

cardiomyocyte). Primary cardiac fibroblasts were obtained by cultivating for up to 7 days the

cells adhered to the T75 flask during the pre-plating. Culture medium for both cardiomyocyte

and fibroblast consisted of Dulbecco’s modified Eagle’s medium (DMEM) with 4.5 g/L glucose,

4 mM L-glutamine, 10% certified fetal bovine serum (FBS), 100 U/mL penicillin, 100 μg/mL

streptomycin and 10 mM 4-2-hydroxyethyl-1-piperazineethanesulphonic acid buffer (HEPES)

(Gibco, Invitrogen, Canada). Human umbilical vein endothelial cells (HUVEC) were purchased

from Lonza, Canada.

3.2.1.4 Cell Patterning

Cell were suspended in a 1:1 ratio of cell suspension solution and RGDS (arg-gly-asp-ser)

peptide-functionalized alginate solution. The cell suspension solution consisted of 12.3%v/v DI

water, 1.2%v/v glucose solution (0.3g/mL), 7.7%v/v 10x Medium 199 (Sigma-Aldrich, Canada),

1.1%v/v NaOH solution (1N), 2.0%v/v NaHCO3 solution (0.075g/mL), 0.8%v/v HEPES

(Invitrogen, Canada), 19.1%v/v MatrigelTM, and 55.9%v/v collagen type I from rat tail

(3.66mg/mL, BD Biosciences, Canada). The peptide-functionalized alginate solution consisted

of 1.5%w.t. RGDS-alginate and 0.08%w.t. collagen type I from rat tail. Peptide-functionalized

alginate was obtained following a previously described procedure177. Briefly, RGDS peptide

(American Peptide 44-0-14) was conjugated to alginate using carbodiimide chemistry with N-

Page 51: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

40

hydroxysulfosuccinimide ester (sulfo-NHS) stabilizer (Pierce, Fisher 24510). The resulted

solution was purified by dialysis, dried by lyophilize, and stored at -20 °C until use.

3.2.1.5 Cell Tracking

CellTracker Red CMPTX (C34552, Molecular Probes, Invitrogen, Canada) was used for

fibroblasts and CellTracker Green for cardiomyocytes (C2925, Molecular Probes). A 10mM

concentration of CellTracker dyes in DMSO was further diluted in serum-free culture medium

(DMEM) to create a working concentration of 10 μM. The cells were incubated in 1mL of dye

solution for 30 min at 37ºC in 5% CO2. Following the incubation step, the dye-cell suspension

was centrifuged and the pellet was washed two times with DMEM.

3.2.1.6 Immunofluorescence Staining

Cell samples were fixed in 4% Paraformaldehyde in PBS at room temperature for 15 minutes

followed by incubation in mouse anti-vimentin (Sigma, 1:100 dilution) overnight at 4°C.

Samples were then incubated with anti-mouse Alexa 488 (Sigma, 1:100) at room temperature for

1 hour and imaged with confocal microscope (Olympus FV5-PSU confocal with IX70

microscope, Canada).

3.2.1.7 Sample Preparation for Scanning Electron Microscopy

Hydrogel samples were fixed in 2% glutaraldehyde in a 0.05M sodium cacodylate buffer at

pH 7.4 for 1hr at room temperature, followed by gradual replacement of the liquid phase with

100% ethanol. Dehydration of the samples was achieved with liquid CO2 at 10°C in a critical

point dryer. Samples were subsequently heated to 31°C with a pressure increase to 7.2MPa,

transitioning the CO2 to supercritical fluid conditions. Lowering the pressure from the

supercritical state allowed a direct transition into the gas phase without causing any unwanted

liquid-gas phase transitions. The dehydrated sample was then transferred into a vacuum and

vapour-deposited with a thin film of gold to render the outer surface of the substrate electrically

conductive.

Page 52: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

41

3.2.1.8 Tensile Test

Samples were cut to lengths of approximately 20 mm and fixed with a cyanoacrylate adhesive

(Krazyglue Advanced Formula, Elmer's Products, Columbus, OH, USA) to cardboard strips,

which were vertically clamped between tensile grips for testing. A ramp of 0.1 mm/s was

applied using a 1000g load cell until failure.

3.2.2 Mosaic Hydrogel Formation

Mosaic hydrogels (thickness δ= 150 - 350µm, width ~3mm) were formed using a multilayer

microfluidic device along with the experimental setup shown in figure 8. Layers were

individually molded and vertically attached using a partial curing process 178, resulting in a 10-

layer-device that was able to withstand pressures of up to 600kPa (Appendix figure A1-b). The

center layer (indicated in blue color in figure 8b and as layer #6 in figure A1-b) carried to the

device exit via a set of parallel microchannels a time-varying content of biopolymer solutions.

Additional layers located above and below delivered, at the device exit, focusing streams

containing the crosslinker (indicated in green color in figure 8b). The produced biopolymer

sheet flowed into a liquid-filled reservoir which contained the same solution as the focusing

streams (Fig. 8a). To reduce the unwanted effect of flow instabilities at the device exit and to

ensure a uniform sheet thickness, δ, two co-flowing fluids were delivered from above and below

the soft biopolymer sheet in a flow-focusing configuration 179,180. The focusing fluids carried

cross-linking ions and induced gelation of the sheet. In a case study, we used a 2%w.t. alginate

solution, a biopolymer with well-known biocompatibility 159 and ionic crosslinking mechanism

181. To increase the fluid viscosity and render the produced biopolymer sheet neutrally buoyant

with respect to the focusing fluids, glycerol was added to both the biopolymer and focusing

streams, with the latter containing CaCl2 as the crosslinker. The focusing fluids were

continuously supplied by an annular gear pump (mzr-2921, HNP Mikrosysteme, Parchim,

Germany) at a rate of 8 mL/min. At a location approximately 50 mm downstream of the device

exit, the sheet was manually attached to a collection drum (21.3 mm in diameter) that rotated at a

constant tangential velocity, UP (Fig. 8a). The focusing fluid was subsequently stopped while the

hydrogel sheet continuously exited the device and was collected by the drum. Although the

shear stress exerted by the focusing fluid alone (Fig. A16) was sufficient to consistently form

hydrogel sheets, we relied on the rotating drum as this configuration allowed the continuous

Page 53: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

42

formation, image-based inspection, and collection of mosaic hydrogels. The sheet thickness δ

was dynamically controlled by varying the flow rate of the base biopolymer, QB, using a syringe

pump (model PHD, Harvard Apparatus, Massachusetts, US) and by varying UP (Fig. 8b). The

employed base biopolymers and their pore sizes (Fig. 8e,f) along with the increased alignment of

the polymer fibers due to the axial stress imposed by the pulling drum affected the elastic moduli

of the produced sheets (Fig. A20).

The continuous collection of the biopolymer sheet onto the rotating drum provides an avenue to

spatially organize soft materials in three dimensions. For the purpose of conveniently imaging

the architecture of produced bulk material, a two-layered hydrogel sheet was produced by adding

an additional layer (blue layer in Fig. 8b) to the microfluidic device, resulting in an 11-layer-

device. The primary biopolymer layer feed a 2%w.t. alginate solution containing a payload of

red fluorescence microbeads, while the secondary layer feed a 2%w.t. alginate solution

containing green fluorescence microbeads. Figure 8g shows a confocal scan of the produced

layered hydrogel.

We now discuss how our experimental strategy allows the various secondary biopolymers to be

incorporated within a planar, unsupported hydrogel sheet, at a spatial resolution of approximately

up to 130µm (Fig. 8e). The ability to precisely control the incorporation of the secondary

hydrogel in the lateral direction and in time allows us to consistently define a variety of

tessellations in the (x, y)-plane. Until now, the ability to controllably form heterogeneous soft

materials in a one-step process was limited to microparticles 182-184 and coded fibers 185.

Recently, a stepwise approach was developed to incorporate within a 200µm thick paper

substrate a secondary biopolymer that consisted of Matrigel containing a payload of cells 186. The

paper substrate was photolithographically patterned. Subsequently, the cell-loaded Matrigel

wicked through spaces between the cellulose fibers, in areas that were not protected by the

photoresist.

Computer-controlled solenoid valves (The Lee Company, Connecticut, US) (Fig. 8h-j and Fig.

A17) initiated the outflow of secondary biopolymers from one of the seven on-chip reservoirs

during a time period tV at which the head pressure was raised from the atmospheric pressure level

Page 54: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

43

P1 to P2. A biopolymer spot was then predictably incorporated within the hydrogel sheet and

cross-talk between different reservoirs was prevented.

3.2.3 Spatiotemporal Control and Payload Incorporation

The experimental configuration shown in Fig. 9a was used to assess the spatiotemporal control

of the process. The seven reservoirs were typically filled with a secondary biopolymer that

consisted of an alginate solution with different payloads (see Experimental section). A custom

computer interface allowed us to individually incorporate localized spots of the secondary

biopolymer on demand. In the first case we substituted the secondary biopolymer with a density-

matched aqueous solution with a composition identical to the focusing fluids (i.e., it contained

100mM CaCl2). A planar soft material sheet with an array of void areas was obtained (Fig. 9b),

at the following experimental conditions Up = 10mm/s, QB = 200µl/min, inlet pressure P =

3.5kPa, and tv = 65ms. Ultimately, the combination of void spaces with the ability to produce

millimeter thick multilayers of hydrogel sheets as illustrated in Fig. 8g, provides a strategy

towards 3D vascularized soft materials.

In a second case, we investigate the extent to which the incorporated biopolymer replaced the

base biopolymer by first considering fluorescently labelled microspheres as the payload. We

performed confocal microscopic scans and found the smallest ellipsoidal spot (lengths of semi-

principal axes: 100µm [w], 130µm [L], 130μm [δ]) that completely replaced the base hydrogel

across the entire sheet, as shown in Fig. 9c. The spot was produced with the conditions

Up = 12 mm/s, QB = 160 µl/min, inlet pressure P = 3.5kPa, and tV = 50ms.

In a third case, we selected viable cells as the payload. Consequently, the secondary biopolymer

was modified to improve cell viability and functionality, since alginate alone is insufficient to

promote cell proliferation, attachment, and migration 187,188. Briefly, a payload of neonatal rat

cardiomyocytes at a density of 10 × 106 cells/mL was suspended in a peptide-functionalized

hydrogel solution (see Methods for more detail). Confocal microscopic scans revealed a uniform

distribution of cardiomyocytes across the hydrogel sheet (Fig. 9d) in a configuration that is not

attainable in a single step using conventional top-down patterning approaches. The cell-loaded

secondary biopolymer pattern clearly extended throughout the entire cross-section (Fig. A19) of

a sheet sufficiently thin (δ~250µm) to be adequately penetrated by oxygen and nutrient

Page 55: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

44

molecules contained in the culture solution 189. The fact that δ is at least tenfold greater than the

average size of the incorporated cells renders mosaic hydrogels a candidate format for 3D cell

culture. Although the described approach can be extended up to δ ~ 700µm (Fig. A15), sheet

thicknesses δ > 250µm were not considered for cell immobilization as they would require

internal vascularization to guarantee viability of the cellular payload throughout the cross

section. Similarly, the fibroblasts were suspended in the same secondary biopolymer as the

cardiomyocytes and were incorporated as patterned spots at a cell density of 10 × 106 cells/mL.

Confocal fluorescence images of the patterned spots were obtained on Day 5, demonstrating the

cells ability to attach onto the biopolymer (Fig. 9e,f). Patterned sheets were fixed and

immunostained following the protocol described in the Experimental section.

In a fourth case, the secondary biopolymer constituted of alginate containing fluorescently

labeled diffusible molecules (concentration 100µM, molecular weights 4kDa and 40kDa FITC-

dextran, and 10kDa rhodamine-dextran, Sigma-Aldrich, Missouri, US). Spots of the secondary

biopolymer (~4 nL) were incorporated in either 2%w.t. alginate (I) or in 1%w.t. pectin-1%w.t.

alginate (II) and the diffusive release of the fluorescent marker was followed in time-sequences

of fluorescence micrographs. Figure 9g shows two fluorescence images that were taken from

such a sequence for a spot with a 40kDa FITC-dextran payload, the first one right after gelation

and the second one 3hrs later. The diffusivity of the three dextran molecules in the base

hydrogels (I) and (II) was determined by fitting similarly measured intensity distributions to an

analytical solution of the concentration field (Fig. 9h). Specifically, the diffusion coefficient for

molecular transport of dextran (4kDa, 10kDa, and 40kDa) through two different hydrogel

matrices that were composed of either 2%w.t. alginate or 1%w.t. pectin-1%w.t. alginate were

calculated by curve fitting the time-lapsed experimental data with the analytical solution for one

dimensional diffusive transport into a semi-infinite domain190.

Dt

xerfcItxI

2),( 0 (7)

We obtained the diffusivity for a best fit using the least mean squares method (LMS). The LMS

value is defined as the sum of the residuals squared,

n

i irS1

2 , where the difference between

the experimental intensity value and the value predicted by the model is ),(exp, txIIr iii . For

Page 56: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

45

all considered molecular payloads, a higher diffusivity was consistently obtained for hydrogel (I)

as compared to hydrogel (II), an effect that we attribute to the larger average pore size of

hydrogel (I) that was confirmed by scanning electron microscope (SEM) images (inset in Figs.

8e,f). As expected, the diffusivity in both base hydrogels decreased with increasing molecular

weight of the payload.

3.2.4 Information Encoding

The ability to incorporate isolated spots of a secondary biopolymer into unsupported soft

material sheets allows information to be encoded in a compact manner. The secondary

biopolymer alginate contained fluorescent microspheres as the payload. At a location

downstream of the device exit, the encoded information was continuously projected onto a line

camera using a fluorescence imaging configuration. The word “TORONTO” was patterned in

14s into a sheet of an alginate base polymer that was supplied at a flow rate QB = 160 μl/min, and

subsequently imaged (Fig. 9i). Valves were actuated at a pressure of P2 = 5 kPa with opening

and closing times of 75ms and 1000ms, respectively. The velocity of the drum was

UP = 12 mm/s. Each letter was represented by 7-20 individual spots and occupied an area of

approximately 6.25 mm2. Similarly, cardiomyocytes as a payload were pre-labelled (CellTracker

Green, Molecular Probes, Invitrogen, Canada) and predictably incorporated in multiple spots that

represented the letters “T” and “O” (Fig. 9j). The base biopolymer was 2%w.t. alginate and the

secondary biopolymer was a suspension of 10 × 106 cells/mL in the same peptide-functionalized

hydrogel as described previously for cardiomyocytes.

The density of the encoded information was increased 19 fold by employing the 7-bit American

Standard Code for Information Interchange (ASCII) where each of the seven solenoid valves was

assigned to one bit. The intensity values recorded from the formed hydrogel sheet were

interpreted by a custom computer program, translated back into text and validated against the

original text. In ASCII format, “TORONTO” was incorporated within a 37.5 mm long hydrogel

sheet during approximately 7.5 s at QB =160 μl/min and UP = 8.25 mm/s (Fig. 9k). To

demonstrate the ability of consistently writing and reading information, article 1, chapter 1 of the

UN Charter (165 words and 1,047 characters including spaces) was encoded in the same format

Page 57: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

46

(see Fig. 9l and Fig. A21). We produced a 5.6 m long sheet within 18.8 min and subsequently

validated the encoded information with 100% accuracy.

3.2.5 Geometric Control over Mosaic Hydrogel Properties

The ability to dynamically control the local material composition provides an effective means of

altering local and bulk material properties, such as the permeability and the elasticity, and of

creating soft materials with directionally dependent properties. We formed and characterized,

using confocal and wide-field fluorescence microscopy, mosaic hydrogels with a variety of

tessellations including square tiles (Fig. 10b), stripe patterns of variable width (Fig. 10c), axially

inter-connected spots (Fig. 10d) and uniform-width stripe sections (Figs. 10a, e-h). In cases

where the base hydrogel and the secondary biopolymer were chemically distinct from each other

(i.e., they did not differ by the presence or absence of a payload only), we investigated how the

different tessellations affected the bulk elastic modulus. Homogeneous and mosaic alginate

sheets were formed via cross-linking with three CaCl2 concentrations, 50, 100, 150mM, and

tensile tests were conducted (Custom 840LE2 tensile tester, Test Resources Inc., Minnesota,

USA) (Fig. 10j, see Methods for sample fixation procedure). Two homogeneous hydrogel

samples with the previously described compositions (I) and (II) were prepared, along with

mosaic hydrogels with the tessellations shown in Figs. 10d and 10f. Figure 10i summarizes the

elastic moduli that were obtained for the different crosslinker concentrations. The values

obtained for mosaic hydrogels fall in between the ones corresponding to homogeneous samples.

A comparison between the two mosaic hydrogels suggests that axially aligned tessellations

(Fig. 10f) resulted in higher elastic moduli than laterally aligned ones (Fig. 10d). All samples

exhibited an increase in the bulk elastic modulus when the crosslinker concentration increased

from 50mM to 100mM. As the concentration of CaCl2 increased, the crosslinking rate increased

proportionally. As a result, a mosaic hydrogel with a locally increased stiffness in proximity of

the sheet surface was formed, limiting the diffusion of CaCl2 into the hydrogel and thereby

creating weaker internal polymer networks 174. We associate the decrease in elastic modulus that

was observed at 150mM with this effect. Elasticity only serves as one illustration for the

acquired ability of tuning a macroscale property with a tailored microscale material composition,

and can be extended to a wide range of other material properties.

Page 58: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

47

3.2.6 Scalable Formation of 3D Hydrogel Assemblies

The obtained mosaic hydrogel sheets may be stacked-up to produce millimeter-size 3D

assemblies with a heterotypic composition. As an example of this approach, five hydrogel sheets

with the tessellations indicated in Fig. 10f were stacked in an alternating orientation (90° offset

between layers). In order to decrease absorbance of the 3D assembly during confocal imaging,

only one of the biopolymers contained a payload of fluorescence microspheres. The resulting 3D

structure had dimensions 5mm[w]×5mm[L]×1.5mm[δ] (Fig. 10j). Alternatively, the presented

platform can be employed for the tubular assembly of hydrogel sheets. As an illustration, the

rotating drum was replaced by a rotating capillary tube (22-690-943, Fisher Scientific, Canada)

which was manually translated to collect a continuously extruded hydrogel sheet with 50%

overlap in the sheet surface area. The overlap ensures the tubular architecture to be retained upon

the removal of the capillary tube. Homogeneous and heterogeneous hydrogel tubes with inner

diameters of approximately 1.5 mm and lengths of up to several centimeters were produced (Fig.

10k-n).

3.2.7 Planar Co-localization of Single and Multiple Cell Types

In a case study, we form mosaic hydrogels by locally incorporating single or multiple cell types

as a payload within the secondary biopolymer. In tissue engineering applications it is necessary

to authentically represent the physiological environmental milieu of a particular tissue or organ.

Resembling the structure and function of tissues and organs requires multiple cell types and

ECM molecules to be co-localized in two or three dimensional patterns at length scales that

exceed several millimeters. Currently available cell patterning methods allow to either

incorporate multiple cell types in microparticles and subsequently organize them in one or two

directions 191-194, or achieve co-localization along one direction within a fiber 185, but do not yet

provide dynamic control over the matrix composition and the incorporation of multiple cell types

in two or more directions.

