154
Université de Sherbrooke Faculté de génie Département de génie chimique DEVELOPMENT OF BIOACTIVE, POROUS SCAFFOLDS FOR THE VASCULARIZATION OF ENGINEERED TISSUES DÉVELOPPEMENT DÉCHAFAUDAGES POREUX, BIOACTIFS POUR LA VASCULARISATION DE TISSUS ARTIFICIELS Thesis submitted for the degree of Doctor of Philosophy Chemical Engineering __________________________________________ Hanna BRAMFELDT Sherbrooke, Québec, CANADA February, 2008

Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

Université de Sherbrooke

Faculté de génie Département de génie chimique

DEVELOPMENT OF BIOACTIVE, POROUS SCAFFOLDS FOR THE VASCULARIZATION OF

ENGINEERED TISSUES

DÉVELOPPEMENT D’ÉCHAFAUDAGES POREUX, BIOACTIFS

POUR LA VASCULARISATION DE TISSUS ARTIFICIELS

Thesis submitted for the degree of Doctor of Philosophy Chemical Engineering

__________________________________________

Hanna BRAMFELDT

Sherbrooke, Québec, CANADA February, 2008

Page 2: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

I

ABSTRACT

The lack of vascular networks to support the growth, function, and survival of three-

dimensional tissues represents one of the greatest challenges of tissue engineering today. This thesis

project was designed to constitute the first part of a work whose final objective is the creation of

functional micro-vessel networks within three-dimensional scaffold constructs for tissue engineering.

Specifically, this part was aimed at creating functional scaffolds, that may be used in the continued

studies towards achieving this long-term goal.

As a first step, three co-polymers on the form poly(ε-caprolactone-co-D,L- lactic acid)-

poly(ethylene glycol)-poly(ε-caprolactone-co-D,L- lactic acid) were synthesized to provide

mechanically stable biomaterials with controllable and tunable material properties. To achieve this, the

monomer content in the side chains was varied and the resulting co-polymers were characterized using 1H-NMR, GPC, DSC, WAXS, and DMA1. Furthermore, the hydrolytic degradation profiles of samples

fabricated from melt-pressed films were studied over six months, as well as the influence of

degradation on tensile strength. Based on obtained results and observations, two of the co-polymers

were judged suitable, in terms of mechanical integrity and degradation profile, for further

investigation.

To improve cell/polymer interactions, co-polymer films were modified by protein

immobilization using imidoester chemistry. The effect of several protein preparations on smooth

muscle cell (SMC) adhesion and proliferation was studied over six days in culture. Cells were counted

at pre-set intervals, and further analyzed by scanning electron microscopy and fluorescent labelling of

the differentiation marker α-actin. Immobilized proteins greatly enhanced adhering cell numbers,

although for stimulation of SMC proliferation covalent immobilization was superior to physisorption.

1 1H-NMR: 1H-nuclear magnetic resonance; GPC: gel permeation chromatography; DSC: differential scanning calorimetry; WAXS: wide-angle x-ray spectrometry; and DMA: dynamic-mechanical analysis

Page 3: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

II

RÉSUMÉ La construction d‘un réseau vasculaire pouvant assurer la croissance, la fonction et la survie de

tissus tridimensionnels représente actuellement l’un des défis les plus importants du génie tissulaire.

Cette thèse fait partie d’un vaste projet ayant pour but de développer des méthodes de culture de

micro-vaisseaux sanguins au sein d’un échafaudage tridimensionnel synthétique qui pourra servir en

génie tissulaire en conditions de bioréacteur.

Trois copolymères composés de poly(ε-caprolactone-co-D,L-acide lactique)-PEG-poly(ε-

caprolactone-co-D,L-acide lactique) ont été synthétisés par la polymérisation par ouverture de cycle. La

composition monomérique a été variée et les polymères obtenus ont été caractérisés par 1H-résonance

magnétique nucléaire, chromatographie de perméation de gel, calorimétrie différentielle à balayage,

diffraction des rayons X aux grands angles, et analyseur mécanique dynamique. De plus, la

dégradation hydrolytique des échantillons, fabriqués par moulage à l’état fondu, a été étudiée, ainsi que

l’impact d’une telle dégradation sur l’endurance à la tension mécanique de ces matériaux. Considérant

les résultats observés au niveau des propriétés mécaniques et leur évolution au cours des tests de

dégradation, deux des copolymères ont été sélectionnés pour être étudiés davantage en lien avec leur

possible utilisation en cultures cellulaire et tissulaire.

Dans l’objectif d’améliorer les interactions cellules/matériaux, des films composés de ces deux

copolymères ont été modifiés en immobilisant des protéines à leur surface. L’effet de plusieurs

protéines sur le comportement de cellules musculaires lisses (smooth muscle cells, SMC) a été étudié

au cours de cultures allant jusqu’à six jours. Les cellules ont été comptées à des intervalles fixes, et

caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

marqueur de différentiation, α-actine. La présence de protéines a amélioré de façon significative les

interactions cellulaires, notamment au niveau de la prolifération qui était davantage stimulée par

l’utilisation de protéines immobilisées par liaison covalente.