We demonstrate continuous two-directional patterning of cardiomyocytes, endothelial cells and

fibroblasts, major components of the native heart 195,196, at a resolution of ~400µm and at length

Page 59: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

48

scales of several millimeters. The cellular payloads suspended in the biopolymer streams are

exposed to shear levels less than 2 dyne/cm2 while passing through the microfluidic device

(Fig. A18), which is within physiological ranges 197-199 and well below shear stresses of 167-

200,000 dyne/cm2 commonly associated with direct-printing26 and ink-jet printing strategies

59,63,200. Neonatal rat fibroblasts were incorporated at a concentration of 10 million cells/mL and

the conditions Up=12mm/s, QB=160µl/min, inlet pressure P=3.5kPa, and tV=65ms (Fig. 10o, top

panel). The homogeneous distribution of cells within the hydrogel sheet was assessed by z-stack

confocal scans of five spots containing cardiomyocytes pre-labelled with CellTracker Green

(Molecular Probes, Invitrogen, Canada) and incorporated at a density of 10 × 106 cells/mL. Z-

stack scans were collected with a 30µm step size to prevent cells from being counted twice. For

each spot sample, three slices located in the middle, top, and bottom, were selected and cells

were counted from these x-y plane slices within an area of 400 × 400µm2 (Fig. A19).

Cell survival of fibroblasts within 15 days of culture and neonatal rat cardiomyocytes within 7

days of culture was investigated using a Live/Dead viability/cytotoxicity kit for mammalian cells

(L3224, Invitrogen, Canada). On Day 15, fibroblasts had a 88.7% survival, and on Day 7,

neonatal rat cardiomyocytes were observed to have 93.1% (Fig. A19). The co-localization of two

cell types (cardiomyocytes or endothelial cells with fibroblasts) within separate tessellations

within a mosaic hydrogel was illustrated by patterning parallel stripes (Fig. 10h) or islands (Fig.

10o, center and bottom figures). The ability to pattern multiple cell types in close geometrical

proximity offers the potential of systematically exploring cell-cell interactions via secreted

factors as well as the interrogation of heterotypic and homeotypic cell interactions. Figure 4p

illustrates how the incorporation of different cell types can be combined with the ability to record

the associated experimental parameters in the form of a barcode that can be tracked throughout

the duration of cell culture. We co-localized patterns consisting of cardiomyocytes (green) and

fibroblasts (red) by using four on-chip reservoirs. The remaining three reservoirs were dedicated

to a 6 bit computer-readable code where a 2%w.t. alginate with a payload of fluorescence

microspheres was used as the secondary biopolymer.

We presented a flowable format for the continuous tessellation of mosaic hydrogels. We further

explored some of the key advantages of our approach that go beyond the encoding of

information and include (a) the ability to predictably, accurately and dynamically control the

Page 60: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

49

composition of a soft material in two dimensions, (b) the continuous and scalable soft material

formation at high throughput, and (c) the mask-free patterning approach. These capabilities

exceed previously demonstrated lower-dimensional heterogeneous soft materials that were

formed using an in-flow lithography technique 201-203 and lead to barcoded polymer particles with

a high information density or to coded fibers 185. The mosaic tessellations created using our

approach can be composed of chemically distinct hydrogels or may differ by the type or

concentration of the payload they carry: diffusing or binding molecules, microscale particles, or

cells. Spatiotemporal control over the hydrogel composition allowed the encoding of information

in the soft material at a resolution of approximately 130µm and 50ms.

The effects of tessellations on local and bulk material properties such as permeability to three

dextran molecules of varying sizes, as well as bulk elastic moduli of mosaic sheets were

investigated. In the cell patterning case studies, the ability to incorporate dense cell populations

within an unsupported soft material sheet with precise spatiotemporal control was demonstrated.

We selected neonatal rat cardiomyocytes to show the compatibility of our method with

patterning a primary cell type that is highly sensitive to adverse culture conditions, including

factors such as supra-physiological shear stress 204,205 and hypoxia.206 Additionally, as

cardiomyocytes are a terminally differentiated cell type that lacks the ability to proliferate, the

achieved spatiotemporal control and the compliance of our strategy with high cell densities is

critical in establishing a physiologically relevant functional tissue. The presented strategy may

enable a fully-automated and continuous format for culturing cells in physiologically relevant

microenvironments, the systematic investigation of cell-cell and cell-matrix interactions and,

ultimately, define 3D functional tissues.

Page 61: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

50

Figure 6 One-step formation of mosaic hydrogels. Schematic illustration of presented

approach. Solutions of two distinct biopolymers, with the option of preloading the second with

microparticles, biomolecular or cellular payloads, are organized into a planar uncrosslinked fluid

network using a microfluidic device (outlined with dashed lines). At the device exit, a mosaic

hydrogel with a well-defined spatial composition is formed upon cross-linking (e.g. ionic

exchange). The mosaic hydrogel properties (e.g. elasticity, diffusivity of different molecular

payloads) can be tailored by controlling its microscale composition. The ability to controllably

define planar biomaterials with heterogeneous properties enables the predictable time dependent

diffusion of molecular payloads. Bottom-up stacking or continuous collection of mosaic

hydrogel sheets onto a rotating drum enables the formation of multilayered soft materials with

compositional control in three dimensions, as well as millimeter to centimeter-scale tubular

structures.

Page 62: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

51

Figure 7. Analytical model of the time-dependent concentration of free cross-linker, free

alginate, and cross-linked gel. (a) Normalized concentration profiles of free Ca2+, free alginate,

and crosslinked gel as a function of time and position x. (b) 2D plot of the concentration profiles

as a function of position x, at five time points t = 0, 200, 400, 600, 1000s.

Page 63: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

52

Page 64: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

53

Figure 8 Continuous formation of hydrogels sheets: experimental setup and device designs.

(a) Experimental setup consisting of a microfluidic device with inlets for a base biopolymer

solution (supplied by a syringe pump) and focusing fluid (supplied by a gear pump). The

extruded soft material sheet is collected on a drum with diameter D. (b) Rendered image and (c)

photograph of device exit section. (d) Control over planar soft material thickness by varying

drum rotation speed UP, with base biopolymer flow rate QB = 80 μl/min (●), 120 μl/min (▲),

160 μl/min (■). (e, f) SEM images showing the pore structure of planar biopolymer of

homogeneous composition: (e) 2%w.t. alginate, (f) 1%w.t. pectin-1%w.t. alginate. (g) Confocal

fluorescence scan of multilayered biopolymer sheets obtained by collecting onto the rotating

drum a continuous sheet composed of two layers of biopolymers (green and red). Final bulk

dimensions of 5mm[w] × 3.5mm[δ] (thickness δ obtained by ~18 double layers 200µm thick)

with a total volume of 2 mL. (h) Rendered image and photograph (i) of multilayered microfluidic

device used for the formation of mosaic hydrogels. The device supplied biopolymer solutions 1-7

from on-chip reservoirs. Individual computer-controlled solenoid valves selected reservoir head

pressures between the states P1 and P2. (j) Schematic illustration of valve actuation. Scale bars

are 1 mm (b, c), 2 μm (e, f), 500 µm (g), 5 mm (h, i).

Page 65: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

54

Page 66: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

55

Figure 9. Dynamically encoded information in planar hydrogels. (a) Illustration of encoding

information by dynamically incorporating spots of a secondary (fluorescently labelled)

biopolymer into a base biopolymer and subsequently decoding the information contained in the

hydrogel sheet. (b) Hydrogel sheet with array of void areas as imaged by confocal fluorescence

(top) and scanning electron microscopy (bottom). (c) Confocal fluorescence image illustrating

dimensions and shape of spots created by incorporating a secondary biopolymer with a payload

of fluorescently labelled microspheres at conditions P = 3.5 kPa, QB = 160 µl/min, UP = 12 mm/s,

tv = 50 ms. Inserts represent the xy-plane (center location of sheet). d) Confocal fluorescence

image (x-z plane) of cardiomyocytes incorporated within a planar biomaterial. The planar

material of thickness δ is outlined by dashed lines. e) Confocal fluorescence image of a patterned

spot of fibroblasts incorporated at a cell density of 10 million cells/mL (40×, Day 5). f) 5x

magnification confocal fluorescence scan of fibroblasts spot shown in (e) (40×, Day 5). g)

Fluorescence image and corresponding intensity curves of 100 μM 40kDa FITC-dextran loaded

in 2%w.t. alginate and incorporated into the same base material. Images captured at times 0 and

3 hrs. h) Diffusivity of 4 kDa, 10 kDa, and 40 kDa dextran in 2%w.t. alginate ( ),

1%w.t. pectin-1%w.t. alginate ( ). i) Line camera intensity scan (top) and fluorescence image

(bottom) of encoded letters. j) Fluorescence image of pattern formed with 10 million/mL

cardiomyocytes in 1.2%w.t. alginate and in 0.08%w.t. collagen type I from rat tail (Day 0).

Approximately 25,000 cells were incorporated, operating conditions: P = 3.5 kPa,

QB = 160 µl/min, UP = 12 mm/s, valve 65 ms open. k) Fluorescence line scan of binary code

(top) and schematic of valve actuation with white sections corresponding to valve open (bottom)

(n = 7 binary characters). l) Sample fluorescence line scan of the UN charter in ASCII code (n=1,

2, ... 1047 binary characters including space). Scale bars are 500 μm (b), 150 μm (c), 200 μm

(d), 50 μm (e), 10 μm (f), 100 μm (g) and 2mm (i-k).

Page 67: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

56

Page 68: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

57

Figure 10 Mosaic hydrogels. Confocal and wide-field fluorescence images of mosaic hydrogels

with various tessellations. Two to three distinct material compositions are illustrated. Inserts

represent schematic of desired patterns including: two parallel stripes of distinct materials (a),

squares (b), alternating wave pattern (c), axially connected spots (d), and multiple parallel stripes

(e-h). Continuous inlet gas pressures ranging from 2-14 kPa were used, with a range of valve

opening times starting from 50 ms to infinity (for continuous stripe pattern). g) SEM image of

striped heterogeneous material (See Supplementary Information for SEM sample preparation). h)

Wide-field fluorescence image of two parallel stripes containing 10 million cells/mL of

cardiomyocytes (green) and fibroblasts (red). i) Modulus of elasticity of homogeneous and

heterogeneous planar materials with CaCl2 concentrations of 50, 100, and 150 mM: 2%w.t.

alginate ( ), 1%w.t. pectin-1%w.t. alginate ( ), 2%w.t. alginate with patterns of 1%w.t. pectin-

1%w.t. alginate ( ) as illustrated in (d), and ( ) in (f). j) Millimeter-scale 3D organization of

heterogeneous soft materials. Obtained by stacking, with alternating orientation, hydrogel sheets

with patterns as illustrated in insert (brightfield image) and (f). k-n) Centimeter length tubular

structures obtained by rolling onto a translating capillary tube. Confocal fluorescence images of

tubes with homogeneous (k) and heterogeneous (l,m) composition (contain fluorescent

microbeads as payload). n) Stereomicroscopy image of tubular structure. o) Single (top) and

multiple (middle and bottom) cell incorporation into a base planar material. Top and bottom

figures are fibroblasts (red) and endothelial cells (green) at a cell density of 10 million cells/mL.

Middle figure consists of fibroblasts (red) and cardiomyocytes (green) at a cell density of 2

million cells/mL. Images were captured on Day 0. p) Combination of multiple cell types

incorporation along with 6-bit barcoding of a planar material. Scale bars 500 μm (a-h, k-m,o, p),

1mm (i), 2mm (n).

.

Page 69: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

58

4 Chapter 4 - Skin Printer: Continuous and Scalable Organization of Layered Tissue

We present a continuous approach for scalable and consistent incorporation of human skin

microtissues into handleable skin substitutes. Our skin printer consists of a microfabricated

cartridge and has a throughput of up to 0.5 square meters per hour (equivalent to ~ 2×108 cells

per hour). Skin equivalents with different sizes and morphologies were bioprinted without

substrate support, and consisted of a biopolymer matrix into which regular arrays of skin

microtissues were incorporated. Bioprinted skin substitutes were consistently produced with

widths between 6mm and 60mm. Tensile properties, cell viability, attachment and proliferation

of bioprinted sheets populated with patterns containing 2 × 106 fibroblasts per ml and 9 × 106

keratinocytes per ml were characterized. The presented approach promises the scalable formation

of cell-populated skin grafts with precise control over the cell types and location, enabling the

physiological properties of the native skin to be reproduced with improved accuracy. In this

approach, the unique ability to create localized arrays of microtissues rather than homogeneously

populating the entire area of the graft with cells may reduce the need for large cell number,

thereby accelerating the post-operative culture time.

4.1 Skin Structure

Many tissues of the body including skin and muscular tissue, e.g., blood vessels and heart

muscle, possess a unique hierarchical organization that is characterized by a layered architecture.

Skin is the largest organ of the body and forms a protective barrier against its environment, and

is composed of an upper epidermal and a lower dermal layer. A unique hierarchical organization

of different cells and extracellular matrix components is critically important for the structure and

biological function of intact skin. The epidermis prevents foreign organisms from entering the

body and limits evaporative water loss. Its thickness varies between 0.05mm on the eyelids to

1.5mm on the palms and soles. It is avascular and populated by 95% keratinocytes, which form

from 37 to 51 layers depending on the region of the body.207-209 The dermis of human skin varies

Page 70: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

59

in thickness between 0.6mm, in the case of the eyelids, and 3mm, in the cases of the back, the

palms, and the soles210,211. The extracellular matrix content of the dermal layer is dominantly

comprised of collagen types I and III (> 90%) secreted by fibroblasts. Collagen is organized in

an irregular intertwined meshwork of fibers to accommodate to minor stresses during normal

activity while resisting severe stretch by aligning in rope-like structures in the direction parallel

to the tension applied207-209.

The dermis and epidermis are separated by a basal lamina, a 50-90nm thin layer that is composed

mainly of collagen Type IV, laminin, fibronectin and proteoglycans212, and prevents the

fibroblasts from the underlying dermis from direct contact with the epidermal keratinocytes,

while allowing immune cells to penetrate. On the epidermal side, the basal lamina nests a source

of self-renewing keratinocytes which, upon reaching the skin surface, stratify, shed their nuclei

and produce keratins that define the epidermal barrier. Instead of a flat boundary between the

epidermal and dermal layers, the basal lamina of intact skin exhibits periodic protrusions or

ridges (also known as rete ridges and papillary projections) that are 50-400µm wide and 50-

200µm deep 213. These ridges increase the area of contact between the dermis and epidermis by

approximately 4 to 7 times, promote a better adhesion between layers211 and enhance the

proliferation and differentiation of keratinocytes 212,214,215. The absence of a basal lamina results

in scaring and loss of biological function during wound healing 216.

The mechanical properties of skin have been investigated in vivo using a variety of approaches

including indentation 217, suction 218,219, tension 220, and torsion 221,222. The resulting values are

summarized in Table A5, and vary from one region of the body to another and are highly

dependent on the methods used (from 1.5kPa on the forearm using indentation 217 to 20MPa on

the leg using tension 220).

4.2 Skin grafts and Artificial Skin Substitutes

In severe burn injuries where both the epidermal and dermal layers are destroyed, effective

strategies that ensure prompt wound closure and result in favorable clinical outcomes and patient

survival rates are essential. Patient survival is inversely proportional to the time required for

wound coverage and stabilization, and the mortality rate increases by 10% for every additional

Page 71: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

60

10% surface area burn223. However, in cases of large burns, rapid wound closure is often

hindered by the insufficient source of autologous skin for grafting. A succession of numerous

surgical interventions are required instead224,225.

An ideal skin substitutes possesses both epidermal and dermal layers, with the respective

keratinocytes and fibroblasts incorporated within each layer. The two cell types are known to

exhibit a strong co-dependence in forming functional and vascularized skin grafts226,227. Graft re-

vascularization is critical in supplying the epidermal layer with nutrients. For grafts thicknesses

larger than 400µm, angiogenesis alone does not establish a rapid enough nutrient supply,

compromising cell viability and often resulting in graft loss228,229.

4.2.1 Skin Grafts

As of today, the golden standard of burn treatment consists in the use of split-thickness autograft,

allograft, acellular or cell-populated artificial substitute, and cell spraying strategies230-237 (Fig.

A22). Limitations associated with autografts and commercially available skin substitutes include

their limited size, the temporary nature of some of the grafts, and the long preparation time of

cell-populated grafts (over two weeks culture prior to surgery).

A split-thickness autograft relies on the use of the patient’s own skin from an unwounded area.

Therefore, the procedure results in a final wound area that is larger than the original burn.

Wound size for split-thickness grafts can be reduced by the use of a meshing technique. It allows

the surface area of the graft to be increased by up 6 times, with however a downside of reduced

cosmesis230. However, grafting is only to a limited extent applicable to severe burn patients due

to the lack of sufficient skin donor sites. The absence of a full dermis limits the extent to wich

the split-thickness autograft resembles the structure and function of the patient’s normal skin230.

Allograft consists of skin from human donors (cadaveric, Alloderm®, LifeCell Corporation, The

Woodlands, TX). It is comprised of a fully formed basement membrane. The associated

morbidity is reduced as compared to split-thickness autografts235,238. However, this strategy

comes at the risk of transmitting pathogens, e.g., the human immunodeficiency virus. Even

though the donor skin becomes vascularized upon grafting, the upper epidermal layer needs to be

removed after approximately 3 weeks to prevent the immunologic rejection by the patient235.

Page 72: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

61

Recently, micrografting techniques from the late 50s have been readapted to regenerate the

epidermis of full-thickness porcine wounds239. Due to the lack of donor sites in large burns, the

use of micrografts from autologous split-thickness skin graft is promising as it enables up to

1:100 expansion ratio, while still enabling full-thickness regeneration in porcine wounds240-242. A

total of 39 micrografts 0.8mm × 0.8mm × 0.35mm were cut using a mincing device and

manually spread over a 5cm × 5cm porcine wound bed 239. This approach was combined with

clinically available moist dressings in view of potential future clinical applications.

4.2.2 Artificial Skin Substitutes

A critical function of skin substitutes is to promote the formation of an intact epidermal barrier

that reduces transepidermal water loss, infection, and hypertrophic scarring by accelerating

wound closure and patient recovery, while enable large surface coverage without the need for

lengthy pre-operative cell culture time243. In order to best reproduce functional elements of intact

skin, biological substitutes must mimic its structure and cellular organization. Such a

requirement must rely on a technology that will enable the simultaneous control over the

combined incorporation of cells, growth factors, and biomaterials.

Boyce et al. have developed skin substitutes consisting of autologous fibroblasts and

keratinocytes cultured for two weeks on collagen sponges. Although limited in thickness

(<500µm), their approach promises qualitative outcomes comparable to meshed split-thickness

autografts, with the advantage of reduced need for donor skin harvesting244,240. Similarly, Kempf

et al. created cultured skin grafts composed of spunges made of electrospun collagen type I

fibers, onto which human fibroblasts and keratinocytes were seeded. Fibroblasts and

keratinocytes naturally separated into two distinct layers over time and resulted in a cell-laden bi-

layer structure with a typical thickness of 50-70µm, smaller than the 1-3mm typical for human

skin245.