Une technique basée sur des mélanges polymériques co-continus a été appliquée à nos

copolymères. Les échafaudages fabriqués par cette technique démontraient une porosité hautement

interconnectée, dont les tailles de pores peuvent être contrôlées par recuit. La perméabilité et la

résistance à la compression des structures tridimensionnelles ont été déterminées. De plus, l‘adhésion

de SMC à des échafaudages modifiés par protéines a indiqué une dépendance des tailles de pores.

Les résultats de cette thèse démontrent que les échafaudages fabriqués et modifiés tel que

spécifié, pourraient trouver usage dans l’ingénierie de tissus à base de SMC. Ces cultures pourraient

aussi servir à la vascularisation de tissus artificiels par voie de co-culture avec des cellules

endothéliales.

Page 4: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

III

GENERAL INTRODUCTION

The in vitro generation of human tissues for transplantation constitutes a long-standing dream

in reparative and reconstructive medicine. Today, the regeneration of tissues such as

cartilage[1], bone[2], and skin[3] is possible with available techniques. However, engineering

(TE) of thick, metabolically demanding tissues is facing an important challenge, as the lack of

functional vascular networks, able to support a growing cell mass, has proven to significantly

limit tissue development[4]. Such tissues, including the pancreas, liver, heart, and kidney,

require intrinsic blood supply for long-term survival and development[5]. To date, the focus of

research has been placed mainly on promoting ingrowth from the pre-existing blood vessels at

the implant site into the TE construct[6]. However, less attention has been given to the

regeneration of viable tissue around pre-constructed vascular networks.

OBJECTIVES. This project was designed to constitute the first part of a work whose final

objective will be the creation of functional vascular beds within three-dimensional scaffold

constructs for tissue engineering. The main objective of the thesis was the development of

three-dimensional, bioactive, porous scaffolds that may be used for advanced tissue culture in

bioreactors. To this end, we wanted to:

1. Synthesize a group of biocompatible and modifiable co-polymers, with

controllable degradation rates and mechanical properties. Characterize and evaluate

these co-polymers for their applicability as scaffold materials.

2. Apply a biochemical surface-modification method pertaining to the optimization of

interactions between the synthesized biomaterials and cells native to blood vessels,

which may eventually serve to improve formation of mature vessel-like structures.

Smooth muscle cells, an essential component of mature vasculature[7,8] was

chosen for this purpose.

3. Fabricate bioactive, three-dimensional, porous scaffolds with highly interconnected

porosity that may withstand long-term culture conditions in a perfusion bioreactor.

To employ the surface modification technique developed to improve cell adhesion

in a three-dimensional environment.

Page 5: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

IV

THESIS STRUCTURE. This thesis was written in a manner that substitutes the classical division

into chapters by the scientific publications that are the results of this project. Some of these

articles have been accepted for publication, whereas others are submitted for review. The

details of the publications and their current status are available in the List of Publications.

Below follows a brief description of their content.

Chapter 1, “Scaffold vascularization: a challenge for three-dimensional tissue

engineering”, consists of a literature review of advancements in three-dimensional tissue

engineering with an emphasis on in vitro tissue culture, and the vascularization of porous

scaffolds. Moreover, the effects of co-culture techniques as well as mechanical conditioning

of tissue constructs within the confines of a bioreactor are discussed. Following this

introductory chapter, the articles resulting from the work carried out within the framework of

this project are presented in Chapters 2-5.

In Chapter 2, “Characterisation, degradation and mechanical strength of poly(D,L

lactide-co-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-co-ε-caprolactone)”, the

detailed characterization of three synthesized co-polymers with varying D,L-lactide and ε-

caprolactone content is described. These co-polymers were chosen due to their

biocompatibility, and the ease with which their material properties can be altered to answer to

the diverse set of demands placed on scaffold materials. The polymers were characterized

using 1H-nuclear magnetic resonance (1H-NMR), gel permeation chromatography (GPC),

differential scanning calorimetry (DSC), wide-angle x-ray spectrometry (WAXS), and

dynamic-mechanical analysis (DMA). The hydrolytic degradation of co-polymer films as well

as the influence of degradation on the tensile strength of these co-polymers is reported. In

addition, blends of these copolymers were prepared and the hydrolytic degradation profiles of

films of different composition and thickness were compared. The results of this study can be

seen in Chapter 3, “Blends as a strategy towards tailored hydrolytic degradation of P(CL-co-

D,L-LA)-PEG-P(CL-co-D,L-LA) copolymers ”.

Chapter 4, “Enhanced smooth muscle cell adhesion and proliferation on protein-

modified polycaprolactone-based copolymers”, describes a two-dimensional cell-culture

study carried out using films from two of the co-polymers that were judged most suitable for

scaffold fabrication. In an effort to create bioactive materials, primary amine groups were

introduced through aminolysis, followed by treatment with dimethyl pimelimidate. The thus

Page 6: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

V

activated surfaces were subsequently reacted with several proteins (fibronectin, fibrinogen,

and fibrin layers), which were covalently immobilized. Co-polymer films exhibiting

physisorbed proteins were also included in the study. The effects of this surface modification

on smooth muscle cell adhesion and proliferation were investigated.