Integra is a bilayered, commercially available skin graft consisting of type I bovine collagen and

chondroitin-6-sulfate, covered with a silicone membrane 231. However, the graft is limited by its

high cost, the two stage procedure, rejection by patients allergic to bovine products, and the

inability of cultured cells to efficiently adhere to and penetrate through the matrix 246.

Page 73: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

62

Cell therapy is another strategy that relies on cultured epidermal cells from patients or cell lines.

Cultured epithelial autografts and skin keratinocyte sheets (Epicel247) are based on keratinocytes

only. Cultured epithelial autografts is an appealing alternative as it enables expansion ratios of up

to 1000 times248, well beyond the maximum 6 times achievable using standard meshing

techniques249,250. However, culture of these cells to desirable numbers takes up to weeks, putting

further at risk the life of patients. In addition, these cellular grafts are fragile to handle and

require a supporting scaffold229,251. A feeder layer of fibroblasts is often required230. Zaulyanov

et al. investigated the use of Apligraf in the treatment of foot and leg ulcers and found it to be

equally fragile, and its incorporated neonatal keratinocytes did not remain viable when applied to

the wound bed. They found its function to be limited to stimulating wound healing through the

delivery of growth factors232,233. Fragility in handling can be overcome by employing a cell

spraying technique252 using a large number of cells (0.16 million cells/cm2), but results in limited

attachment of keratinocytes to the underlying dermis234.

4.2.3 Microfluidic approaches

The formation of spatially organized, cell-laden biomaterials has also been accomplished through

microfluidic-based approaches such as sequential replica molding or templating24,253, or by using

continuous-flow microfluidic formats that form cell-laden fibers or sheets185,254,255.

Atac et al. proposed a micro-bioreactor for the long-term dynamic culture of skin equivalents, ex

vivo skin tissue samples, and single hair follicular units 256. Similarly, Morimoto et al.257

prepared spheroid-based skin microtissues enabling the co-culture of both fibroblasts and

keratinocytes and their study as a simplified skin tissue model. The spheroids were 150µm in

diameter, composed of collagen Type I, and encapsulated human dermal fibroblasts at a density

of 1×106cells/mL. Human epidermal keratinocytes were subsequently seeded and attached to the

collagen spheroids at a density of 0.06×106 cells/cm2. Although promising for screening

purposes, microsystems-based advances have to date lacked the ability to routinely define skin

substrates larger than 1cm2 in size that closely recapitulate the structure and function of human

skin. Cerqueira et al. have remedied to the small scale limitation by utilizing a cell-sheet

technique 250, resulting in 0.8cm2 human adipose stem cell derived sheets for the treatment of

murine full-thickness excisional wounds239. Although there is potential in producing larger

Page 74: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

63

sheets 139, this approach is time consuming as it requires manual stacking of up to three

individual sheets, or the use of a support substrate to produce sheets that are sufficiently robust

and handleable for grafting. 258

4.3 3D Printer Strategy for the Continuous Formation of Skin Substitute for Wound Healing

Here, we present a skin printer for the scalable formation of cell-populated wound dressings that

accurately reproduce key features of human skin, and are scalable to grafts with large surface

areas (tens of cm2). Our approach is based on a microfabricated printer cartridge that enables the

scalable incorporation of skin microtissue regions within continuously produced hydrogel sheets.

The printed skin grafts consist of multiple viable cell types, have a precisely controlled graft

thickness, structure (e.g. epidermis and dermis), and biomolecular composition that are

characteristic of human skin. We demonstrate the provided skin substitutes are easy to routinely

handle and apply to the wound bed, and promise to reduce wound recovery times.

4.3.1 Materials and Methods

4.3.1.1 Materials

The following materials were purchased for the preparation of both the base graft material and

the cell printing solution: sodium alginate (Molecule-R), high G:M alginate (G:M >1.5,

Novamatrix), Hyaluronic acid (HA Grade 80, Novamatrix), collagen type I from rat tail (BD),

Matrigel (BD).

The base sheet material consisted of 1%w.t. sodium alginate (Molecule-R). Cells were

suspended in 3.5%v/v DI water, 0.6 %v/v glucose solution (0.3g/mL), 3.7%v/v 10× Medium 199

(Sigma-Aldrich, Canada), 4.5%v/v NaOH solution (1 N), 0.96%v/v NaHCO3 solution

(0.075g/mL), 0.4%v/v HEPES (Invitrogen, Canada), 18.7%v/v Matrigel, 2.5mg/mL collagen

type I from rat tail (BD Biosciences, Canada), 42%v/v culture media (DMEM and Epilife for

fibroblasts and keratinocytes respectively), and 8.34%v/v of cell printing alginate. The cell

printing alginate differed for each cell type to be printed. Fibroblast suspension material was

Page 75: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

64

composed of 0.05%w.t. sodium alginate (Molecule-R), whereas keratinocyte suspension material

consisted of 0.2%w.t. high G:M ratio alginate (G:M>1.5) with 0.23mg/mL hyaluronic acid (HA

Grade 150).

4.3.1.2 Printer Cartridge Fabrication

The printer cartridge consisted of four to six layers, which were fabricated using standard soft

lithography techniques139. Masters with 150µm tall features were defined, from which each

layers were individually molded and vertically attached using a partial curing process178. The

cartridge design for bilayered skin grafts consisted of six layers (Fig. A1c, left). Layers 1 and 4

distribute the flow focusing solution. Layer 5 and 6 distribute the respective fibroblast and

keratinocyte-loaded solutions into layer 2 and 3 respectively. Each layer 2 and 3 are populated

with additional channels for the distribution of a base material. This configuration results in the

formation of a bilayered skin graft containing keratinocytes on the top layer and fibroblasts on

the bottom layer. Both cartridges for single layered and 6cm wide grafts consisted of four

aligned and bonded layers. The design for the formation of a single layered graft is composed of

layers 1, 2, 4, and 5 in figure A1c (left). The cartridge design for a 60mm wide graft is

represented in figure A1c (right). Layers 1 and 3 distribute the focusing solution. Layer 4

distributes the cell-populated solution into layer 2, which guides both the base material solution

and cell-populated solution in a planar fashion until they exit the device as an organized planar

graft.

4.3.1.3 Tensile Measurements

Tensile measurements of patterned sheets were measured using a custom tensile tester (840LE2

tensile tester, Test Resources Inc., Minnesota, USA). Wet samples were cut to lengths of

approximately 20 mm and sandwhiched between two cardboard strips, which were vertically

clamped between tensile grips for testing. A ramp of 0.1 mm/s was applied using a 1000g load

cell until failure. Results are expressed as mean ± standard deviations (each sample has n = 5

experiments). All data were analysed using multiple comparison ANOVA tests in SigmaStat 3.5

(Systat Software Inc.; San Jose, CA, USA) using the Tukey method.

Page 76: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

65

4.3.1.4 Cell Sources

Fibroblasts were obtained from healthy human normal skin after surgery. Cells were cultured in

growth medium (DMEM, 10% fetal bovine serum (FBS) and 1% antibiotic/antimycotic

(Ab/Am)) until near confluency and split into further passages by trypsin-EDTA treatment

(0.05% Trypsin-EDTA). Human umbilical epidermal keratinocytes (Gibco Invitrogen, C-011-

5C) were cultured according to company instructions in EpiLife Medium with 1% HKGS and

1% Ab/Am and trypsinized using the same trypsin-EDTA solution as for fibroblasts.

4.3.1.5 Immunofluorescent Staining

Cells in sheets were fixed with 4% paraformaldehyde in HBSS for 1 hour at room temperature

then washed with HBSS. They were permeabilized with 0.5% Triton X-100 in HBSS for 30 min

at room temperature and then washed with HBSS. Cells were blocked with block buffer (1%

BSA in 0.25% Triton X-100 in HBSS) for 1 hour. Antibodies were diluted in block buffer and

incubated overnight at 4°C. Primary antibodies included fluorescein phalloidin (Life

Technologies) and cytokeratin 14 (Santa Cruz Biotechnology). If only phalloidin was used, the

mounting step was performed next. With keratinocytes, samples were washed with HBSS then

incubated with secondary Alexa Fluor antibodies (Life Technologies). Cells were washed then

slides were mounted with Vectashield mounting medium with DAPI (Vector Laboratories).

Images were taken on Apotome Axiovert fluorescent imaging system or Zeiss Observer

Z1 spinning disk confocal microscope.

4.3.1.6 Murine Model

Male nude mice (J:NU, Jackson Laboratory) at 6-8 weeks old were ordered and allowed to

acclimate to the environment for a week prior to use. Mice were anesthetized with isofluorane

and buprenorphine was given as an analgesic. Two 6mm diameter full-thickness excisional

wounds were created on the back of each mouse using a biopsy punch. Silicon rings were

sutured onto the wound to prevent contraction. Circular patches of the same diameter were

punched from the printed sheets and placed onto the wound model, followed by coverage with

Page 77: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

66

Tegaderm. The mice were monitored and health scored twice daily for 10 days. Mice were

sacrificed using CO2 euthanasia and the wound area was taken and stored in 10% neutral-

buffered formalin for histological analysis.

4.3.1.7 Histology

Tissue specimens were fixed in 10% buffered formalin overnight at 4°C, and stored in 70%

ethanol and embedded in paraffin. Specimens were cut into 5 µm sections in the centre of the

wound.

Trichrome reagents were from EMS (Hatfield, PA) unless otherwise stated. Briefly, paraffin

embedded slides were deparaffinized with citrosol, followed by rehydration through grades of

ethanol to water. Slides were placed in Bouin’s solution 1 hour at 60°C and washed in water.

Hematoxylin (Sigma) and Biebrich scarlet-acid fuchsin solution were stained for 10 minutes

each, respectively with washes in between. Slides were differentiated in phosphomolybdic-

tungstic acid for 15 min, and transferred to aniline blue for 5 min. Slides were rinsed and

differentiated in 1% acetic acid for 2 min. Slides were dehydrated through 95% ethanol and

absolute ethanol followed by clearing in citrosol. Slides were mounted with SHUR/Mount

xylene-based liquid mounting media (Triangle Biomedical Sciences). Images were acquired

using LeicaDM 2000LED light microscope.

For immunohistochemistry staining, paraffin embedded skin tissue slides were deparaffinized

with citrosol followed by rehydration. Antigen decloaker (1X, Biocare) was added to the slides

in a preheated decloaking chamber for 4 minutes at 110°C. Samples were blocked with 3% H2O2

for 10 min, then washed with washing buffer (0.05 M Tris-HCl, 0.15 M NaCl, 0.05% Tween 20

in DI water). Primary antibody was diluted in PBS and incubated at room temperature for 1 h.

Primary antibody used was cytokeratin 14 (Santa Cruz Biotechnology). Next, slides were

incubated for 15 minutes first with goat-on-rodent probe (Biocare Medical), and secondly with

goat-on-rodent HRP-polymer. The betazoid DAB chromogen kit (Biocare Medical) was added

for 5-10 min and the reaction was terminated with running water. Nuclear staining was done with

hematoxylin for 30 sec, followed by differentiation with 3 dips in 1.5% acid alcohol and bluing

in 0.1% sodium bicarbonate for 10 sec. Sections were dehydrated through 95% and absolute

Page 78: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

67

ethanol to citrosol and mounted with SHUR/Mount as previously described. Images were

acquired using LeicaDM 2000LED light microscope.

4.3.2 Formation of Cell-Populated Skin Grafts

The skin printer shown in figure 12a consists of a microfabricated cartridge, a receiving

reservoir, a stepper motor-controlled drum and a pneumatic control unit. The printer cartridge

distributes and organizes in a planar fashion the biopolymer matrix that provides mechanical

support for the printed grafts, along with the printing solution that contains a mixture of

biopolymers with dermal and epidermal cells. Three microfabricated cartridges were fabricated

for the formation of single layered, bilayered, and wide (60mm in width) skin substitutes

respectively. The biopolymer solution for the support structure consisted of 1%w.t. sodium

alginate (Molecule-R) and was supplied to the cartridge using syringe pumps. Up to two

different biopolymer-cell mixtures were supplied at the top side of the cartridge, from separate

on-chip wells. Pneumatic control of the individual head pressures along with the microfabricated

channel networks within the cartridge enabled us with precise spatio-temporal patterning of cells

within the continuously formed skin graft 255. Dedicated cartridges allowed single and bilayered

sheets that were populated with viable human fibroblasts and keratinocytes to be continuously

formed and resembled the dermal and epidermal layers of intact human skin. To reduce the

unwanted effect of flow instabilities at the device exit and to ensure a uniform sheet thickness, a

co-flowing focusing solution was delivered from above and below the soft biopolymer sheet in a

flow-focusing configuration.179 The focusing fluids carried cross-linking ions and induced

gelation of the sheet. The ability to spatially control the cellular composition of these grafts

allowed the formation of skin substitutes with defined epidermis and dermis, appropriate cell

types, and control over the cell density. The exit of the microfabricated printer cartridge was

immersed in a liquid-filled reservoir. The calcium-loaded liquid within the reservoir facilitated

cross-linking of the skin graft and guided the cross-linked graft at a distance 50 mm away from

the cartridge exit to a rotating drum (drum velocities: 1-5mm/s). Varying the velocity of drum

rotation allowed the sheet thickness to be precisely tailored. Skin grafts with thicknesses between

100µm and 500µm, and widths between 3mm and 60mm were prepared.

Page 79: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

68

Cells were incorporated in a variety of patterns within single and bilayered grafts, as alternating

spots, parallel spots, parallel void regions, and parallel stripes (Fig. 12b). The skin grafts were

subsequently collected, cultured for three days in vitro, and implanted in vivo on murine wound

models (Fig. 12c). A photograph of a microfabricated cartridge within the liquid-filled reservoir

is shown in figure 12d. Figure 12e focuses on the exit region of the cartridge, where a parallel

array of a secondary material (labelled with rhodamine for visualization) is printed into the

continuously extruded sheet.

4.3.3 Pattern Formation – No Cells

Controlled 3D cell printing and incorporation of void spaces within the hydrogel sheets was

achieved through up to two computer-controlled solenoid valves. Hydrogel solutions with a

known concentration of a particular cell type were loaded into the on-chip wells. After loading,

the wells were closed and connected to individual solenoid valves that selected between two

pressure levels. Upon electrical valve actuation, the well head-pressure either suddenly increased

or decreased, thereby initiating or ending the local incorporation of a cellular pattern within a

continuously formed hydrogel sheet. Different patterns in the form of stripes and co-localized

spots were formed, defined by the selected valve actuation times and well head-pressures (Fig.

13a). Spot volumes were precisely controlled and varied between 45-450nL (Fig. A23).

As an illustration of the patterning ability and accuracy, and to distinguish between the base

material and the patterning solution, hydrogel sheets were initially patterned with an alginate

solution (0.75%w.t. sodium alginate) while the base material was loaded with 0.5%w.t. BaSO4.

Single-layer hydrogel sheets that contained regular arrays of parallel spots, stripes, alternating

spots and void regions were continuously incorporated into the continuously extruded base

material (Fig. 13b).

4.3.4 Characterization of Printed Tissue Substitutes

An important practical requirement for skin grafts is their handleability during tissue culture and

the application to the wound bed. The elastic moduli were measured for sheets printed using

Page 80: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

69

different biopolymer solutions and patterns. Materials included homogeneous sheets of 1%w.t.

sodium alginate (Molecule-R), 4%w.t. sodium alginate (Sigma), sheets patterned with parallel

stripes, parallel spots, and alterning spots of cell printing solution, and alternating void regions.

Degradation studies over the course of 11 days were performed for all skin graft compositions

and their elastic moduli were measured during a period of four days culture in both fibroblast and

keratinocyte culture media (DMEM and Epilife, respectively) (Figs. A24, 25). Mechanical

properties of the printed skin grafts were strongly dependent on three parameters: base material,

pattern, culture media, and culture time.

For the homogeneous case, graft degradation was characterized by incubating in DMEM and

Epilife culture media over the course of 11 days. The dry weight of five samples for each time

point and condition (material type and culture media) was measured and normalized by the value

of dry weight after 2h incubation on Day 0 (Fig. A24a). Epilife culture media degraded the

materials more rapidly than in DMEM culture media, with the 1%w.t. sodium alginate

(Molecule-R) exhibiting a slower degradation than the 4%w.t. sodium alginate (Sigma). The

elastic moduli measurements shown in figure A25 confirmed our degradation results. Sheets

made of 1%w.t. sodium alginate (Molecule-R) exhibited a higher elastic modulus than the

4%w.t. sodium alginate (Sigma). Specifically, the initial elastic modulus of 4%w.t. sodium

alginate (Sigma) of 51.4 ± 4.6kPa (Day 0) was reduced to 18.8 ± 8.2kPa and 14.2 ± 7.6kPa on

Day 4 in DMEM and Epilife respectively, equivalent to a reduction by 63.4% and 72.4% of the

original value. The elastic modulus of 1%w.t. sodium alginate (Molecule-R) in DMEM however

remained constant at approximately 51kPa ± 4.4kPa during the first four days and reduced by

36.4% by day 4 during culture in Epilife. Similar trends were observed for the ultimate tensile

strength of 1%w.t. sodium alginate (Molecule-R) and 4%w.t. sodium alginate (Sigma). For the

first, the initial value of 62.5 ± 2.1kPa was reduced by 28% in DMEM and 56.8% in Epilife on

day 4. For the second, the initial ultimate tensile strength of 29.2 ± 1.9kPa dropped by 58.7% in

DMEM and 86.3% in Epiflife on day 4. We therefore selected 1%w.t. sodium alginate

(Molecule-R) as the base material of our printed skin grafts.

Although non-toxic to cells, commonly employed for cell encapsulation and 3D culture, and

FDA approved for clinical use in skin grafts, alginate alone does not provide an adequate

environment for cell attachment, proliferation, and migration. The cell printing solution was

Page 81: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

70

therefore adapted to cater to the above requirements. Briefly, two cell printing solutions were

optimized for fibroblasts and keratinocytes respectively, with a general decrease in alginate

concentration of 80-95% as compared to the base material. We compensated for the reduced

alginate concentration by selecting an alginate with a higher elastic modulus. Alginate is a linear

co-polymer composed of blocks of mannuronic (M) and gluoronic (G) residues covalently linked

in different concentrations and sequences. Alginate gelation is triggered through an ionic

mechanism. Divalent cations (Ca2+) dissolved in the liquid-filled reservoir interacted with blocks

of G residues to form a 3D gel network. A higher G:M ratio is equivalent with an increased

number of binding sites and therefore produced mechanically stronger gels.

In the case of keratinocytes, hyaluronic acide (HA) was added to the material composition. HA is

a nonsulfated glycosaminoglycan widely found in epithelial and connective tissues. It is a major

component of the ECM, contributes significantly to cell proliferation and migration, exhibits

anti-inflammatory behavior and, reduces scar formation and graft contraction by regulating

collagen synthesis85,242,259,260.