In the last phase of this work, three-dimensional scaffolds were fabricated using a

binary polymer blend technique, which resulted in highly interconnected pores and allowed

close control of pore size through static annealing. With the aim of evaluating the applicability

of such scaffolds in tissue engineering, the compressive strength and permeability of non-

annealed and annealed structures were investigated. Moreover, scaffolds of two different pore

sizes were fibronectin-modified using the method used in Chapter 4, and compared to evaluate

cell adhesion and seeding efficiency of smooth muscle cells. The results are presented and

discussed in Chapter 5, “Smooth muscle cell adhesion in surface-modified three-dimensional

co-polymer scaffolds prepared from co-continuous blends”.

CONTRIBUTIONS. During the course of this thesis, I have had important help from several

people, to whom I am very grateful. The co-author of three of the publications making up this

document, Dr Pierre Sarazin, has assisted in the execution and interpretation of results from

DSC, WAXS, and DMA. The DMA analysis itself was carried out by Dr Patrice Cousin,

Department of Civil Engineering, at the Université de Sherbrooke. Pierre Sarazin has

furthermore prepared all the binary polymer blends used for the fabrication of porous

scaffolds, and taught me the techniques to carry out static annealing, and preparation of

samples for SEM analysis. Finally, he contributed the idea to prepare co-polymer blend films

to study the influence of such compositions and of film thickness on degradation.

The testing of compressive mechanical properties was carried out at Industrial

Materials Institute in Boucherville, QC, where Dr Martin Bureau and Manon Plourde assisted

in the use of the equipment. WAXS and 1H-NMR analysis were carried out by the designated

technicians. All other laboratory and and interpretive work was carried out by myself.

Page 7: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

VI

References

(1) Stock U, Vacanti JP. Tissue engineering: current state and prospects. Annu. Rev. Med. 2001; 52:443-51.

(2) Middleton JC, Tipton AJ. Synthetic biodegradable polymers as orthopedic devices. Biomaterials 2000; 21:2335-2346.

(3) Balasubramani M, Kumar TR, Babu M. Skin substitutes: a review. Burns 2001; 27:534-544.

(4) Kannan RY, Salacinski HJ, Sales K, Butler P, Seifalian AM. The roles of tissue engineering and vascularisation in the development of micro-vascular networks: a review. Biomaterials 2005; 26:1857-1875.

(5) Borenstein JT, Terai H, King KR, Weinberg EJ, Kaazempur-Mofrad MR, Vacanti JP. Microfabrication technology for vascularized tissue engineering. Biomed. Microdevices 2002; 4:167-175.

(6) Cassell OCS, Hofer S, Morrison WA, Knight KR. Vascularization of tissue-engineered grafts: the regulation of angiogenesis in reconstructive surgery and in disease states. Brit. J. Plastic Surg. 2002; 55:603-610.

(7) Drake CJ, Hungerford JE, Little CD. Morphogenesis of the first blood vessels. Morphogenesis: Cell. Interactions 1998; 857:155-179.

(8) Vailhe B, Vittet D, Feige JJ. In vitro models of vasculogenesis and angiogenesis. Lab. Invest. 2001; 81:439-452.

Page 8: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

VII

LIST OF PUBLICATIONS

The thesis contains modified versions of the below publications.

• Bramfeldt, H., Vermette, P. Scaffold vascularization: A challenge for three-dimensional

tissue engineering. In preparation for Frontiers in Biomedicine.

• Bramfeldt, H., Sarazin, P., and Vermette, P. Characterisation, degradation and

mechanical strength of poly(D,L-lactide-co-ε-caprolactone)-poly(ethylene glycol)-

poly(D,L-lactide-co-ε-caprolactone). Journal of Biomedical Materials, Part A, 2007;

83A : 503-511.

• Bramfeldt, H., Sarazin, P., and Vermette, P. Blends as a strategy towards tailored

hydrolytic degradation of P(CL-co-D,L-LA)-PEG-P(CL-co-D,L-LA) copolymers. Submitted to Polymer Stability and Degradation.

• Bramfeldt, H., Vermette, P. Enhanced smooth muscle cell adhesion and proliferation on

protein-modified polycaprolactone-based copolymers. Journal of Biomedical Materials,

Part A, accepted for publication. Reference number: JBMR-A-07-0467.R1

• Bramfeldt, H., Sarazin, P., Vermette, P. Smooth muscle cell adhesion in surface-modified

three-dimensional co-polymer scaffolds prepared from co-continuous blends. Submitted to Biomacromolecules.

Page 9: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

VIII

ACKNOWLEDGEMENTS

First of all, I would like to extend a special thank you to my supervisor, Dr. Patrick

Vermette. You have reinforced in me the values of independence and perseverance. You were

never too busy to meet or to discuss. I truly feel I have become a better scientist under your

supervision.

Dr. Pierre Sarazin, a post-doctoral fellow in our group (presently at École

Polytechnique in Montréal). I would like to thank you for everything you have helped me

with: from the technical know-how and the pursuit of perfection to the constant asking of

nagging questions and general merriment. Not only has your knowledge and clear-sightedness

motivated me and helped me to restructure and refocus on many an occasion, but I find your

philosophy on life, your curiosity, and your sense of humour very refreshing!