Similar to the homogeneous samples, the degradation of the cell printing solutions was

characterized by incubating in DMEM and Epilife culture media over the course of 11 days (Fig.

A24b). Overall, Epilife culture medium degraded the materials more rapidly than in DMEM

culture medium, with the fibroblast printing solution exhibiting slower degradation than

keratinocyte printing solution for both culture media.

The cell printing solution remains considerably weaker than the base material. The composition

of cell-laden printed grafts with different patterns was therefore optimized to ensure consistent

handling without rupture. The elastic moduli of four distinct patterns were investigated and

compared (Fig. 13c). Tensile tests were performed on printed sheets in both the extrusion

direction (x) and the lateral direction (y) (Fig. 13c insert), and on Day 0 and Day 1 during culture

in DMEM. Sheets patterned with alternating spots and void regions showed comparable elastic

modulus of approximately 35kPa, and exhibited satisfactory stability after one day of culture in

DMEM, with approximately 12% decrease in elastic modulus in both x and y. Patterns of parallel

spots and stripes however, showed much weaker properties when tested in the y-direction, with a

drop of 50% in elastic modulus as compared to measurements performed in the x-direction.

Similarly, the ultimate tensile strength of sheets formed with alternating spots and void regions

Page 82: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

71

was around 13kPa in x and 8kPa in y on day 0, with a decrease of approximately 15% on day 1.

On the other hand, patterns consisting of parallel stripes and spots exhibited up to 87% drop in

UTS in the y direction compared to x direction.

4.3.5 Cell-Populated Skin Grafts in Vitro

4.3.5.1 Single Layer Grafts

The presented platform to continuously pattern soft material sheets with spots and void regions

was utilized to controllably incorporate human fibroblasts and human epidermal keratinocytes

within single and bilayer skin grafts. Fibroblasts were incorporated at a concentration of

2×106cells/mL, while keratinocytes were incorporated at a concentration of 4×106cells/mL.

Figures 14a-f reveal fibroblasts printed within single layer sheets in the form of spots and

parallel stripes. The cell printing solution used consisted of 0.05%w.t. sodium alginate

(Molecule-R), 2.5mg/mL collagen type I, and 18.7%v/v Matrigel. After two days of culture in

DMEM, the printed fibroblasts were fully confluent within the three-dimensional environment of

the skin graft. The cell printing solution for keratinocytes was optimized to maximized cell

proliferation. Four candidates were tested, namely: (1) 0.05% w.t. sodium alginate (Molecule-R)

with 2.5mg/mL HA grade 80 (MW ~ 620-1200kDa), (2) 0.05% w.t. sodium alginate (Molecule-

R) with 2.5mg/mL HA grade 150 (MW ~ 1200-1900kDa), (3) 0.2% w.t. high G:M alginate

with 2.5mg/mL HA grade 80, and (4) 0.2% w.t. high G:M alginate with 2.5mg/mL HA grade

150, all of which also contained 2.5mg/mL collagen type I and 18.7%v/v Matrigel. Material

samples were prepared by pipetting 500µl of each solution into individual wells in a 24 well

plate. The solutions were gelled by gently pipetting in equal volume of alginate cross-linking

solution containing 50mM CaCl2 in DI water. The well plate was then placed in the incubator for

an additional 30min to further cross-link the collagen component of the solutions. Subsequently,

the excess calcium solution was removed and gels were gently rinsed twice with Epilife.

Keratinocytes were then seeded onto each gel sample at a concentration of 0.1million cells/cm2

and cultured over 13 days in Epilife culture media. Their proliferation was quantified by

measuring the percentage increase in surface area covered by the cells at different type points

throughout the culture period. Keratinocytes began to form clusters on Day 2. On Day 3 of

culture, samples were placed on 6-wellplate transwells and brought to the air-liquid interface.

Page 83: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

72

Compiled results in figure 15 confirmed that material 3 had the optimal composition, with a

645±28% increase in cell surface area on day 13. The insert in figure 15 shows a fluorescence

microscopy image of keratinocytes stained with Dapi, Phalloidin, and Keratin14.

4.3.6 Scalable Formation of Tissue Substitutes

The presented platform is scalable to produce skin grafts up to 10 folds wider than the smaller

grafts, equivalent to up to 60mm wide sheets. These larger scale grafts are obtained using a

microfluidic device with a 75mm wide exit section. The widths and thicknesses of skin grafts as

a function of pulling velocity obtained using four cartridges of different width were measured

and summarized in figure 17. Overall, 10mm, 25mm, 60mm, and 75mm wide devices were

characterized. The matrix flow rates used for the 10mm cartridge were: 100µl/min (green),

150µl/min (red), 200µl/min (blue). For the other three devices, flow rates were scaled

proportionally by multiplying by 2.5, 6, and 7.5 respectively. Rendered designs of 10mm and

75mm wide cartridges are represented in figure 16a. For all devices, there is on average a 35%

decrease in final sheet width compared to the initial cartridge width (Fig. 16b). We hypothesize

this phenomenon to be caused by the pulling of the alginate sheet as it exits the cartridges and

enters a stagnant liquid reservoir. This width reduction must be taken into account when

designing new cartridges for the formation of different skin grafts of specific widths. Finally,

since the matrix flow rates were scaled proportionally to the changes in cartridge widths, the

thicknesses obtained within the range of pulling velocity was consistent for all devices (Fig.

16c).

4.3.7 Cell-Populated Skin Grafts In Vivo

In vivo validation of cell-populated bilayered skin grafts was performed on 6 mm diameter

incision wounds in immune-incompetent mice. Samples were prepared and cultured in Epilife

culture media containing 1:100 BrdU for three days at 37 ̊C with 5% O2, following which they

were implanted onto murine models (fig. 17a). A representation of a bilayered skin graft was

obtained by co-localizing green fluorescent microbeads on the top layer (keratinocytes within the

epidermis) and red microbeads on the bottom layer (fibroblasts within the dermis) (fig. 17b). The

Page 84: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

73

main sheet material was loaded with blue fluorescent microbeads. All microbeads were 0.1µm in

diameter and incorporated into the solutions at a density of 1%v/v.

Control consisted of 1%w.t. sodium alginate sheets (Molecule-R) patterned with cell-free cell

printing solution, whereas cell-populated sheets were patterned with fibroblasts at a

concentration of 2×106cells/mL. Our preliminary data suggests that our printed biopolymer skin

grafts led to an improved skin regeneration compared to our control (fig. 17c-f). Specifically,

trichrome staining showed better integration and collagen formation in the cell-populated graft as

opposed to control. Similarly, Keratin 14 staining revealed keratinization post-wounding on the

cell-populated graft compared to none in the control, where keratinocytes are rather migrating

underneath the alginate graft.

We have demonstrated a scalable approach for the formation of human skin grafts that were up

to 6cm in width and meters in length. The high-throughput of this platform enabled skin grafts to

be printed at a rate of up to 0.5m2/hr, equivalent to ~2×108 cells/hr. These grafts composed of

human skin microtissue islands (2 × 106 fibroblasts per ml and 9 × 106 keratinocytes per ml)

were printed in a variety of patterns that affected the overall mechanical properties of the grafts

and determined their handleability for application onto wound beds of in vivo murine models.

Based on the tensile properties of the printed grafts, an optimal pattern design was chosen for our

in vivo experiments and consisted of microtissue islands printed in an array of alternating spots,

which resulted in the highest mechanical properties (i.e. E = 30 ± 2.1kPa and UTS = 12 ± 1.5kPa

after one day of culture in DMEM). Preliminary in vivo data obtained over 10 days post grafting

suggests improved wound healing of wound grafted with cell-populated bilayer skin substitutes

in comparison to fibroblasts-only or cell-free substitutes. This approach offers the scalable

formation of handleable skin grafts with control over the cell type and concentration, and

promises the formation of meter scale grafts with reduced requirement for cell number.

Page 85: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

74

Figure 11. Intact human skin and bioprinted skin grafts. (a) Multi-scale organization of skin,

where the epidermis is populated with keratinocytes, and the dermal layer with fibroblasts and

collagen type I being the predominant component of the ECM forming the dermis. The

epidermal and dermal layers are separated by a basement membrane. (b) Scalable, one-step

incorporation of microtissue arrays into bioprinted skin graft.

Page 86: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

75

Figure 12 Skin Printer. (a) Schematic illustration of skin printer operation for continuous

formation of cell-populated skin graft by means of a microfabricated printer cartridge. The

microfluidic device distributes base materials, with the ability to locally and controllably pattern

both keratinocytes and fibroblasts. A streaming solution is distributed on the top and bottom

layers of the printer cartridge and focuses the skin graft as it is continuously extruded into a

liquid-filled reservoir and collected onto a rotating drum. (b) A variety of patterns in the form of

(1) alternating spots, (2) parallel spots, (3) void regions, (4) and parallel stripes can be formed in

both single and bilayer skin grafts (5). (c) The skin grafts are subsequently cultured in vitro, and

Page 87: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

76

applied on murine wound models in vivo. (d) Photograph of the extrusion reservoir, with the

ability to form up to 6cm wide skin grafts. Scale bars 1cm (a), 4cm (d), 2cm (e).

Page 88: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

77

Figure 13. Patterned Single-Layered Sheets. (a) Patterns formed with cell printing solution or

void regions are printed by pressure-driven flow of various solutions loaded into on-chip

cartridges. (b) The patterns generated include: (1) parallel spots, (2) alternating spots, (3) parallel

stripes, and (4) alternating void regions. Actuation pressure constant at 3.5psi, with duty cycles

(1) 300ms open-300ms close, (2) 600ms open-600ms close, (3) fully open, (4) 600ms open-

600ms close. Scale bars 500µm. (c) Mechanical properties of skin grafts (E and UTS, *p<0.001

Day 0 and **p<0.001 Day1, n = 5). Sheets with patterns in (b) were subjected to both x and y

strain (insert, scale bar 5mm). Skin graft patterns consist of (I) base matrix without pattern, (II)

parallel spot (b-1), (III) parallel stripe (b-3), (IV) alternating void (b-4), (V) alternating spot (b-

2). Day 1 corresponds to samples cultured in DMEM during one day.

Figure 14. In vitro characterization of printed single layer skin grafts. Bright field and

fluorescence images of fibroblasts printed as spots (a-c) and stripes (d-f). (a, b, d, e) Day 3, (c, f)

Day 5. Fibroblasts were printed at a concentration of 2×106cells/mL. Actuation pressure of

3.5psi, with duty cycle 300ms open-300ms close for the spotted pattern, and fully open for the

stripe pattern. Cell staining and fluorescence imaging obtained by Cassandra Belo. Scale bars

250µm (a, d), 100µm (b, c, e, f).

Page 89: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

78

Figure 15. Material optimization for keratinocyte printing. Human keratinocytes were seeded

at a concentration of 0.1million cells/cm2 on four different material compositions, and their

proliferation assessed as a measure of increase in surface area covered by cells. Materials:M1%

w.t. sodium alginate (Molecule-R) with 2.5mg/mL HA grade 80, M1% w.t. sodium alginate

(Molecule-R) with 2.5mg/mL HA grade 150, 2.15% w.t. high G:M alginate with 2.5mg/mL

HA grade 80, and M 2.15% w.t. high G:M alginate with 2.5mg/mL HA grade 150 (statistical

comparison p<0.001 with * Day 2, ** Day 6, *** Day 9, **** Day 13; n = 5). Samples were

lifted at the air-liquid interface on Day 3. Insert: fluorescence image of keratinocytes stained

with Dapi, Phalloidin, and Keratin14. Cell staining and fluorescence imaging obtained by

Cassandra Belo. Scale bar 10µm.

Page 90: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

79

Figure 16. Scalable formation of skin grafts. (a) Rendered images of a 10mm wide (1) and

75mm wide (2) cartridge. (b, c) Measured graft widths and thicknesses versus pulling velocity

produced using four different cartridges: 10mm (●), 25mm (▲), 60mm (♦), and 75mm (■). Flow

rates 100µl/min (green), 150µl/min (red), 200µl/min (blue) for 10mm device. Flow rates for

other cartridges were scaled proportionally by multiplying by 2.5, 6, and 7.5 respectively. Scale

bars 20mm (a-1), 25mm (a-2).

Page 91: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

80

Figure 17 In vivo characterization of printed skin grafts. (a) Schematic of animal wound bed

with implanted skin graft. (b) Confocal scan of bilayered graft. Support material and cell printing

solution were loaded with fluorescent microbeads to distinguish between the different regions.

(c, d) Trichrome staining of wounded skin 10 days post-wounding. (e, f) Keratin14 staining

showing keratinization post-wounding. (c, e) Control, no cells. (d, f) 4%w.t. alginate patterned

with human fibroblasts. Histology stains obtained by Cassandra Belo. Scale bars 500µm (b, e, f),

1mm (c, d).

Page 92: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

81

Summary and Future Work

This thesis presented a novel microfluidic approach for the continuous and scalable formation of

biologically relevant soft materials with dynamically tunable composition. Directionally varying

compositions lead to directionally-dependent properties such as elastic modulus, ultimate tensile

strength, and diffusivity to various molecules. Viable cells were printed within biomaterial

sheets, promising the generation of heterogeneous soft materials with physiologically relevant

composition. The capability of the presented platform was investigated, namely: the control over

the printed material size, shape, and patterns generated, the incorporation of a variety of

materials possessing distinct mechanical properties and their impact on the macroscale properties

of the printed sheets, and viability and proliferation of various cells post-printing. The potential

application of this technology was demonstrated in two groups of work.

In a first demonstration, we have generated high aspect ratio collagen sheets with widths ranging

from 3-17mm and thicknesses ranging from 30-250µm. The formation of highly aligned collagen

fibers within these high aspect ratio sheets was possible through a combination of flow-induced

shear through a shallow geometric constriction, and strain-induced stretching in the longitudinal

direction through a collection drum. The resulting anisotropic alignment led to high mechanical

strength (E and UTS) comparable to native blood vessels. Cellular alignment and morphological

changes of smooth muscle cells and endothelial cells seeded and cultured on these aligned

collagen sheets was observed over a period of 3 days, demonstrating the direct impact of

collagen fibril alignment on cellular behavior (data obtained by our collaborator Dr. Stephanie

Grainger but not shown in this thesis). Future work will involve the application of these cell-

populated collagen sheets as tubular constructs for replacement arteries in mice and rat models.

In a second demonstration, biomaterial sheets with structural organization and cellular

composition mimicking the physiological composition of native skin were printed. Dermal

fibroblasts were incorporated at a concentration of 2-6×106cells/mL in arrays of spots and

parallel stripes and their viability and attachment was investigated. Epidermal keratinocytes were

incorporated in similar patterns at a higher concentration of 9×106 cells/mL. Material

compositions for cell printing were optimized for each cell types to promote their viability,

Page 93: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

82

attachment, and proliferation. A variety of patterns in the form of parallel stripes, parallel spots,

and alternating spots were investigated to select the optimal mechanical strength of the printed

skin grafts for improved handling. In our preliminary in vivo experiments, we have investigated

wound healing on immunodeficient murine models treated with a variety of skin grafts

consisting of single layers without cells or fibroblasts-populated spots. Histological data were

promising as wounds treated with cell-populated grafts regenerated with a full epidermal layer

populated with keratinocytes that had migrated from the host tissue.

As the next steps, we are currently investigating the difference between printed skin grafts

composed of spotted arrays of concentrated cells versus homogeneously cell-populated grafts.

The effects of cell concentration and cell clustering are investigated by comparing single layer

sheets containing: (1) fibroblasts homogeneously incorporated at a concentration of

2×106cells/mL, (2) fibroblasts incorporated at a concentration of 2×106cells/mL within spots

occupying 40% of the sheet surface area, and (3) fibroblasts incorporated at a concentration of

5×106cells/mL within spots occupying 40% of the sheet surface area (Fig. 18). We investigate

the effect of localizing cells within spots versus homogeneously distributing them within an

entire skin graft, while maintaining the same cell density (fig. 18i). Cases 1 and 3 will allow us

to study the effect of increased local cell density within spots, with the same total cell number in

the spotted candidate versus homogeneously populated candidate (Fig. 18ii). The resulting cell

viability and proliferation for all samples will be characterized through live/dead staining, where

cell concentration will be determined as a measure of cell-to-cell distance.

Finally, the co-culture of bilayered sheets containing both keratinocytes and fibroblasts

(epidermal and dermal layers respectively) will be investigated in vitro over a period of 14 to 28

days with the goal of obtaining a basement membrane between the two respective layers,

essential in the formation of functional skin substitutes. Two systems will be compared: bilayer

grafts containing fibroblasts in the bottom layer and keratinocytes in the top layer, and single

layer grafts containing a mixed solution of both keratinocytes and fibroblasts. In the first case,

fibroblasts will be incorporated at a concentration of 2×106cells/mL and keratinocytes at a

concentration of 9×106cells. In the second case, cell-populated spots will be printed with a

solution containing a mixture of 2×106cells/mL fibroblasts and 7×106cells/mL keratinocytes. The

total cell concentration to be printed must not exceed 9×106cells/mL, as higher cell concentration

would result in device clogging. These skin grafts printed with both cell types will be immersed

Page 94: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

83

in culture media and cultured over a period of 3 days, following which the grafts will be raised to

the air-liquid interface and further cultured up to 28 days.

These samples will also be grafted onto immunodeficient murine models for wound healing

studies. As a preliminary set of experiments (animal surgeries and histology data obtained by Dr.

Amini-Nik and Cassandra Belo), we have grafted bilayered skin grafts patterned with: cell-free

spots, fibroblasts only, keratinocytes only, and keratinocytes with fibroblasts. In difference to our

in vivo data in figure 17, a plastic dome designed by Dr. Amini-Nik was placed onto 8mm

diameter wound, sealing the wound boundary and preventing in the process both wound

contraction and migration of host cells. As a result, we can postulate that wound healing is

strictly due to our printed skin grafts, as healing from wound contraction and host cell invasion

was inhibited. In our preliminary data, figure 19 shows histology samples obtained by Cassandra

Belo where keratinization was observed after 15 days post-surgery. Samples were stained with

trichrome (a), Keratin 14 (b), and Keratin 10 (c).

Page 95: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

84

Figure 18. Benefit of cell clustering and characterization of skin microtissues. (i)

Comparison of homogeneously populated skin graft (1) versus locally printed cells (2). Local

cell concentration are equal between the two samples. (ii) Comparison of homogeneously

populated skin graft (1) versus patterned graft with a 2.5× higher cell density (3). Increased local

cell density in (3) vs (1), with an equal total cell number between the two candidates. Base

material (blue), cell printing solution (red), human dermal fibroblasts (yellow).

Page 96: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

85

Figure 19. In vivo characterization of bilayered cell-populated skin grafts. An 8mm diameter

dome with a ring bracket was used to segregate the wound region from the host tissue,

preventing wound contraction and migration of host cells into the wound area. Histology samples

were stained with (a) Trichrome, (b) Keratin 14, and (c) Keratin 10. Preliminary data shows the

onset of keratinocyte layer formation after 15 days post-surgery. Scale bars 500µm.