My very own Matej. What can I say? Without you, what would all this mean anyway?

You have supported me and challenged me, and your good mood makes my good mood. I

thank you also for all the weekends spent at home or outdoors, alone or in the company of

good friends, exploring the most remote corners of Québec and elsewhere. It has been a

constant reminder of how much more there is to life than lab bench tops and evenings spent in

the less stimulating company of primary cell cultures.

To my family: I have always been able to count on you. Thanks for cheering me on,

for pretending to listen when I talk about my work, for the cross-words and the Christmas

decorations, and for never telling me it was a bad idea to come here. Here’s to many more

years of me boring you with fascinating scientific details!

Much appreciation is extended to my friends and colleagues, the patient and fantastic

people who have endured me on a close-to-daily basis for these four years. Without you I

would have given up a long time ago. Good luck to you all in the future, whatever it may

bring!

Finally, I would like to express my gratitude to Drs. Pentti Tengwall and Fredrik Höök

for first inspiring in me the courage to enter into the world of research, and for being the

scientific and personal models to which I shall always aspire.

Page 10: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

TABLE OF CONTENTS

CHAPTER 1: SCAFFOLD VASCULARIZATION: A CHALLENGE FOR THREE-

DIMENSIONAL TISSUE ENGINEERING .....................................................................1

1.1 Abstract .............................................................................................................................................................1

1.2 Introduction ......................................................................................................................................................1

1.3 Scaffold materials .............................................................................................................................................3 1.3.1 Biologically derived polymers .........................................................................................................3 1.3.2 Synthetic biomaterials........................................................................................................................5 1.3.3 Scaffold porosity ..................................................................................................................................8

1.4 Vascularization of biomaterials.......................................................................................................................9 1.4.1 Blood vessels and vascularization in vivo....................................................................................9 1.4.2 Angiogenic Factors...........................................................................................................................12 1.4.3 Tube formation in biological scaffolds .......................................................................................13 1.4.4 Tube formation in synthetic scaffolds..........................................................................................18 1.4.5 Co-culture models .............................................................................................................................20

1.5 Bioreactor culture and mechanical conditioning.........................................................................................23 1.5.1 Flow conditioning .............................................................................................................................24 1.5.2 Cyclic strain........................................................................................................................................25

1.6 Conclusions .....................................................................................................................................................27

1.7 References .......................................................................................................................................................29

CHAPTER 2: CHARACTERISATION, DEGRADATION AND MECHANICAL

STRENGTH OF POLY(D,L-LACTIDE-CO-Ε-CAPROLACTONE)-POLY(ETHYLENE

GLYCOL)-POLY(D,L-LACTIDE-CO-Ε-CAPROLACTONE) ...................................... 45

2.1 Abstract ...........................................................................................................................................................45

2.2 Introduction ....................................................................................................................................................45

2.3 Materials and Methods ..................................................................................................................................47 2.3.1 Materials..............................................................................................................................................47 2.3.2 Polymer synthesis..............................................................................................................................47 2.3.3 Film preparation ...............................................................................................................................48 2.3.4 Characterization................................................................................................................................48 2.3.5 Hydrolytic degradation ...................................................................................................................49 2.3.6 Tensile testing.....................................................................................................................................50

2.4 Results and Discussion ...................................................................................................................................50 2.4.1 Characterization of PEG-containing co-polymers..................................................................50

Page 11: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

2.4.2 Hydrolytic degradation ...................................................................................................................53 2.4.3 Tensile strength..................................................................................................................................59

2.5 Conclusions .....................................................................................................................................................60

2.6 Acknowledgments...........................................................................................................................................61

2.7 References .......................................................................................................................................................62

CHATPER 3: BLENDS AS A STRATEGY TOWARDS TAILORED HYDROLYTIC

DEGRADATION OF P(CL-CO-D,L-LA)-PEG-P(CL-CO-D,L-LA) COPOLYMERS .......... 66

3.1 Abstract ...........................................................................................................................................................66

3.2 Introduction ....................................................................................................................................................66

3.3 Materials and Methods ..................................................................................................................................68 3.3.1 Materials..............................................................................................................................................68 3.3.2 Copolymers, blend preparation and sample fabrication .......................................................68 3.3.3 Hydrolytic degradation ...................................................................................................................68 3.34 Statistical Analysis.....................................................................................................................................69

3.4 Results and Discussion ...................................................................................................................................70 3.4.1 Mass loss and buffer absorption ...................................................................................................70 3.4.2 Evolution of pH of degradation media........................................................................................71 3.4.3 Morphology.........................................................................................................................................73

3.5 Conclusions .....................................................................................................................................................74

3.6 Acknowledgments...........................................................................................................................................75

3.7 References .......................................................................................................................................................76

CHAPTER 4: ENHANCED SMOOTH MUSCLE CELL ADHESION AND

PROLIFERATION ON PROTEIN-MODIFIED POLYCAPROLACTONE-BASED

COPOLYMERS ...................................................................................................... 79

4.1 Abstract ...........................................................................................................................................................79

4.2 Introduction ....................................................................................................................................................79