Page 97: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

86

References

1 Fratzl, P. & Weinkamer, R. Nature's hierarchical materials. Progress in Materials

Science 52, 1263-1334, doi:10.1016/j.pmatsci.2007.06.001 (2007).

2 Lakes, R. Materials with structural hierarchy. Nature 361, 511-515,

doi:10.1038/361511a0 (1993).

3 Xu, S. Q. et al. Generation of monodisperse particles by using microfluidics: Control

over size, shape, and composition. Angewandte Chemie-International Edition 44, 724-

728, doi:10.1002/anie.200462226 (2005).

4 Shui, L. L., Kooij, E. S., Wijnperle, D., van den Berg, A. & Eijkel, J. C. T. Liquid

crystallography: 3D microdroplet arrangements using microfluidics. Soft Matter 5, 2708-

2712 (2009).

5 Wu, H. K., Thalladi, V. R., Whitesides, S. H. & Whitesides, G. M. Using hierarchical

self-assembly to form three-dimensional lattices of spheres. Journal of the American

Chemical Society 124, 14495-14502 (2002).

6 van der Net, A., Delaney, G. W., Drenckhan, W., Weaire, D. & Hutzler, S. Crystalline

arrangements of microbubbles in monodisperse foams. Colloids and Surfaces 309, 117-

124 (2006).

7 Cheung, Y. K., Gillette, B. M., Zhong, M., Ramcharan, S. & Sia, S. K. Direct patterning

of composite biocompatible microstructures using microfluidics. Lab on a Chip 7, 574-

579, doi:10.1039/b700869d (2007).

8 Chan, V., Zorlutuna, P., Jeong, J. H., Kong, H. & Bashir, R. Three-dimensional

photopatterning of hydrogels using stereolithography for long-term cell encapsulation.

Lab on a Chip 10, 2062-2070, doi:10.1039/c004285d (2010).

9 Park, H. et al. The significance of pore microarchitecture in a multi-layered elastomeric

scaffold for contractile cardiac muscle constructs. Biomaterials 32, 1856-1864,

doi:10.1016/j.biomaterials.2010.11.032 (2011).

10 Dendukuri, D. & Doyle, P. S. The Synthesis and Assembly of Polymeric Microparticles

Using Microfluidics. Advanced Materials 21, 4071-4086, doi:10.1002/adma.200803386

(2009).

11 Dendukuri, D., Hatton, T. A. & Doyle, P. S. Synthesis and self-assembly of amphiphilic

polymeric microparticles. Langmuir 23, 4669-4674, doi:10.1021/la062512i (2007).

Page 98: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

87

12 Dendukuri, D., Pregibon, D. C., Collins, J., Hatton, T. A. & Doyle, P. S. Continuous-flow

lithography for high-throughput microparticle synthesis. Nature Materials 5, 365-369

(2006).

13 Panda, P. et al. Stop-flow lithography to generate cell-laden microgel particles. Lab on a

Chip 8, 1056-1061 (2008).

14 Lee, S. A., Chung, S. E., Park, W., Lee, S. H. & Kwon, S. Three-dimensional fabrication

of heterogeneous microstructures using soft membrane deformation and optofluidic

maskless lithography. Lab on a Chip 9, 1670-1675, doi:10.1039/b819999j (2009).

15 Tolley, M. T., Krishnan, M., Erickson, D. & Lipson, H. Dynamically programmable

fluidic assembly. Applied Physics Letters 93, doi:10.1063/1.3048562 (2008).

16 Krishnan, M., Tolley, M. T., Lipson, H. & Erickson, D. Hydrodynamically Tunable

Affinities for Fluidic Assembly. Langmuir 25, 3769-3774, doi:10.1021/la803517f (2009).

17 Chung, S. E., Park, W., Shin, S., Lee, S. A. & Kwon, S. Guided and fluidic self-assembly

of microstructures using railed microfluidic channels. Nature Materials 7, 581-587

(2008).

18 Yin, Y., Lu, Y., Gates, B. & Xia, Y. Template-assisted self-assembly: a practical route to

complex aggregates of monodispersed colloids with well-defined sizes, shapes, and

structures. Journal of the American Chemical Society 123, 8718-8729 (2001).

19 Matsunaga, Y. T., Morimoto, Y. & Takeuchi, S. Molding Cell Beads for Rapid

Construction of Macroscopic 3D Tissue Architecture. Advanced Materials 23, H90-H94,

doi:10.1002/adma.201004375 (2011).

20 Ballyns, J. J. et al. Image-guided tissue engineering of anatomically shaped implants via

MRI and micro-CT using injection molding. Tissue Engineering Part A 14, 1195-1202,

doi:10.1089/ten.tea.2007.0186 (2008).

21 Malda, J., Martens, D. E., Tramper, J., van Blitterswijk, C. A. & Riesle, J. Cartilage

tissue engineering: Controversy in the effect of oxygen. Critical Reviews in

Biotechnology 23, 175-194, doi:10.1080/714037688 (2003).

22 Nishiguchi, A., Yoshida, H., Matsusaki, M. & Akashi, M. Rapid Construction of Three-

Dimensional Multilayered Tissues with Endothelial Tube Networks by the Cell-

Accumulation Technique. Advanced Materials 23, 3506-+, doi:10.1002/adma.201101787

(2011).

23 Tan, W. & Desai, T. A. Layer-by-layer microfluidics for biomimetic three-dimensional

structures. Biomaterials 25, 1355-1364, doi:10.1016/j.biomaterials.2003.08.021 (2004).

24 Choi, N. W. et al. Microfluidic scaffolds for tissue engineering. Nature Materials 6, 908-

915, doi:10.1038/nmat2022 (2007).

Page 99: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

88

25 Ling, Y. et al. A cell-laden microfluidic hydrogel. Lab on a Chip 7, 756-762,

doi:10.1039/b615486g (2007).

26 Vozzi, G., Flaim, C., Ahluwalia, A. & Bhatia, S. Fabrication of PLGA scaffolds using

soft lithography and microsyringe deposition. Biomaterials 24, 2533-2540,

doi:10.1016/s0142-9612(03)00052-8 (2003).

27 Bettinger, C. J. et al. Three-dimensional microfluidic tissue-engineering scaffolds using a

flexible biodegradable polymer. Advanced Materials 18, 165-169 (2006).

28 Golden, A. P. & Tien, J. Fabrication of microfluidic hydrogels using molded gelatin as a

sacrificial element. Lab on a Chip 7, 720-725, doi:10.1039/b618409j (2007).

29 Tang, M. D., Golden, A. P. & Tien, J. Fabrication of collagen gels that contain patterned,

micrometer-scale cavities. Advanced Materials 16, 1345-+, doi:10.1002/adma.200400766

(2004).

30 Yoshida, H., Matsusaki, M. & Akashi, M. Multilayered Blood Capillary Analogs in

Biodegradable Hydrogels for In Vitro Drug Permeability Assays. Advanced Functional

Materials 23, 1736-1742, doi:10.1002/adfm.201201905 (2013).

31 Nazhat, S. N. et al. Controlled microchannelling in dense collagen scaffolds by soluble

phosphate glass fibers. Biomacromolecules 8, 543-551, doi:10.1021/bm060715f (2007).

32 Chen, C. H. et al. Novel living cell sheet harvest system composed of thermoreversible

methylcellulose hydrogels. Biomacromolecules 7, 736-743, doi:10.1021/bm0506400

(2006).

33 Haraguchi, Y. et al. Fabrication of functional three-dimensional tissues by stacking cell

sheets in vitro. Nature Protocols 7, 850-858, doi:10.1038/nprot.2012.027 (2012).

34 Masumoto, H. et al. Pluripotent Stem Cell-Engineered Cell Sheets Reassembled with

Defined Cardiovascular Populations Ameliorate Reduction in Infarct Heart Function

Through Cardiomyocyte-Mediated Neovascularization. Stem Cells 30, 1196-1205,

doi:10.1002/stem.1089 (2012).

35 Tsuda, Y. et al. The use of patterned dual thermoresponsive surfaces for the collective

recovery as co-cultured cell sheets. Biomaterials 26, 1885-1893,

doi:10.1016/j.biomaterials.2004.06.005 (2005).

36 Yang, J. et al. Cell sheet engineering: Recreating tissues without biodegradable scaffolds.

Biomaterials 26, 6415-6422, doi:10.1016/j.biomaterials.2005.04.061 (2005).

37 Yang, J. et al. Reconstruction of functional tissues with cell sheet engineering.

Biomaterials 28, 5033-5043, doi:10.1016/j.biomaterials.2007.07.052 (2007).

38 Nishida, K. et al. Functional bioengineered corneal epithelial sheet grafts from corneal

stem cells expanded ex vivo on a temperature-responsieve cell culture surface.

Transplantation 77, 379-385, doi:10.1097/01.tp.0000110320.45678.30 (2004).

Page 100: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

89

39 Shimizu, T. et al. Fabrication of pulsatile cardiac tissue grafts using a novel 3-

dimensional cell sheet manipulation technique and temperature-responsive cell culture

surfaces. Circ.Res. 90, E40-E48, doi:10.1161/hh0302.105722 (2002).

40 Kim, B. S., Choi, J. S., Kim, J. D., Choi, Y. C. & Cho, Y. W. Recellularization of

decellularized human adipose-tissue-derived extracellular matrix sheets with other human

cell types. Cell and Tissue Research 348, 559-567, doi:10.1007/s00441-012-1391-y

(2012).

41 Boland, T., Mironov, V., Gutowska, A., Roth, E. A. & Markwald, R. R. Cell and organ

printing 2: Fusion of cell aggregates in three-dimensional gels. Anatomical Record Part

a-Discoveries in Molecular Cellular and Evolutionary Biology 272A, 497-502,

doi:10.1002/ar.a.10059 (2003).

42 Boland, T., Xu, T., Damon, B. & Cui, X. Application of inkjet printing to tissue

engineering. Biotechnology Journal 1, 910-917, doi:10.1002/biot.200600081 (2006).

43 Mironov, V., Boland, T., Trusk, T., Forgacs, G. & Markwald, R. R. Organ printing:

computer-aided jet-based 3D tissue engineering. Trends in Biotechnology 21, 157-161,

doi:10.1016/s0167-7799(03)00033-7 (2003).

44 Roth, E. A. et al. Inkjet printing for high-throughput cell patterning. Biomaterials 25,

3707-3715, doi:10.1016/j.biomaterials.2003.10.052 (2004).

45 Wilson, W. C. & Boland, T. Cell and organ printing 1: Protein and cell printers.

Anatomical Record Part a-Discoveries in Molecular Cellular and Evolutionary Biology

272A, 491-496, doi:10.1002/ar.a.10057 (2003).

46 Norotte, C., Marga, F. S., Niklason, L. E. & Forgacs, G. Scaffold-free vascular tissue

engineering using bioprinting. Biomaterials 30, 5910-5917,

doi:10.1016/j.biomaterials.2009.06.034 (2009).

47 Mironov, V., Reis, N. & Derby, B. Bioprinting: A beginning. Tissue Engineering 12,

631-634, doi:10.1089/ten.2006.12.631 (2006).

48 Campbell, P. G. & Weiss, L. E. Tissue engineering with the aid of inkjet printers. Expert

Opinion on Biological Therapy 7, 1123-1127, doi:10.1517/14712598.7.8.1123 (2007).

49 Jakab, K. et al. Tissue engineering by self-assembly and bio-printing of living cells.

Biofabrication 2, doi:10.1088/1758-5082/2/2/022001 (2010).

50 Derby, B. Printing and Prototyping of Tissues and Scaffolds. Science 338, 921-926,

doi:10.1126/science.1226340 (2012).

51 Pataky, K. et al. Microdrop Printing of Hydrogel Bioinks into 3D Tissue-Like

Geometries. Advanced Materials 24, 391-+, doi:10.1002/adma.201102800 (2012).

Page 101: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

90

52 Bassoli, E., Gatto, A., Iuliano, L. & Violante, M. G. 3D printing technique applied to

rapid casting. Rapid Prototyping J. 13, 148-155, doi:10.1108/13552540710750898

(2007).

53 Ebert, J. et al. Direct Inkjet Printing of dental Prostheses Made of Zirconia. Journal of

Dental Research 88, 673-676, doi:10.1177/0022034509339988 (2009).

54 Mott, M., Song, J. H. & Evans, J. R. G. Microengineering of ceramics by direct ink-Jet

printing. Journal of the American Ceramic Society 82, 1653-1658 (1999).

55 Symes, M. D. et al. Integrated 3D-printed reactionware for chemical synthesis and

analysis. Nat. Chem. 4, 349-354, doi:10.1038/nchem.1313 (2012).

56 Di Biase, M., Saunders, R. E., Tirelli, N. & Derby, B. Inkjet printing and cell seeding

thermoreversible photocurable gel structures. Soft Matter 7, 2639-2646,

doi:10.1039/c0sm00996b (2011).

57 Klebe, R. J. Cytoscribing - A method for micropositioning cells and the construction of

two-dimensional and 3-dimensional synthetic tissues. Experimental Cell Research 179,

362-373, doi:10.1016/0014-4827(88)90275-3 (1988).

58 Wang, Y., Bokor, J. & Lee, A. in Emerging Lithographic Technologies Viii Vol. 5374

Proceedings of the Society of Photo-Optical Instrumentation Engineers (Spie) (ed R. S.

Mackay) 628-636 (2004).

59 Saunders, R. E., Gough, J. E. & Derby, B. Delivery of human fibroblast cells by

piezoelectric drop-on-demand inkjet printing. Biomaterials 29, 193-203 (2008).

60 Hockaday, L. A. et al. Rapid 3D printing of anatomically accurate and mechanically

heterogeneous aortic valve hydrogel scaffolds. Biofabrication 4, 1-22 (2012).

61 Yang, S. F., Leong, K. F., Du, Z. H. & Chua, C. K. The design of scaffolds for use in

tissue engineering. Part II. Rapid prototyping techniques. Tissue Engineering 8, 1-11,

doi:10.1089/107632702753503009 (2002).

62 Cohen, D. L., Malone, E., Lipson, H. & Bonassar, L. J. Direct freeform fabrication of

seeded hydrogels in arbitrary geometries. Tissue Engineering 12, 1325-1335 (2006).

63 Nair, K. et al. Characterization of cell viability during bioprinting processes.

Biotechnology Journal 4, 1168-1177, doi:10.1002/biot.200900004 (2009).

64 Phillippi, J. A. et al. Microenvironments engineered by inkjet bioprinting spatially direct

adult stem cells toward muscle- and bone-like subpopulations. Stem Cells 26, 127-134,

doi:10.1634/stemcells.2007-0520 (2008).

65 Cao, Y. L., Vacanti, J. P., Paige, K. T., Upton, J. & Vacanti, C. A. Transplantation of

chondrocytes utilizing a polymer-cell construct to produce tissue-engineered cartilage in

the shape of a human ear. Plastic and Reconstructive Surgery 100, 297-302 (1997).

Page 102: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

91

66 Kojima, K. et al. Comparison of tracheal and nasal chondrocytes for tissue engineering of

trachea. The Annals of Thoracic Surgery 76, 1884 (2003).

67 Kamil, S. H. et al. In vitro tissue engineering to generate a human-sized auricle and nasal

tip. Laryngoscope 113, 90 (2003).

68 Puelacher, W. C. et al. Design of nasoseptal cartilage replacements synthesized from

biodegradable polymers and chondrocytes. Biomaterials 15, 774 (1994).

69 Jakab, K. et al. Tissue engineering by self-assembly of cells printed into topologically

defined structures. Tissue Engineering Part A 14, 413-421, doi:10.1089/tea.2007.0173

(2008).

70 Mironov, V. Printing technology to produce living tissue. Expert Opinion on Biological

Therapy 3, 701-704, doi:10.1517/14712598.3.5.701 (2003).

71 Wang, N., Liu, W., Huang, J. & Ma, K. The structure-mechanical relationship of palm

vascular tissue. Journal of the mechanical behavior of biomedical materials 36, 1-11,

doi:10.1016/j.jmbbm.2014.04.001 (2014).

72 Hutmacher, D. W. Scaffold design and fabrication technologies for engineering tissues -

state of the art and future perspectives. J. Biomater. Sci.-Polym. Ed. 12, 107-124,

doi:10.1163/156856201744489 (2001).

73 Holmes, D. F. et al. Corneal collagen fibril structure in three dimensions: Structural

insights into fibril assembly, mechanical properties, and tissue organization. Proceedings

of the National Academy of Sciences of the United States of America 98, 7307-7312,

doi:10.1073/pnas.111150598 (2001).

74 Boland, E. D. et al. Electrospinning collagen and elastin: Preliminary vascular tissue

engineering. Frontiers in Bioscience 9, 1422-1432, doi:10.2741/1313 (2004).

75 Ottani, V., Raspanti, M. & Ruggeri, A. Collagen structure and functional implications.

Micron 32, 251-260, doi:10.1016/s0968-4328(00)00042-1 (2001).

76 Muller, L. J., Marfurt, C. F., Kruse, F. & Tervo, T. M. T. Corneal nerves: structure,

contents and function. Exp. Eye Res. 76, 521-542, doi:10.1016/s0014-4835(03)00050-2

(2003).

77 Ruberti, J. W., Roy, A. S. & Roberts, C. J. in Annual Review of Biomedical Engineering,

Vol 13 Vol. 13 Annual Review of Biomedical Engineering (eds M. L. Yarmush, J. S.

Duncan, & M. L. Gray) 269-295 (2011).

78 Muller, L. J., Pels, E. & Vrensen, G. The specific architecture of the anterior stroma

accounts for maintenance of corneal curvature. British Journal of Ophthalmology 85,

437-443, doi:10.1136/bjo.85.4.437 (2001).

Page 103: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

92

79 Mackenzie, I. C. & Hill, M. W. Connective-tissue influences on patterns of epithelial

architecture and keratinization in skin and oral-mucosa of the adult-mouse. Cell and

Tissue Research 235, 551-559 (1984).

80 Verhaegen, P. et al. Differences in collagen architecture between keloid, hypertrophic

scar, normotrophic scar, and normal skin: An objective histopathological analysis. Wound

Repair and Regeneration 17, 649-656, doi:10.1111/j.1524-475X.2009.00533.x (2009).

81 Chen, X., Nadiarynkh, O., Plotnikov, S. & Campagnola, P. J. Second harmonic

generation microscopy for quantitative analysis of collagen fibrillar structure. Nature

Protocols 7, 654-669, doi:10.1038/nprot.2012.009 (2012).

82 Birk, D. E. & Trelstad, R. L. Extracellular compartments in tendon morphogenesis -

collagen fibril, bundle, and macroaggregate formation. Journal of Cell Biology 103, 231-

240, doi:10.1083/jcb.103.1.231 (1986).

83 Sharma, P. & Maffulli, N. Current concepts review tendon injury and tendinopathy:

Healing and repair. Journal of Bone and Joint Surgery-American Volume 87A, 187-202,

doi:10.2106/jbjs.d.01850 (2005).

84 Gelse, K., Poschl, E. & Aigner, T. Collagens - structure, function, and biosynthesis.

Advanced Drug Delivery Reviews 55, 1531-1546, doi:10.1016/j.addr.2003.08.002 (2003).

85 Zhao, J.-Y. et al. Influence of hyaluronic acid on wound healing using composite porcine

acellular dermal matrix grafts and autologous skin in rabbits. International Wound

Journal 10, 562-572, doi:10.1111/j.1742-481X.2012.01023.x (2013).

86 Brodsky, B., Eikenberry, E. F. & Cassidy, K. Unusual collagen periodicity in skin.

Biochimica Et Biophysica Acta 621, 162-166, doi:10.1016/0005-2795(80)90072-0

(1980).

87 Hofmann, H., Fietzek, P. P. & Kuhn, K. Role of polar and hydrophobic interactions for

molecular packing of type-I collagen - 3-dimensional evaluation of amino-acid sequence.

Journal of Molecular Biology 125, 137-165, doi:10.1016/0022-2836(78)90342-x (1978).

88 Amiel, D., Frank, C., Harwood, F., Fronek, J. & Akeson, W. Tendons and ligaments: A

morphological and biochemical comparison. Journal of Orthopaedic Research 1, 257-

265 (1984).

89 Kastelic, J., Galeski, A. & Baer, E. Multicomposite structure of tendon. Connective

Tissue Research 6, 11-23, doi:10.3109/03008207809152283 (1978).

90 Diamant, J., Arridge, R. G. C., Baer, E., Litt, M. & Keller, A. Collagen-ultrastructure and

its relation to mechanical properties as a function of aging. Proceedings of the Royal

Society Series B-Biological Sciences 180, 293-+, doi:10.1098/rspb.1972.0019 (1972).

91 Misof, K., Rapp, G. & Fratzl, P. A new molecular model for collagen elasticity based on

synchrotron x-ray scattering evidence. Biophysical Journal 72, 1376-1381 (1997).

Page 104: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

93

92 Komai, Y. & Ushiki, T. The 3-dimensional organization of collagen fibrils in the human

cornea and sclera. Investigative Ophthalmology & Visual Science 32, 2244-2258 (1991).

93 Daxer, A., Misof, K., Grabner, B., Ettl, A. & Fratzl, P. Collagen fibrils in the human

corneal stroma: Structure and aging. Investigative Ophthalmology & Visual Science 39,

644-648 (1998).

94 Rafat, M. et al. PEG-stabilized carbodiimide crosslinked collagen-chitosan hydrogels for

corneal tissue engineering. Biomaterials 29, 3960-3972,

doi:10.1016/j.biomaterials.2008.06.017 (2008).

95 Wollensak, G., Spoerl, E. & Seiler, T. Stress-strain measurements of human and porcine

corneas after riboflavin-ultraviolet-A-induced cross-linking. Journal of Cataract and

Refractive Surgery 29, 1780-1785, doi:10.1016/s0886-3350(03)00407-3 (2003).

96 Bigi, A., Ripamonti, A. & Roveri, N. X-ray investigation of the orientation of collagen

fibers in aortic media layer under distending pressure. International Journal of Biological

Macromolecules 3, 287-291, doi:10.1016/0141-8130(81)90044-1 (1981).

97 van Zuijlen, P. P. et al. Collagen morphologyin human skin and scar tissue: no

adaptations in response to mechanical loading at joints. Burns 29, 423-431 (2003).

98 Purslow, P. P., Bigi, A., Ripamonti, A. & Roveri, N. Collagen fiber reorientation around

a crack in biaxially stretched aortic media. International Journal of Biological

Macromolecules 6, 21-25, doi:10.1016/0141-8130(84)90042-4 (1984).

99 Jarrett, A. The structure of collagen and elastic tissues in unprocessed skin. Br. J.

Dermatol. 70, 343-347, doi:10.1111/j.1365-2133.1958.tb13794.x (1958).

100 Beenakker, J. W. M., Ashcroft, B. A., Lindeman, J. H. N. & Oosterkamp, T. H.

Mechanical properties of the extracellular matrix of the aorta studied by enzymatic

treatments. Biophysical Journal 8, 1731-1737 (2012).

101 Berillis, P. The role of collagen in the aorta's structure. The Open Circulation and

Vascular Journal 6, 1-8 (2013).

102 Canham, P. B., Finlay, H. M. & Boughner, D. R. Contrasting structure of the saphenous

vein and internal mammary artery used as coronary bypass vessels. Cardiovascular

Research 34, 557-567, doi:10.1016/s0008-6363(97)00056-4 (1997).

103 Shadwick, R. E. Mechanical design in arteries. Journal of Experimental Biology 202,

3305-3313 (1999).

104 Achilli, M. & Mantovani, D. Tailoring mechanical properties of collagen-based scaffolds

for vascular tissue engineering: the effects of pH, temperature and ionic strength on

gelation. Polymers 2, 664-680, doi:10.3390/polym2040664 (2010).

105 Weinberg, C. B. & Bell, E. A blood vessel model constructed from collagen and cultured

vascular cells. Science 231, 397-400 (1986).

Page 105: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

94

106 L'Heureux, N., Pâquet, S., Labbé, R., Germain, L. & Auger, F. A. A completely

biological tissue-engineered human blood vessel. FASEB J. 12, 47-56 (1998).

107 Berglund, J. D., Mohseni, M. M., Nerem, R. M. & Sambanis, A. A biological hybrid

model for collagen-based tissue engineered vascular constructs. Biomaterials 24, 1241-

1254 (2003).

108 Cukierman, E., Pankov, R., Stevens, D. R. & Yamada, K. M. Taking cell-matrix

adhesions to the third dimension. Science 294, 1708-1712, doi:10.1126/science.1064829

(2001).

109 O'Shaughnessy, T. J., Lin, H. J. & Ma, W. Functional synapse formation among rat

cortical neurons grown on three-dimensional collagen gels. Neuroscience Letters 340,

169-172, doi:10.1016/s0304-3940(03)00083-1 (2003).

110 Ruggeri, Z. M. Platelets in atherothrombosis. Nature Medicine 8, 1227-1234 (2002).

111 Matsumoto, N. et al. Effect of alignment of the transplanted graft extracellular matrix on

cellular repopulation and newly synthesized collagen. Archives of Orthopaedic and

Trauma Surgery 117, 215-221, doi:10.1007/s004020050232 (1998).

112 Caves, J. M. et al. Fibrillogenesis in Continuously Spun Synthetic Collagen Fiber.

Journal of Biomedical Materials Research Part B-Applied Biomaterials 93B, 24-38,

doi:10.1002/jbm.b.31555 (2010).

113 Koester, S., Evans, H. M., Wong, J. Y. & Pfohl, T. An in situ study of collagen self-

assembly processes. Biomacromolecules 9, 199-207, doi:10.1021/bm700973t (2008).

114 Hakansson, K. M. O. et al. Hydrodynamic alignment and assembly of nanofibrils

resulting in strong cellulose filaments. Nature Communications 5,

doi:10.1038/ncomms5018 (2014).

115 Lanfer, B. et al. Aligned fibrillar collagen matrices obtained by shear flow deposition.

Biomaterials 29, 3888-3895, doi:10.1016/j.biomaterials.2008.06.016 (2008).

116 Lai, E. S., Huang, N. F., Cooke, J. P. & Fuller, G. G. Aligned nanofibrillar collagen

regulates endothelial organization and migration. Regenerative Medicine 7, 649-661,

doi:10.2217/rme.12.48 (2012).

117 Eastwood, M., Porter, R., Khan, U., McGrouther, G. & Brown, R. Quantitative analysis

of collagen gel contractile forces generated by dermal fibroblasts and the relationship to

cell morphology. Journal of Cellular Physiology 166, 33-42, doi:10.1002/(sici)1097-

4652(199601)166:1<33::aid-jcp4>3.0.co;2-h (1996).

118 Thomopoulos, S., Fomovsky, G. M. & Holmes, J. W. The development of structural and

mechanical anisotropy in fibroblast populated collagen gels. Journal of Biomechanical

Engineering-Transactions of the Asme 127, 742-750, doi:10.1115/1.1992525 (2005).

Page 106: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

95

119 Lee, P., Lin, R., Moon, J. & Lee, L. P. Microfluidic alignment of collagen fibers for in

vitro cell culture. Biomedical Microdevices 8, 35-41, doi:10.1007/s10544-006-6380-z

(2006).

120 Cheng, X. et al. An electrochemical fabrication process for the assembly of

anisotropically oriented collagen bundles. Biomaterials 29, 3278-3288,

doi:10.1016/j.biomaterials.2008.04.028 (2008).

121 Xu, B., Chow, M.-J. & Zhang, Y. Experimental and modeling study of collagen scaffolds

with the effects of crosslinking and fiber alignment. International journal of biomaterials

2011, 172389-172389, doi:10.1155/2011/172389 (2011).

122 Guo, C. & Kaufman, L. J. Flow and magnetic field induced collagen alignment.

Biomaterials 28, 1105-1114, doi:10.1016/j.biomaterials.2006.10.010 (2007).

123 Novak, T., Shannon, G., Mousoulis, C., Voytik-Harbin, S. L. & Neu, C. P. Controlled

fibrillogenesis for improved magnetic alignment of collagen. J. Tissue Eng. Regen. Med.

8, 268-269 (2014).

124 Torbet, J. & Ronziere, M. C. Magnetic alignment of collagen during self-assembly.

Biochem. J. 219, 1057-1059 (1984).

125 Barocas, V. H., Girton, T. S. & Tranquillo, R. T. Engineered alignment in media

equivalents: Magnetic prealignment and Mandrel compaction. Journal of Biomechanical

Engineering-Transactions of the Asme 120, 660-666, doi:10.1115/1.2834759 (1998).

126 Torbet, J. et al. Orthogonal scaffold of magnetically aligned collagen lamellae for corneal

stroma reconstruction. Biomaterials 28, 4268-4276,

doi:10.1016/j.biomaterials.2007.05.024 (2007).

127 Oryan, A., Moshiri, A. & Meimandi-Parizi, A. In vitro characterization of a novel tissue

engineered based hybridized nano and micro structured collagen implant and its in vivo

role on tenoinduction, tenoconduction, tenogenesis and tenointegration. Journal of

Materials Science-Materials in Medicine 25, 873-897, doi:10.1007/s10856-013-5110-3

(2014).

128 Zhong, S. P. et al. An aligned nanofibrous collagen scaffold by electrospinning and its

effects on in vitro fibroblast culture. J. Biomed. Mater. Res. Part A 79A, 456-463,

doi:10.1002/jbm.a.30870 (2006).

129 Xie, J. W., Li, X. R. & Xia, Y. N. Putting Electrospun Nanofibers to Work for

Biomedical Research. Macromol. Rapid Commun. 29, 1775-1792,

doi:10.1002/marc.200800381 (2008).

130 Builles, N. et al. Use of magnetically oriented orthogonal collagen scaffolds for hemi-

corneal reconstruction and regeneration. Biomaterials 31, 8313-8322,

doi:10.1016/j.biomaterials.2010.07.066 (2010).

Page 107: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

96

131 Namani, R., Wood, M. D., Sakiyama-Elbert, S. E. & Bayly, P. V. Anisotropic mechanical

properties of magnetically aligned fibrin gels measured by magnetic resonance

elastography. J. Biomech. 42, 2047-2053, doi:10.1016/j.jbiomech.2009.06.007 (2009).

132 Ng, C. P. & Swartz, M. A. Fibroblast alignment under interstitial fluid flow using a novel

3-D tissue culture model. American Journal of Physiology-Heart and Circulatory

Physiology 284, H1771-H1777, doi:10.1152/ajpheart.01008.2002 (2003).

133 Ng, C. P. & Swartz, M. A. Mechanisms of interstitial flow-induced remodeling of

fibroblast-collagen cultures. Annals of Biomedical Engineering 34, 446-454,

doi:10.1007/s10439-005-9067-3 (2006).

134 Dahl, S. L. M., Vaughn, M. E. & Niklason, L. E. An ultrastructural analysis of collagen

in tissue engineered arteries. Annals of Biomedical Engineering 35, 1749-1755,

doi:10.1007/s10439-007-9340-8 (2007).

135 Dahl, S. L. M., Rhim, C., Song, Y. C. & Niklason, L. E. Mechanical properties and

compositions of tissue engineered and native arteries. Annals of Biomedical Engineering

35, 348-355, doi:10.1007/s10439-006-9226-1 (2007).

136 Gervais, T., El-Ali, J., Gunther, A. & Jensen, K. F. Flow-induced deformation of shallow

microfluidic channels. Lab on a Chip 6, 500-507, doi:10.1039/b513524a (2006).

137 Caves, J. M. et al. The use of microfiber composites of elastin-like protein matrix

reinforced with synthetic collagen in the design of vascular grafts. Biomaterials 31, 7175-

7182, doi:10.1016/j.biomaterials.2010.05.014 (2010).

138 Caves, J. M. et al. Elastin-like protein matrix reinforced with collagen microfibers for

soft tissue repair. Biomaterials 32, 5371-5379, doi:10.1016/j.biomaterials.2011.04.009

(2011).

139 McDonald, J. C. et al. Fabrication of microfluidic systems in poly(dimethylsiloxane).

Electrophoresis 21, 27-40, doi:10.1002/(sici)1522-2683(20000101)21:1<27::aid-

elps27>3.3.co;2-3 (2000).

140 Cuneo, P., Magri, E., Verzola, A. & Grazi, E. Macromolecular crowding is a primary

factor in the organization of the cytoskeleton. Biochem. J. 281, 507-512 (1992).

141 Zhou, H. X., Rivas, G. & Minton, A. P. Macromolecular crowding and confinement:

biochemical, biophysical, and potential physiological consequences. Annual Review

Biophysics 37, 375-397 (2008).

142 Minton, A. P. The influence of macromolecular crowding and macromolecular

confinement on biochemical reactions in physiological media. Journal of Biological

Chemistry 276, 10577-10580 (2001).

143 Saeidi, N. et al. Molecular crowding of collagen: A pathway to produce highly-organized

collagenous structures. Biomaterials 33, 7366-7374,

doi:10.1016/j.biomaterials.2012.06.041 (2012).

Page 108: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

97

144 Cavallaro, J. F., Kemp, P. D. & Kraus, K. H. Collagen fabrics as biomaterials.

Biotechnology and Bioengineering 44, 146 (1994).

145 Kato, Y. P. & Silver, F. H. Formation of continuous collagen fibres: evaluation of

biocompatibility and mechanical properties. Biomaterials 11, 169-175 (1990).

146 Paten, J. A. et al. Utility of an optically-based, micromechanical system for printing

collagen fibers. Biomaterials 34, 2577-2587, doi:10.1016/j.biomaterials.2012.12.028

(2013).

147 Kemp, P. D., Cavallaro, J. F. & Hastings, D. N. Effects of carbodiimide crosslinking and

load environment on the remodeling of collagen scaffolds. Tissue Engineering 1, 71-79

(1995).

148 Zeugolis, D. I., Paul, R. G. & Attenburrow, G. Extruded collagen fibres for tissue-

engineering applications: influence of collagen concentration and NaCl amount. J.

Biomater. Sci.-Polym. Ed. 20, 219-234 (2009).

149 Silver, F. H. & Trelstad, R. L. Type-I collagen in solution - structure and properties of

fibril fragments. Journal of Biological Chemistry 255, 9427-9433 (1980).

150 Saeidi, N., Sander, E. A. & Ruberti, J. W. Dynamic shear-influenced collagen self-

assembly. Biomaterials 30, 6581-6592, doi:10.1016/j.biomaterials.2009.07.070 (2009).

151 Goublomme, A., Draily, B. & Crochet, M. J. Numerical prediction of extrudate swell of a

high-density polyethylene. J. Non-Newton. Fluid Mech. 44, 171-195, doi:10.1016/0377-

0257(92)80050-8 (1992).

152 Mitsoulis, E., Abdali, S. S. & Markatos, N. C. Flow simulation of herschel-bulkley fluids

through extrusion dies. Can. J. Chem. Eng. 71, 147-160 (1993).

153 Kumar, V. A. C., J. M.; Haller, C. A.; Dai, E.; Liu, L.; Grainger, S.; Chaikof, E. L.

Acellular vascular grafts generated from collagen and elastin analogs. Acta biomaterialia

9, 8067-8074 (2013).

154 Tonnesen, H. H. & Karlsen, J. Alginate in drug delivery systems. Drug Development and

Industrial Pharmacy 28, 621-630 (2002).

155 Kikuchi, A. & Okano, T. Pulsatile drug release control using hydrogels. Advanced Drug

Delivery Reviews 54, 53-77 (2002).

156 Joki, T. et al. Continuous release of endostatin from microencapsulated engineered cells

for tumor therapy. Nature biotechnology 19, 35-39 (2001).

157 Drury, J. L., Boontheeku, T. & Mooney, D. J. Cellular cross-linking of peptide modified

hydrogels. Journal of Biomechanical Engineering-Transactions of the Asme 127, 220-

228, doi:10.1118/1.1865194 (2005).

Page 109: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

98

158 Rowley, J. A., Madlambayan, G. & Mooney, D. J. Alginate hydrogels as synthetic

extracellular matrix materials. Biomaterials 20, 45-53, doi:10.1016/s0142-

9612(98)00107-0 (1999).

159 Augst, A. D., Kong, H. J. & Mooney, D. J. Alginate hydrogels as biomaterials.

Macromolecular Bioscience 6, 623-633 (2006).

160 Smidsrod, O. & Skjakbraek, G. Alginate as immobilization matrix for cells. Trends in

Biotechnology 8, 71-78, doi:10.1016/0167-7799(90)90139-o (1990).

161 Johnson, F. A., Craig, D. Q. M. & Mercer, A. D. Characterization of the block structure

and molecular weight of sodium alginates. Journal of Pharmaceutical and Pharmacology

49, 639-643 (1997).

162 Smidrod, O. & Skjakbraek, G. Alginate as immobilization matrix for cells. TIBTECH 8,

71-78 (1990).

163 Paige, K. T., Cima, L. G., Yaremchuk, M. J., Vacanti, J. P. & Vacanti, C. A. Injectable

cartilage. Plastic and Reconstructive Surgery 96, 1390-1398 (1995).

164 Chang, S. C. N. et al. Injection molding of chondrocyte/alginate constructs in the shape

of facial implants. Journal of Biomedical Materials Research 55, 503-511 (2001).

165 Stevens, M. M., Qanadilo, H. F., Langer, R. & Shastri, V. P. A rapid-curing alginate gel

system: utility in periosteum-derived cartilage tissue engineering. Biomaterials 25, 887-

894, doi:10.1016/j.biomaterials.2003.07.002 (2004).

166 Lee, K. Y., Alsberg, E. & Mooney, D. J. Degradable and injectable poly(aldehyde

guluronate) hydrogels for bone tissue engineering. J Biomed Mater Res 56, 228-233

(2001).

167 Alsberg, E., Anderson, K. W., Albeiruti, A., Franceschi, R. T. & Mooney, D. J. Cell-

interactive alginate hydrogels for bone tissue engineering. Journal of Dental Research

80, 2025-2029 (2001).