4.3 Materials and Methods ..................................................................................................................................81 4.3.1 Materials..............................................................................................................................................81 4.3.2 Co-polymer synthesis .......................................................................................................................81 4.3.3 Film preparation ...............................................................................................................................82 4.3.4 Surface modification ........................................................................................................................82 4.3.5 Characterization of amine content ...............................................................................................83 4.3.6 Cell culture..........................................................................................................................................83 4.3.7 Cytotoxicity .........................................................................................................................................83

Page 12: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

4.3.8 Cell adhesion and proliferation ....................................................................................................84 4.3.9 Statistical analysis ............................................................................................................................85 4.4.1 Cytotoxicity .........................................................................................................................................86 4.4.2 Surface modification ........................................................................................................................87 4.4.3 Smooth muscle cell culture .............................................................................................................89

4.5 Conclusions .....................................................................................................................................................96

4.6 Acknowledgements .........................................................................................................................................97

4.7 References .......................................................................................................................................................98

CHAPTER 5: SMOOTH MUSCLE CELL ADHESION IN SURFACE-MODIFIED

THREE-DIMENSIONAL CO-POLYMER SCAFFOLDS PREPARED FROM CO-

CONTINUOUS BLENDS ........................................................................................ 102

5.1 Abstract .........................................................................................................................................................102

5.2 Introduction ..................................................................................................................................................102

5.3 Materials and Methods ................................................................................................................................105 5.3.1 Materials............................................................................................................................................105 5.3.2 Co-polymers......................................................................................................................................105 5.3.3 Preparation of blends ....................................................................................................................105 5.3.4 Annealing and scaffold fabrication............................................................................................105 5.3.5 SEM imaging ....................................................................................................................................106 5.3.6 Mechanical properties ...................................................................................................................106 5.3.7 Permeability of scaffolds...............................................................................................................107 5.3.8 Protein immobilization and smooth muscle cell seeding.....................................................107

5.4 Results and Discussion .................................................................................................................................110 5.4.1 Effects of annealing on pore size ................................................................................................110 5.4.2 Compressive strength .....................................................................................................................113 5.4.3 Permeability of scaffolds...............................................................................................................114 5.4.4 Smooth muscle cell adhesion on protein-modified scaffolds..............................................116

5.5 Conclusions ...................................................................................................................................................119

5.6 Acknowledgements .......................................................................................................................................119

5.7 References .....................................................................................................................................................120

GENERAL CONCLUSION AND SUGGESTED FUTURE WORK ................................ 123

APPENDIX .......................................................................................................... 129

A1. Protocol for synthesis of P(ε-CL-co-D,L-LA)-PEG-P(ε-CL-co-D,L-LA)................................................129

Page 13: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

A2. Protocol for surface modification, cell seeding, cell count, and imaging. ...............................................133

A3. Standard curve established for the ninhydrin assay used in Chapter 4. ................................................136

A4. Control α-actin staining of SMC in Chapter 4..........................................................................................137

Page 14: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

LIST OF FIGURES

Chapter 1

Figure 1.1 Polymeric structure of poly(glycolic acid), poly(lactic acid), and poly(ε-caprolactone) 6

Figure 1.2 The structure of a large blood vessel, depicting the intima, the media,

and the adventitia 10

Chapter 2

Figure 2.1 WAXD curves of films of PEG-containing copolymers 52

Figure 2.2 Water absorption and mass loss of P(100/0), P(70/30), and P(30/70)

during degradation in PBS (37°C, pH 7,4) for up to 26 weeks 54

Figure 2.3 Gel permeation chromatograms showing the effect of degradation on the polydispersity and molecular weights 55

Figure 2.4 The effect of degradation on the average molecular weight, Mn, as

determined by GPC and the effect of degradation on monomer composition as calculated from 1H-NMR spectra 57

Chapter 3

Figure 3.1 Mass loss (% of initial weight) and water absorption (% of dry weight) presented as a function of degradation time over 26 weeks 71

Figure 3.2 Effect of sample thickness and blend composition on the pH of the

degradation media. 72

Figure 3.3 Scanning electron micrographs of partially degraded P100/P30 copolymer blends. 73

Figure 3.4 Scanning electron micrographs of cross-sections of partially degraded PCL/P70, P100/P70, and P100/P30 copolymer blends. 74

Chapter 4

Figure 4.1 MTT mitochondrial activity assay of the cytocompatibility of degradation medium. 86

Figure 4.2 Absorbance at 570nm of aminolyzed copolymer films. 88

Page 15: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

Figure 4.3 Cell adhesion and proliferation on aminolyzed and protein-immobilized P(100/0) films. 90

Figure 4.4 Cell adhesion and proliferation on aminolyzed and protein-immobilized

P(70/30) films. 92

Figure 4.5 Scanning electron micrographs of SMC on modified P(100/0), 8h and 48h after seeding. 93

Figure 4.6 α-actin staining of SMC on modified P(100/0), 8h and 48h after seeding. 94

Figure 4.7 α-actin staining of SMC on modified P(70/30), 8h and 48h after seeding. 95

Chapter 5

Figure 5.1 Schematic drawing of cell-seeding device. 108 Figure 5.2 SEM images confirming continuous porosity in non-annealed samples of