168 Okano, T. & Matsuda, T. Tissue engineered skeletal muscle: preparation of highly dense,

highly oriented hybrid skeletal muscular tissues. Cell Transplantation 7, 71-82 (1998).

169 Prabakaran, G. & Hoti, S. L. Immobilization of alginate-encapsulated Bacillus

thuringiensis var. israelensis containing different multivalent counterions for mosquito

control. Current Microbiology 57, 111-114, doi:10.1007/s00284-008-9159-z (2008).

170 Kachouie, N. N. et al. Directed assembly of cell-laden hydrogels for engineering

functional tissues. Organogenesis 6, 234-244, doi:10.4161/org6.4.12650 (2010).

171 Li, X. et al. Culture of neural stem cells in calcium alginate beads. Biotechnology

Progress 22, 1683-1689, doi:10.1021/bp060185z (2006).

Page 110: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

99

172 Skjakbraek, G., Grasdalen, H. & Smidrod, O. Inhomogeneous polysaccharide ionic gels.

Carbohydrate Polymers 10, 31-54 (1989).

173 Tanaka, H., Matsumura, M. & Veliky, I. A. Diffusion characteristics of substrates in Ca-

alginate gel beads. Biotechnology and Bioengineering 26, 53-58 (1984).

174 Kuo, C. K. & Ma, P. X. Ionically crosslinked alginate hydrogels as scaffolds for tissue

engineering: Part 1. Structure, gelation rate and mechanical properties. Biomaterials 22,

511-521, doi:10.1016/s0142-9612(00)00201-5 (2001).

175 A., M. & Elgsaeter, A. Density distribution of calcium-induced alginate gels.

Biopolymers 36, 17-41 (1995).

176 Iyer, R. K., Chui, J. & Radisic, M. Spatiotemporal tracking of cells in tissue-engineered

cardiac organoids. J. Tissue Eng. Regen. Med. 3, 196-207 (2009).

177 Plouffe, B. D., Brown, M. A., Iyer, R. K., Radisic, M. & Murthy, S. K. Controlled

capture and release of cardiac fibroblasts using peptide-functionalized alginate gels in

microfluidic channels. Lab on a Chip 9, 1507-1510 (2009).

178 Unger, M. A., Chou, H. P., Thorsen, T., Scherer, A. & Quake, S. R. Monolithic

microfabricated valves and pumps by multilayer soft lithography. Science 288, 113-116

(2000).

179 Knight, J. B., Vishwanath, A., Brody, J. P. & Austin, R. H. Hydrodynamic focusing on a

silicon chip: Mixing nanoliters in microseconds. Physical Review Letters 80, 3863-3866

(1998).

180 Anna, S. L., Bontoux, N. & Stone, H. A. Formation of dispersions using "flow focusing"

in microchannels. Applied Physics Letters 82, 364-366 (2003).

181 Fang, Y. et al. Binding behavior of calcium to polyuronates: Comparison of pectin with

alginate. Carbohydrate Polymers 72, 334-341 (2008).

182 Chen, C. H., Shah, R. K., Abate, A. R. & Weitz, D. A. Janus particles templated from

double emulsion droplets generated using microfluidics. Langmuir 25, 4320-4323 (2009).

183 Bong, K. W., Bong, K. T., Pregibon, D. C. & Doyle, P. S. Hydrodynamic focusing

lithography. Angewandte Chemie-International Edition 49, 87-90 (2010).

184 Nie, Z. H., Li, W., Seo, M., Xu, S. Q. & Kumacheva, E. Janus and ternary particles

generated by microfluidic synthesis: Design, synthesis, and self-assembly. Journal of the

American Chemical Society 128, 9408-9412 (2006).

185 Kang, E. et al. Digitally tunable physicochemical coding of material composition and

topography in continuous microfibres. Nature Materials 10, 877-883 (2011).

186 Derda, R. et al. Multizone paper platform for 3D cell cultures. Plos One 6 (2011).

Page 111: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

100

187 Nagakura, T. et al. Effect of viscous injectable pure alginate sol on cultured fibroblasts.

Plastic and Reconstructive Surgery 116, 831-838 (2005).

188 Pokrywczynska, M., Drewa, T., Jundzill, A. & Lysik, J. Alginate is not a good material

for growth of rapidly proliferating cells. Transplantation Proceedings 40, 1664-1667

(2008).

189 Radisic, M. et al. Oxygen gradients correlate with cell density and cell viability in

engineered cardiac tissue. Biotechnol Bioeng 93, 332-343 (2006).

190 Bird, R. B., Stewart, W. E. & Lightfoot, E. N. Transport Phenomena. Second edn, 18

(Wiley, 2007).

191 Du, Y., Lo, E., Ali, S. & Khademhosseini, A. Directed assembly of cell-laden microgels

for fabrication of 3D tissue constructs. Proc Natl Acad Sci U S A 105, 9522-9527 (2008).

192 Bruzewicz, D. A., McGuigan, A. P. & Whitesides, G. M. Fabrication of a modular tissue

construct in a microfluidic chip. Lab Chip 8, 663-671 (2008).

193 Qi, H. et al. Patterned differentiation of individual embryoid bodies in spatially organized

3D hybrid microgels. Advanced Materials 22, 5276-5281 (2010).

194 Fernandez, J. G. & Khademhosseini, A. Micro-masonry: construction of 3D structures by

microscale self-assembly. Advanced Materials 22, 2538-2541 (2010).

195 Naito, H. et al. Optimizing engineered heart tissue for therapeutic applications as

surrogate heart muscle. Circulation 114, I72-78 (2006).

196 Brown, M. A., Iyer, R. K. & Radisic, M. Pulsatile perfusion bioreactor for cardiac tissue

engineering. Biotechnol Prog 24, 907-920 (2008).

197 Davies, P. F. Flow-mediated endothelial mechanotransduction. Physiological Reviews 75,

519-560 (1995).

198 Dewey, C. F., Bussolari, S. R., Gimbrone, M. A. & Davies, P. F. The dynamic response

of vascular endothelial cells to fluid shear stress. J. Biomech. Eng.-Trans. ASME 103,

177-185 (1981).

199 Born, C., Zhang, Z., Alrubeai, M. & Thomas, C. R. Estimation of disruption of animal-

cells by laminar shear-stress. Biotechnology and Bioengineering 40, 1004-1010 (1992).

200 Lee, W. et al. On-demand three-dimensional freeform fabrication of multi-layered

hydrogel scaffold with fluidic channels. Biotechnology and Bioengineering 105, 1178-

1186 (2010).

201 Pregibon, D. C., Toner, M. & Doyle, P. S. Multifunctional encoded particles for high-

throughput biomolecule analysis. Science 315, 1393-1396 (2007).

Page 112: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

101

202 Chapin, S. C., Pregibon, D. C. & Doyle, P. S. High-throughput flow alignment of

barcoded hydrogel microparticles. Lab on a Chip 9, 3100-3109 (2009).

203 Lee, H., Kim, J., Kim, H. & Kwon, S. Colour-barcoded magnetic microparticles for

multiplexed bioassays. Nature Materials 9, 745-749 (2010).

204 Carrier, R. L. et al. Control of the structure and metabolism of engineered cardiac muscle

by direct perfusion of culture medium. Abstracts of Papers American Chemical Society

219, 76 (2000).

205 Carrier, R. L. et al. Perfusion improves tissue architecture of engineered cardiac muscle.

Tissue Engineering 8, 175-188 (2002).

206 Casey, T. M. & Arthur, P. G. Hibernation in noncontracting mammalian cardiomyocytes.

Circulation 102, 3124-3129 (2000).

207 Ribeiro, J. F., Martins dos Anjos, E. H., Mello, M. L. S. & Vidal, B. d. C. Skin Collagen

Fiber Molecular Order: A Pattern of Distributional Fiber Orientation as Assessed by

Optical Anisotropy and Image Analysis. Plos One 8, doi:10.1371/journal.pone.0054724

(2013).

208 Gibson, T., Kenedi, R. M. & Craik, J. E. Mobile micro-architecture of dermal collagen -

A bio-engineering study. British Journal of Surgery 52, 764-&,

doi:10.1002/bjs.1800521017 (1965).

209 Sakai, S., Yamanari, M., Lim, Y., Nakagawa, N. & Yasuno, Y. In vivo evaluation of

human skin anisotropy by polarization-sensitive optical coherence tomography.

Biomedical Optics Express 2, 2623-2631 (2011).

210 Ross, M. H. P., W. in Integumentary System 442-475 (2003).

211 Kanitakis, J. Anatomy, histology and immunohistochemistry of normal human skin.

European Journal of Dermatology 12, 390-400 (2002).

212 Pins, G. D., Toner, M. & Morgan, J. R. Microfabrication of an analog of the basal

lamina: biocompatible membranes with complex topographies. Faseb Journal 14, 593-

602 (2000).

213 Odland, G. F. The morphology of the attachment between the dermis and the epidermis.

Anatomical Record 108, 399-413, doi:10.1002/ar.1091080305 (1950).

214 Downing, B. R., Cornwell, K., Toner, M. & Pins, G. D. The influence of microtextured

basal lamina analog topography on keratinocyte function and epidermal organization.

Journal of Biomedical Materials Research Part A 72A, 47-56, doi:10.1002/jbm.a.30210

(2005).

215 Bush, K. A. & Pins, G. D. Development of Microfabricated Dermal Epidermal

Regenerative Matrices to Evaluate the Role of Cellular Microenvironments on Epidermal

Page 113: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

102

Morphogenesis. Tissue Engineering Part A 18, 2343-2353,

doi:10.1089/ten.tea.2011.0479 (2012).

216 Vracko, R. Basal lamina scaffold-anatomy and significance for maintenance of orderly

tissue structure. American Journal of Pathology 77, 314-346 (1974).

217 Bader, D. L. & Bowker, P. Mechanical characteristics of skin and underlying tissues in

vivo. Biomaterials 4, 305-308, doi:10.1016/0142-9612(83)90033-9 (1983).

218 Hendriks, F. M. et al. A numerical-experimental method to characterize the non-linear

mechanical behaviour of human skin. Skin Research and Technology 9, 274-283,

doi:10.1034/j.1600-0846.2003.00019.x (2003).

219 Diridollou, S. et al. In vivo model of the mechanical properties of the human skin under

suction. Skin Research and Technology 6, 214-221, doi:10.1034/j.1600-

0846.2000.006004214.x (2000).

220 Manschot, J. F. M. & Brakkee, A. J. M. The measurement and modeling of the

mechanical properties of human skin in vivo. 1. The measurement. J. Biomech. 19, 511-

515, doi:10.1016/0021-9290(86)90124-7 (1986).

221 Escoffier, C. et al. Age-related mechanical-properties of human-skin - An in vivo study.

Journal of Investigative Dermatology 93, 353-357, doi:10.1111/1523-1747.ep12280259

(1989).

222 Agache, P. G., Monneur, C., Leveque, J. L. & Derigal, J. Mechanical properties and

Young's modulus of human skin in vivo. Archives of Dermatological Research 269, 221-

232, doi:10.1007/bf00406415 (1980).

223 Miller, S. F. et al. National burn repository 2007 report: A synopsis of the 2007 call for

data. Journal of Burn Care & Research 29, 862-870,

doi:10.1097/BCR.0b013e31818cb046 (2008).

224 Tanner, J. C., Jr., Vandeput, J. & Olley, J. F. The mesh skin graft. Plastic and

Reconstructive Surgery 34, 287-292 (1964).

225 Warden, G. D., Saffle, J. R. & Kravitz, M. A 2-stage technique for excision and grafting

of burn wounds. Journal of Trauma-Injury Infection and Critical Care 22, 98-103,

doi:10.1097/00005373-198202000-00004 (1982).

226 Delvoye, P. et al. Fibroblasts induce the assembly of the macromolecules of the

basement-membrane. Journal of Investigative Dermatology 90, 276-282,

doi:10.1111/1523-1747.ep12456042 (1988).

227 Demarchez, M., Hartmann, D. J., Regnier, M. & Asselineau, D. The role of fibroblasts in

dermal vascularization and remodeling of reconstructed human skin after transplantation

onto the nude-mouse. Transplantation 54, 317-326, doi:10.1097/00007890-199208000-

00023 (1992).

Page 114: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

103

228 Sahota, P. S. et al. Development of a reconstructed human skin model for angiogenesis.

Wound Repair and Regeneration 11, 275-284, doi:10.1046/j.1524-475X.2003.11407.x

(2003).

229 MacNeil, S. Progress and opportunities for tissue-engineered skin. Nature 445, 874-880,

doi:10.1038/nature05664 (2007).

230 Sheridan, R. Closure of the excised burn wound: autografts, semipermanent skin

substitutes, and permanent skin substitutes. Clinics in Plastic Surgery 36, 643-+,

doi:10.1016/j.cps.2009.05.010 (2009).

231 Burke, J. F., Yannas, I. V., Quinby, W. C., Bondoc, C. C. & Jung, W. K. Successful use

of a physiologically acceptable skin in the treatment of extensive burn injury. Annals of

Surgery 194, 413-428, doi:10.1097/00000658-198110000-00005 (1981).

232 Streit, M. & Braathen, L. R. Apligraf - a living human skin equivalent for the treatment

of chronic wounds. International Journal of Artificial Organs 23, 831-833 (2000).

233 Zaulyanov, L. & Kirsner, R. S. A review of a bi-layered living cell treatment (Apligraf

(R)) in the treatment of venous leg ulcers and diabetic foot ulcers. Clinical Interventions

in Aging 2, 93-98, doi:10.2147/ciia.2007.2.1.93 (2007).

234 Wood, F. M., Stoner, M. L., Fowler, B. V. & Fear, M. W. The use of a non-cultured

autologous cell suspension and Integra (R) dermal regeneration template to repair full-

thickness skin wounds in a porcine model: A one-step process. Burns 33, 693-700,

doi:10.1016/j.burns.2006.10.388 (2007).

235 Cuono, C., Langdon, R. & McGuire, J. Use of cultured epidermal autografts and dermal

allografts as skin replacement after burn injury. Lancet 1, 1123-1124 (1986).

236 Jones, I., Currie, L. & Martin, R. A guide to biological skin substitutes. British Journal of

Plastic Surgery 55, 185-193, doi:10.1054/bjps.2002.3800 (2002).

237 Boyce, S. T. et al. Surface electrical capacitance as a noninvasive index of epidermal

barrier in cultured skin substitutes in athymic mice. Journal of Investigative Dermatology

107, 82-87, doi:10.1111/1523-1747.ep12298286 (1996).

238 Boyce, S. T., Supp, A. P., Swope, V. B. & Warden, G. D. Vitamin C regulates

keratinocyte viability, epidermal barrier, and basement membrane in vitro, and reduces

wound contraction after grafting of cultured skin substitutes. Journal of Investigative

Dermatology 118, 565-572, doi:10.1046/j.1523-1747.2002.01717.x (2002).

239 Hackl, F. et al. Moist dressing coverage supports proliferation and migration of

transplanted skin micrografts in full-thickness porcine wounds. Burns 40, 274-280,

doi:10.1016/j.burns.2013.06.002 (2014).

240 Hackl, F. et al. Epidermal Regeneration by Micrograft Transplantation with Immediate

100-Fold Expansion. Plastic and Reconstructive Surgery 129, 443E-452E,

doi:10.1097/PRS.0b013e318241289c (2012).

Page 115: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

104

241 Breuing, K., Eriksson, E., Liu, P. & Miller, D. R. Healing of partial thickness porcine

skin wounds in a liquid environment. J. Surg. Res. 52, 50-58, doi:10.1016/0022-

4804(92)90278-8 (1992).

242 Svensjo, T., Pomahac, B., Yao, F., Slama, J. & Eriksson, E. Accelerated healing of full-

thickness skin wounds in a wet environment. Plastic and Reconstructive Surgery 106,

602-612, doi:10.1097/00006534-200009030-00012 (2000).

243 Klingenberg, J. M. et al. Engineered human skin substitutes undergo large-scale genomic

reprogramming and normal skin-like maturation after transplantation to athymic mice.

Journal of Investigative Dermatology 130, 587-601, doi:10.1038/jid.2009.295 (2010).

244 Boyce, S. T. et al. Cultured skin substitutes reduce requirements for harvesting of skin

autograft for closure of excised, full-thickness burns. Journal of Trauma-Injury Infection

and Critical Care 60, 821-829, doi:10.1097/01.ta.0000196802.91829.cc (2006).

245 Kempf, M. et al. A denatured collagen microfiber scaffold seeded with human fibroblasts

and keratinocytes for skin grafting. Biomaterials 32, 4782-4792,

doi:10.1016/j.biomaterials.2011.03.023 (2011).

246 Boyce, S. T. et al. Comparative-assessment of cultured skin substitutes and native skin

autograft for treatment of full-thickness burns. Annals of Surgery 222, 743-752,

doi:10.1097/00000658-199512000-00008 (1995).

247 Wright, K. A. et al. Alternative delivery of keratinocytes using a polyurethane membrane

and the implications for its use in the treatment of full-thickness burn injury. Burns 24, 7-

17, doi:10.1016/s0305-4179(97)00075-2 (1998).

248 Atiyeh, B. S. & Costagliola, M. Cultured epithelial autograft (CEA) in burn treatment:

Three decades later. Burns 33, 405-413, doi:10.1016/j.burns.2006.11.002 (2007).

249 Wood, F. M., Kolybaba, M. L. & Allen, P. The use of cultured epithelial autograft in the

treatment of major burn injuries: A critical review of the literature. Burns 32, 395-401,

doi:10.1016/j.burns.2006.01.008 (2006).

250 Lumenta, D. B., Kamolz, L.-P. & Frey, M. Adult Burn Patients With More Than 60%

TBSA Involved-Meek and Other Techniques to Overcome Restricted Skin Harvest

Availability-The Viennese Concept. Journal of Burn Care & Research 30, 231-242,

doi:10.1097/BCR.0b013e318198a2d6 (2009).

251 Yannas, I. V., Burke, J. F., Orgill, D. P. & Skrabut, E. M. Wound tissue can utilize a

polymeric template to synthesize a functional extension of skin. Science 215, 174-176,

doi:10.1126/science.7031899 (1982).

252 Navarro, F. A. et al. Sprayed keratinocyte suspensions accelerate epidermal coverage in a

porcine microwound model. Journal of Burn Care & Rehabilitation 21, 513-518,

doi:10.1067/mbc.2000.110291 (2000).

Page 116: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

105

253 Baranski, J. D. et al. Geometric control of vascular networks to enhance engineered

tissue integration and function. Proceedings of the National Academy of Sciences of the

United States of America 110, 7586-7591, doi:10.1073/pnas.1217796110 (2013).

254 Onoe, H. et al. Metre-long cell-laden microfibres exhibit tissue morphologies and

functions. Nature Materials 12, 584-590, doi:10.1038/nmat3606 (2013).

255 Boyce, S. T. et al. Skin wound closure in athymic mice with cultured human-cells,

biopolymers, and growth-factors. Surgery 110, 866-876 (1991).