P(100/0), P(70/30), and polystyrene. 110 Figure 5.3 Porous P(100/0): effect of annealing on pore size. 112 Figure 5.4 Porous P(70/30): effect of annealing on pore size. 112 Figure 5.5 Representative stress-strain curves for porous P(100/0) and P(70/30). 114 Figure 5.6 Compressive Young’s moduli and stress at 10% strain of porous P(100/0)

and P(70/30) for different times of annealing. 114 Figure 5.7 A log-linear plot of the permeability constant κ (m2) vs. annealing time

for porous P(100/0) and P(70/30). 115 Figure 5.8 Aminated copolymer scaffold cross-sections stained blue by ninhydrin

reactant. 117 Figure 5.9 Scanning electron micrographs of SMC adhering in a P(100/0) scaffold,

8h after cell seeding. 118 Figure 5.10 Number of SMC adhering on porous P(100/0) scaffolds 7h after cell seeding. 119 Appendix

Figure A.1 Setup for copolymer synthesis. 134 Figure A.2 Standard curve established for ninhydrin assay in Chapter 4. 136 Figure A.3 Control α-actin staining of SMC in Chapter 4. 139

Page 16: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

LIST OF TABLES

Chapter 1

Table 1.1 Some angiogenic factors used to promote vascularization of biomaterials. 14 Chapter 2

Table 2.1 Characterization of the three copolymers synthesized by ring- opening polymerization . 50

Table 2.2 Thermodynamic transitions, fusion energy, and % of crystallinity

of P(100/0), P(70/30), and P(30/70) films. 51 Table 2.3 Changes in mechanical properties with time of degradation. 59 Chapter 5

Table 5.1 Pore size and approximate porosity as a function of annealing time and temperature. 111

Page 17: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

LIST OF ABBREVIATIONS bFGF basic fibroblast growth factor PDLLA poly(D,L-lactic acid)

DMA dynamic-mechanical analysis PEG poly(ethylene glycol)

DMP dimethyl pimelimidate PEO poly(ethylene oxide)

DSC differential scanning calorimetry PGA poly(glycolic acid)

EC endothelial cells PHEMA poly(2-hydroxyethyl methacrylate)

ECM extracellular matrix PI polydispersity index

EDTA ethylenediamine tetraacetic acid, PLA poly(lactic acid)

ePTFE expanded polytetrafluoroethylene PLGA poly(lactic-co-glycolic acid)

FAP focal adhesion plaques PLLA poly(L-lactic acid)

Fgn fibrinogen PMA 4β-phorbol 12-myristate 13-acetate

FITC fluorescein isothiocyanate PS polystyrene

Fn fibronectin PTFE polytetrafluoroethylene

GAG glucosaminoglycan PVA poly(vinyl alcohol)

GPC gel permeation chromatography RGD arginine-glycine-aspartic acid peptide

HA hyaluronic acid SEM scanning electron microscopy

HMVEC human microvascular endothelial cells

SMC smooth muscle cells

HUVEC human umbilical vein endothelial cells

SnOct stannous octoate

MMP matrix metalloproteinase Tc, Tg, Tm

temperatures of crystallization, glass transition, and melting

MTT 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide

TCPS tissue culture polystyrene

NMR nuclear magnetic resonance TE tissue engineering

PBS phosphate buffered saline TGF- β Transforming growth factor-β

PCL polycaprolactone VEGF vascular endothelial growth factor

PDGF platelet-derived growth factor WAXD wide-angle x-ray diffraction

Page 18: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

1

CHAPTER 1:

SCAFFOLD VASCULARIZATION: A CHALLENGE FOR THREE-DIMENSIONAL TISSUE

ENGINEERING

1.1 Abstract

The prevalent challenge facing tissue engineering today is the lack of adequate vascularization

to support the growth, function, and viability of tissues. Researchers rely on the increasing

knowledge of angiogenic and vasculogenic processes to stimulate vascular network formation

within three-dimensional tissue constructs. These processes are mainly endothelial cell-

regulated, although in the context of tissue engineering, specific interactions with scaffold

materials, growth factors, and other cell types may require in vitro vascularization schemes to

be altered accordingly. In order to better mimic the complete in vivo environment, increasing

attention is given to the integration of co-cultures and mechanical conditioning in bioreactors.

Such approaches show great promise for the enhancement of the functionality and clinical

applicability of tissue engineering constructs.

1.2 Introduction

Microvessel beds constitute the infrastructure that provides nutrient delivery and waste

removal in living tissues. In the past few decades many an effort has been made to engineer

functional vascular networks in vitro. Paralleled by progress in tissue-engineered blood

vessels and in the decoding of the complex mechanisms of blood vessel formation, this work

has also generated insight valuable for the repair and reconstruction of three-dimensional

human tissue. There are two principal reasons for promoting the formation of vascular

networks within three-dimensional cell/scaffold constructs. Firstly, in vitro angiogenesis

assays provide useful tools for detailed study of this process by identifying and evaluating the

role of specific biochemical pathways[1]. Secondly, in the absence of a vascular network that

can ensure efficient exchange of metabolites in a growing cell mass, artificial tissue constructs

will remain severely limited in obtainable size, function, and applicability in vivo[2].

Page 19: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

2

Vascularization of newly formed tissue in vivo takes place by angiogenesis; the process

of sprouting of new capillaries from pre-existing vasculature[3]. Angiogenesis is essential for

wound healing, tissue repair, and the growth of newly formed tissue. Tissue engineering (TE)

constructs may also become vascularized via a process known as vasculogenesis; the in situ

reassembly of undifferentiated endothelial cells into capillaries and networks[4]. In vivo, this

process normally takes place in developing organs in the embryo[5]. Specific molecules,

called angiogenic factors, that control blood vessel formation have been identified[3] and

some of these are frequently used to enhance vascularization processes. Following the creation

of vascular networks within a TE construct, whether established in vitro or in vivo, these must

ultimately connect to the host vasculature after implantation to fulfill their intended function.

This event is called inosculation and is the process by which the vasculature is rendered

continuous. Finally, a tissue repaired or regenerated with the help of an implanted material,

like any other tissue, will be subject to vascular remodeling with time.

At present, regeneration/repair of cartilage[6], bone[7], and skin[8] – although still

dealing with issues of predictability and long-term stability – can be achieved with available

techniques. However, larger and/or more metabolically demanding tissues such as the

pancreas, liver, heart, and kidney require intrinsic blood supply for long-term survival and

development[9]. For example, it has been suggested that hepatocytes may not be grown

further away than 200-300µm from nutrient supply in order to survive[10]. Vascular networks

are therefore critical for the efficient distribution of oxygen and nutrients to all tissues, as well

as for the removal of waste products. In fact, for any cell mass large enough that it cannot rely

on nutrient supply solely by diffusion, growth, function, and survival post-implantation are

improved by and dependent on ingrowth of blood vessels from the host or, alternatively, by

inosculation of an already existing network[11-14]. The idea of prevascularized TE scaffolds,

as brought forward by Mikos et al.[15], was shown to enhance the performance of forming

tissues compared to non-vascularized constructs[11,16]. However, most attempts to promote

vascularization of scaffolds have been pursued in vivo, by allowing for ingrowth of

vascularized fibrous tissue, which then comes to occupy an important fraction of the available

void space, thus limiting the space available for engraftment of tissue-specific cells[15]. At

present, the development of functional vascular beds in three-dimensional constructs presents

Page 20: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

3

a great challenge to researchers, but is no less central to the successful therapeutic application

of tissue engineering principles.

This review treats the background and recent advancements in the vascularization of

porous scaffolds for three-dimensional tissue engineering, with an emphasis on in vitro tissue

culture. Moreover, the effects of co-culture techniques as well as mechanical conditioning of

tissue constructs within the confines of a bioreactor are discussed.

1.3 Scaffold materials

The requirements for three-dimensional TE scaffolds are generally defined in terms of

biocompatibility, degradability, porosity (void volume), pore size, and mechanical properties.

In addition, scaffolds should promote cell-material interactions suitable for a particular

application, either inherently or as a result of surface modification. The choice of a suitable

biomaterial is the first step in the scaffold design process. Polymeric biomaterials can be of

either biological or synthetic origin, and they offer tremendous variation as pure preparations

or as co-polymeric devices. Control over scaffold properties is crucial in the design of

successful scaffolds. Ideally, a scaffold should be highly porous, permeable to flow, and

possess mechanical strength to match the requirements of the tissue to be replaced. On the

same note, it is important that the rate of degradation match an appropriate rate of tissue

growth as it proceeds within the scaffold. Adequate pore size, pore morphology, and a high

degree of interconnectivity are important to achieve efficient cell seeding and growth.

Furthermore, scaffolds should be favorable to fluid flow, ensuring sufficient delivery of

nutrients and removal of waste and scaffold degradation products[17].

1.3.1 Biologically derived polymers

ECM-DERIVED POLYMERS, such as collagens, glycoseaminoglycans (GAGs), and fibrin

gels, are attractive scaffold materials in that they originate from the very tissues that need to

be regenerated. As components of the extracellular matrix (ECM) they offer natural sites for

cell attachment, and signaling that regulate cell phenotype. Traditionally, this group of ECM-

based biomaterials has been the most popular choice for angiogenesis assays and

vascularization studies. They are also used as coating materials to improve cell-material

interactions on synthetic polymeric devices.

Page 21: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

4

Collagen is the most abundant protein in mammalian tissues, and is therefore an

interesting scaffold material. Chemical or physical cross-linking may be employed to improve

the mechanical performance of collagen fibers and scaffolds and to moderate enzymatic

degradation of these scaffolds, although cross-link density and porous microstructure are

known to influence cellular interactions with these templates[18,19]. Immunogenic issues,

especially with animal collagen, have been a concern for the application of collagen-based

biomaterials. This effect is primarily due to telopeptides, which may be efficiently removed

enzymatically[20]. However, the full immunogenic potential may not be completely

removable from non-human proteins, and methods for production of recombinant collagen I

and III have been proposed as an alternative source[21].

In response to injury, a fibrin clot forms spontaneously from fibrinogen and thrombin, and

are further cross-linked by factor XIIIa. The implication of fibrin in blood vessel repair makes

it a natural candidate for angiogenic assays[22,23], and has also been used for controlled

release of growth factors in vascular graft engineering and bone regeneration[24-26].