256 Atac, B. et al. Skin and hair on-a-chip: in vitro skin models versus ex vivo tissue

maintenance with dynamic perfusion. Lab on a Chip 13, 3555-3561,

doi:10.1039/c3lc50227a (2013).

257 Morimoto, Y., Tanaka, R. & Takeuchi, S. Construction of 3D, layered skin, microsized

tissues by using cell beads for cellular function analysis. Advanced Healthcare Materials

2, 261-265, doi:10.1002/adhm.201200189 (2013).

258 Pereira Gomes, J. A. et al. Corneal Reconstruction with Tissue-Engineered Cell Sheets

Composed of Human Immature Dental Pulp Stem Cells. Investigative Ophthalmology &

Visual Science 51, 1408-1414, doi:10.1167/iovs.09-4029 (2010).

259 Galassi, G. et al. In vitro reconstructed dermis implanted in human wounds: degradation

studies of the HA-based supporting scaffold. Biomaterials 21, 2183-2191,

doi:10.1016/s0142-9612(00)00147-2 (2000).

260 Zacchi, V. et al. In vitro engineering of human skin-like tissue. Journal of Biomedical

Materials Research 40, 187-194, doi:10.1002/(sici)1097-4636(199805)40:2<187::aid-

jbm3>3.0.co;2-h (1998).

261 Cicchi, R. et al. From molecular structure to tissue architecture: collagen organization

probed by SHG microscopy. Journal of Biophotonics 6, 129-142,

doi:10.1002/jbio.201200092 (2013).

262 Chan, E.-S. et al. Effect of formulation of alginate beads on their mechanical behavior

and stiffness. Particuology 9, 228-234, doi:10.1016/j.partic.2010.12.002 (2011).

263 Dewey, C. F., Bussolari, S. R., Gimbrone, M. A. & Davies, P. F. The dynamic-response

of vascular endothelial-cells to fluid shear-stress. Journal of Biomechanical Engineering-

Transactions of the Asme 103, 177-185 (1981).

264 Young, E. W. K., Wheeler, A. R. & Simmons, C. A. Matrix-dependent adhesion of

vascular and valvular endothelial cells in microfluidic channels. Lab on a Chip 7, 1759-

1766, doi:10.1039/b712486d (2007).

265 Hendriks, F. M., Brokken, D., Oomens, C. W. J., Bader, D. L. & Baaijens, F. P. T. The

relative contributions of different skin layers to the mechanical behavior of human skin in

vivo using suction experiments. Medical Engineering & Physics 28, 259-266,

doi:10.1016/j.medengphy.2005.07.001 (2006).

Page 117: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

106

Page 118: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

107

Appendix

Microfluidic Device Fabrication

The multi-layered microfluidic devices were produced from vertically stacked and bonded

PDMS layers that were individually obtained by moulding from different masters. Figure A1

represents rendered views of the microfluidic device designs presented.

Masters with 150µm tall features were defined by spincoating negative photoresist SU8-2050

(MicroChem Corp, Newton, MA, USA) onto clean glass substrate. The final feature height was

achieved by two spincoating steps at 1900rpm (30s with a 5s linear ramp to 1900rpm), producing

a 75µm thick resist layer in each step. This two-step procedure ensured thickness uniformity

across the entire master. After the first spincoating step, the substrate was postbaked for 6min at

a temperature of 65˚C, followed by 15min at 95˚C. Following the second spincoating step, the

substrate was baked for 10min at 65˚C and 35 min at 95˚C. Features with minimum width of

230 µm at the device exit section were patterned by soft lithography with 24mW/cm2 UV

intensity and 9s exposure time (total energy of 220 mJ). The exposed substrate was baked for

30 s at 65˚C and 20 min at 95˚C, left to cool to room temperature, and developed under constant

shaking for 12 min with SU8 developer (MicroChem Corp).

Individual layers of the microfluidic device were defined by spincoating178

poly(dimethylsiloxane) (1:10 ratio of curing agent to monomer) (PDMS, Sylgard 184 Silicone

Elastomer Kit, Dow Corning, Midland, MI, USA). Spincoating PDMS at 450 rpm for 30 s

resulted in layers with uniform thickness of 400 ±7 µm. The multilayer device was obtained by

sequentially aligning and bonding individual layers that were previously partially cured for 8 min

at 80ºC, producing a final multilayer device composed of up to 10 layers. On-chip reservoirs

were obtained from 3ml BD syringe barrels cut in half, resulting in a total fluid storage volume

of 1.5ml. The section of the barrel containing the female luer lock connector was used for easy

connection to the computer-controlled solenoid valves using male luer lock connectors

(Upchurch Scientific, Oak Harbor, Washington, USA). These reservoirs were implemented onto

the microfluidic device by first fixing with epoxy and subsequently pouring a 1cm thick uncured

PDMS layer over the final device, preventing the reservoirs from delaminating. The completed

multilayer microfluidic device was further cured for 8 hrs at 80ºC. Devices consistently

withstood inlet pressures up to 600 kPa without any delamination.

Page 119: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

108

Page 120: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

109

Figure A1. Rendered device designs. (a) Designs for the formation of collagen sheets. Focusing

solution layer (left) and collagen layer (right). (b) Designs for the formation of patterned single

layered sheet with the ability to incorporate up to seven distinct payloads or material solutions.

(c) Designs for the formation of bilayered skin grafts 15mm wide (left) and single layered skin

grafts 60mm wide (right). On the left, the layer numbers correspond to: (1, 4) focusing solution,

(2) bottom matrix layer, (3), top matrix layer, (5) distribution of botton cell patterning solution,

(6) distribution of top cell patterning solution. On the right, the layer numbers correspond to: (1,

3) focusing solution, (2) matrix layer, (4) distribution of cell patterning solution. Scale bars 5mm

(a), 10mm (b), 20mm (c).

CHAPTER 2 – Appendix

Tissues Fiber Molecule Fibril Diameter

[nm] Fiber Diameter [µm]

UTS

[MPa]

E

[MPa]

Palm Tree73,261 -- -- 1.1-19 89-222 440-1090

Spruce Wood1,73 Cellulose -- 34.5 ± 5.2 -- 14000 ± 2000

Rat Tail Skin 86 Collagen -- -- 1-20 --

Human

Cornea 92-95 Collagen 25-35 (stroma) -- 3.3 ± 0.2 15.9 ± 2.0

Rat Tail

Tendon 1,86 Collagen 50-500 -- 100 1000-2000

Native Porcine

Arteries 134,135 Collagen 16-500 -- 6.58 ± 0.97 45.1 ± 16.8

Table A1. Natural tissues and their mechanical properties.

Page 121: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

110

Collagen Formation Degree of

Alignment

Dimensions of

Collagen Construct

[W]×[L] [mm]

Collagen Fibril

Diameter [nm]

UTS

[MPa]

E

[MPa]

Ionic Strength, pH,

Temperature 104 Random 0.01 0.1

Microfluidic – Flow 119 20-40% aligned

Flow/Extrusion 112 aligned 53 ± 14 × 21 ± 3 94 ± 19 775 ± 173

Flow/Extrusion 116 aligned 1 × 1.2 × 30 45 - 3.6 ± 1.9

Electrochemical 120 aligned 0.050-0.4 × 30-70 24-88 277-671

Electrospun and

embedded 127

71.27 ± 12.81%

aligned 3 × 20 272 ± 183 2.69 ± 0.47 0.044 ± 0.0042

Table A2. Mechanical and structural properties of natural and synthetic collagen gels created

using a variety of strategies.

Figure A2. Characterization of velocity profiles within the constriction. Velocity streamlines

of focusing solution loaded with 0.08% v/v fluorescent microbeads were imaged using long-

exposure fluorescence microscopy. (a) Entrance region to the constriction. QM = 100µl/min,

QF = 1ml/min, V* = 10. (b) No collagen flow and no pulling. QF = 3ml/min (1), 6ml/min (2).

Scale bars 250µm.

Page 122: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

111

Numerical simulation of the velocity profile within the constriction.

A multiphysics solver based on the finite element method (Comsol v3.5 fluid dynamics module)

was used to model the velocity profile of the focusing solution within the constriction, under

steady flow conditions. The system was treated as a two-dimensional problem, with the focusing

solution consisting of 10% w/v PEG (dynamic viscosity = 54.4cP) treated as a Newtonian fluid.

The collagen sheet was assumed a gelled sheet and modelled as a moving wall with a velocity

equal to the pulling velocity VP = 2mm/s. The focusing solution flow rate QF was varied from 1-

16ml/min and input as a constant mean velocity, which was calculated by dividing the flow rate

by the cross-sectional area of the microfluidic device (10mm [W] × 0.12mm [H]). The outlet was

set at 0Pa gauge pressure. The parameters used in the numerical study are listed in table A3, and

a schematic of the model and the resulting velocity profiles are summarized in figure A3.

Focusing Flow Rate QF [mL/min] 1-16

Pulling Velocity VP [mm/s] 2

Dynamic Viscosity Focusing Solution [Pa•s] 0.0544

Density Focusing Solution [mg/mL] 100

½ Constriction Height [mm] 0.5

Table A3. Parameters used in numerical model, treating the focusing solution as a Newtonian

fluid.

Page 123: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

112

Figure A3. Numerical simulation of the flow behavior of the focusing solution within ½ of

the constriction (½ Hc). The cross-linked sheet is considered a moving wall with velocity equal

to the pulling velocity VP = 2mm/s. (a) Schematic of the region investigated. (b) Velocity profiles

of the focusing solution within half the constriction height HC, with QF = 1, 3, 6, 9, 12ml/min.

Focusing solution viscosity = 54.4cP.

Figure A4. Characterization of collagen sheet width and thickness produced at varying V*

and QF, with constant QM. QM = 100µl/min, QF = 1ml/min (red), QF = 3ml/min (blue), QF =

6ml/min (green). Data obtained for 10mm wide device with flow-focusing manifold, HC= 1mm.

Page 124: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

113

Figure A5. Mechanical properties of collagen sheets following various treatments: with/without

48hr incubation in fibril incubation buffer (FIB) post-extrusion, and with/without air-drying step

post-extrusion. Young’s modulus and upper tensile strength are reported for all conditions.

Figure A6. (a) TEM images of collagen sheets produced at V* = 0.6 (1) and 4.5 (2). (b) SEM

images of collagen sheets produced at V* = 0.6 (1) and 4.5 (2). Magnification: 20× (left), 100×

(right). Scale bars 200nm (a), 1µm (b, left), 500nm (b, right).

Page 125: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

114

Figure A7. Characterization of fibril spacing and compaction by autocorrelation of TEM

(a) and SEM (b) images. Experimental conditions are QM = 100µl/min, QF = 1ml/min, V* = 0.1,

0.6, 7, 10 (1 to 4 respectively). (c) SEM image of collagen sheet produced with QM = 100µl/min,

QF = 1ml/min, and V* = 7, showing the repeated banding pattern (D-period). Scale bar 500nm.

Page 126: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

115

Figure A8. Photograph of microfluidic device with constriction unit. Scale bar 10mm.

Page 127: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

116

Design Drawings of the Constriction Manifold

Figure A9. Constriction manifold assembly.

Figure A10. Constriction manifold – Top piece.

Page 128: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

117

Figure A11. Constriction manifold – Bottom piece.

Design Drawings of the XZ Constriction Manifold

Figure A12. XZ constriction manifold assembly.

Page 129: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

118

Figure A13. XZ constriction manifold - Frame.

Figure A14. XZ constriction manifold – Left constriction bracket.

Page 130: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

119

Figure A15. XZ constriction manifold – Right constriction bracket.

Page 131: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

120

CHAPTER 3 – Appendix

Referenc

e

Alginate

Concentrati

on

[% w.t.]

Alginate

M:G

Ratio

Cross-

Linker

Cross-Linker

Concentration

[mM]

Cross-

Linking

Time

[min]

Elastic

Modulus

[kPa]

UTS

[kPa]

Drury et

al.157

2 0.29 CaSO4 1540 20 52 ± 12 32 ± 3

2 0.85 CaSO4 1540 20 14 ± 5 5 ± 2

Chan et

al.262

2 0.59 CaCl2 135.2 30 300 ± 45 --

2 1.56 CaCl2 135.2 30 250 ± 10 --

4.5 0.59 CaCl2 135.2 30 570 ± 15 --

4.5 1.56 CaCl2 135.2 30 400 ± 30 --

1.5 0.59 BaCl2 50 30 592 --

1.5 0.59 CuCl2 50 30 802 --

Stevens et

al. 165

2 0.33-0.54 CaCl2 75 -- 172 ± 7 --

2 0.33-0.54 CaCl2 300 -- 155 ± 3 --

4 0.33-0.54 CaCl2 75 -- 489 ± 21 --

4 0.33-0.54 CaCl2 300 -- 471 ± 19 --

Table A4 Mechanical properties of alginate gels.

Page 132: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

121

Control of Soft Material Thickness by Fluid Focusing

Planar soft material samples with well-defined thicknesses were also produced by employing the

outlined strategy while relying exclusively on the shear imposed by the focusing fluid, i.e.,

without additional pulling via the collection drum. Thickness of samples produced at QB =

120µl/min and QF = 2-10ml/min was characterized. Thicknesses ranging from 170-700µm were

obtained and measured by optical microscopy of the cross section.

Figure A16. Control over soft material thickness as a function of the flow rate of the focusing

stream QF, and for QB = 120µl/min

Page 133: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

122

Dynamics of Valve Actuation for Soft Material Coding

Characterization of the dynamic behaviour of the computer-actuated solenoid valves (model

LHLA0521111H, The Lee Company, Westbrook, CT, USA) was achieved using piezoresistive

pressure transducers (pressure range: 0-30psi, time resolution: 1ms, model HSCDIP030PGAA5,

Honeywell, Morristown, New Jersey, USA). On-chip measurements obtained in the reservoirs

during valve actuation in terms of a voltage were converted to a pressure reading using a

calibration curve. Two actuation cycles were considered: 0.15 s open and 2 s closed, and 0.25 s

open and 2 s closed. The measured pressures and valve actuation times were found to be in good

agreement with the programmed input parameters.

Figure A17. Characterization of pressure in on-chip reservoirs. Valve actuation pattern: (left)

0.15 ms open – 2 s close, (right) 0.25 ms open – 2s close. Input pressure 7kPa. Inserts represent

magnified view of pressure evolution during valve actuation.

Page 134: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

123

Shear Stress During Cell Patterning

Shear stresses during the flow of cell suspension into the microfluidic device were calculated to

ensure that the shear stress experienced by the cells did not exceed physiological levels. Given

the employed microfluidic channel geometry and experimental conditions, the inlet pressures

(wells) of 2-4 kPa and a viscosity of the (uncrosslinked) biopolymers of approximately 0.05 Pas,

the shear stress is linearly distributed between the location of the channel center (zero) and its

maximum value of 13-26 dyne/cm2 at the wall (Poiseuille flow). The cell suspended within the

biopolymer are therefore subjected to shear stresses less than 2 dyne/cm2, a level well within

physiological conditions. Endothelial cells, e.g., experience 15-20 dyne/cm2 in undisturbed

regions of the vascular system, can be transiently exposed to 40-50 dyne/cm2 in areas of

disturbed flow 197,263 and exhibit reduced adhesion above 100 dyne/cm2 264. The calcium

chloride concentrations of 50-100mM that we used for cross-linking of hydrogel sheets are

consistent with conditions previously employed for cell encapsulation and are not detrimental to

cells 171,177,185.

Figure A18. Shear stress profile within a microfluidic channel.

Page 135: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

124

Figure A19. Viability and distribution of printed cells. (a) Statistics of cell distribution within

a single pattern (n=5). (b) Confocal fluorescence image of fibroblasts incorporated in 2%w.t.

alginate (Day 5). (c) Assessment of patterned cell survival (n=5). Neonatal rat fibroblast (left)

and cardiomyocyte (right). Scale bars 200µm (a), 50µm (b).

Page 136: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

125

Figure A20. Modulus of elasticity for a homogeneous soft material composed of 2%w.t.

alginate produced in the free-extrusion and pulled-extrusion modes.

Page 137: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

126

Figure A21. Line camera intensity measurements of the UN Charter, Chapter 1, Article1, “The

purposes of the United Nations”.

Page 138: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

127

CHAPTER 4 – Appendix

Method Test Region E [kPa]

Indentation [217]

Male thigh 1.99 ± 0.43

Male forearm 1.51 ± 0.32

Female forearm 1.09 ± 0.54

Suction [219,265] Forearm 56-217

Tensile [220]

Leg (parallel to muscle) 20000

Leg (perpendicular to muscle) 4600

Torsion [222, 221]

Dorsal forearm (<30yr old) 420

Dorsal forearm (>30yr old) 850

Ventral forearm 1120

Table A5. Literature data of elastic moduli of human skin measured in vivo using various

strategies.

Page 139: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

128

Figure A22. Full-thickness burns and current treatments. (a) Pathophysiology of skin and

comparison to full-thickness burns, (b) Clinical approaches for treatment of full-thickness burns.

Figure A23. Characterization of spot volume as a function of valve actuation.

Page 140: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

129

Figure A24. Degradation studies of skin grafts. (a) Grafts composed of base material solutions

consisting of 1%w.t. sodium alginate (Molecule-R) (circle) and 4%w.t. sodium alginate (Sigma)

(triangle). (b) Grafts composed of cell printing solutions consisting of 0.05%w.t. sodium alginate

(Molecule-R) (circle), and 0.2%w.t. high G:M alginate with 0.23mg/mL Hyaluronic acid

(triangle). Samples were incubated at 37 ̊C in DMEM (full symbols) and Epilife (hollow

symbols), and their dry weight measured over a period of 11 days.

Page 141: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

130

Figure A25. Mechanical properties of skin grafts as a function of culture time. Elastic

modulus and UTS of: 1%w.t. sodium alginate (Molecule-R) and 4%w.t. sodium alginate (Sigma)

as base materials. Samples were cultured in DMEM (*) and Epilife (**) over four days.

Page 142: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

131

Sample Stress-Strain Curves of Printed Skin Grafts

Figure A26. Stress-Strain curves of skin graft materials as a function of culture time. (a)

1%w.t. sodium alginate (Molecule-R) and (b) 4%w.t. sodium alginate (Sigma). Samples were

cultured in DMEM and Epilife over four days.

Page 143: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

132

Figure A27. Stress-Strain curves of skin graft printed with parallel spots. Tensile tests

conducted in both x and y direction on day 0 and day 1 of culture in DMEM.

Page 144: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

133

Figure A28. Stress-Strain curves of skin graft printed with parallel stripes. Tensile tests

conducted in both x and y direction on day 0 and day 1 of culture in DMEM.

Page 145: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

134

Figure A29. Stress-Strain curves of skin graft printed with alternating voids. Tensile tests

conducted in both x and y direction on day 0 and day 1 of culture in DMEM.

Page 146: Microfluidic Approach for the Scalable Formation of 3D Soft ......We have developed a microfluidic-based printing platform for the continuous and scalable formation of planar soft

135

Figure A30. Stress-Strain curves of skin graft printed with alternating spots. Tensile tests

conducted in both x and y direction on day 0 and day 1 of culture in DMEM.