GAGs occur naturally as linear polysaccharide branches of certain protein structures called

proteoglycans. Hyaluronic acid is the simplest of the GAGs, and is present in nearly all

mammalian tissues. Its high molecular weight (50-500kDa) confers mechanical properties that

make it interesting as a scaffold material, and it has been used mainly in ophthalmology and

experimental treatments of joint dysfunction[27]. Pioneering work using collagen-based

materials was carried out by Yannas et al.[28], who developed templates for regeneration of

skin and nerve from co-polymers of collagen and another type of GAG, namely chondroitin 6-

sulfate. Fibrous materials based on fibronectin have also been investigated for similar

applications[29,30].

POLYSACCHARIDES, such as chitosan and alginate, that have been shown to elicit a

relatively mild immunogenic response, have also been investigated for various TE

applications. Chitosan has a molecular structure similar to GAGs, and is susceptible to

enzymatic degradation by the human enzyme lysozyme. The rate of degradation depends on

the degree of crystallinity, which can be easily manipulated by varying the acetyl

content[19,31]. Chitosan has been investigated for such applications as encapsulation,

membrane barriers, and contact lens material[31]. Moreover, DNA complex formation is

favored by its cationic nature, making it an ideal candidate for gene delivery[32]. More

Page 22: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

5

recently, hybrid scaffolds containing a synthetic component have also been proposed for drug

release and engineering of cartilage and bone[33,34]. Alginate gels are fabricated under

relatively mild conditions, making it suitable for application in drug delivery and cell

transplantation (e.g. chondrocytes and insulin-producing cells)[19,35]. In particular, growth

factors incorporated into alginate beads have been shown to retain >90% of their

bioactivity[36]. Disadvantages include issues of poorly regulated degradation and the lack of

natural cell adhesion motifs[37].

1.3.2 Synthetic biomaterials

Despite the obvious advantages of using biologically derived materials, important

drawbacks limit their use as scaffolds. These include poor mechanical properties, and inherent

batch-to-batch variability. Synthetic polymers therefore represent an attractive alternative,

especially for the fabrication of large and/or load-bearing scaffolds. These materials offer a

high level of control, as well as seemingly endless variations, through the manipulation of

monomer composition, molecular weight, degree of cross-linking and/or branching, co-

polymerization, blending, etc. These and other polymer processing techniques may be used as

tools to manipulate such material properties as degradability, mechanical strength, and surface

chemistry.

Many classes of resorbable synthetic polymers are being studied for use in tissue

engineering[38]. Degradable polyesters, such as poly(lactic acid) (PLA), poly(glycolic acid)

(PGA), poly(ε-caprolactone) (PCL), and their co-polymers have been approved by the FDA

for a number of medical and drug delivery devices, and are thus very attractive candidates for

scaffold fabrication. These biocompatible polymers exhibit interesting and tunable mechanical

properties and predictable degradation kinetics. PLA, PGA and their copolymers (PLGA)

were utilized in early applications for sutures and orthopedic fixations[39]. PGA is an inelastic

and highly crystalline material. Poly(lactic acid) exists in two practically employed isoforms,

namely the highly crystalline poly(L-lactic acid) (PLLA) and the amorphous poly(D,L-lactic

acid) (PDLLA). Although less crystalline than PGA, PLA is more hydrophobic, and less

susceptible to hydrolysis[2]. PCL possesses properties that set it apart somewhat from the

other polyesters mentioned. It has a very low glass transition temperature (Tg = -60ºC), low

melting temperature (Tm = 55-60ºC), and shows high thermal stability. Its decomposition

temperature is approximately 100ºC higher (Td = 350ºC) than PLA and PGA, which makes it

Page 23: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

6

better suited for some processing techniques. Due to its low Tg, PCL is rubbery at room and

physiological temperatures, a temperature that contributes to its high permeability for many

drugs[31]. PCL is furthermore a flexible plastic that degrades at a much slower rate than either

PLA or PGA.

H [ O (CH2) C ] OH H [ O CH2 C ] OH

O

n

H [ O CH C ] OH

O

n

CH3

O

n5

a) b) c)

H [ O (CH2) C ] OH H [ O CH2 C ] OH

OO

n

H [ O CH C ] OH

OO

n

CH3

O

n5

a) b) c)

Figure 1.1: Polymeric structure of a) poly(glycolic acid), b) poly(lactic acid), and c) poly(ε-

caprolactone).

Polyesters can be easily co-polymerized, to achieve materials with intermediate

properties. For example, block co-polymers of PCL and PLA combines the rapid degradation

of PLA with the flexibility of PCL[40]. Incorporation of poly(ethylene glycol) (PEG) has also

been shown to improve water uptake, and speed up degradation of PCL[41]. Vert and

coworkers synthesized a series of block co-polymers containing different combinations of

PLA/PCL/PEG to achieve materials exhibiting a wide range of mechanical properties and

degradation rates[42-45]. This work clearly shows that such properties may be efficiently

tailored to meet the demands of particular applications. In addition, PLA-PGA co-polymers

(PLGA) are commonly occurring biomaterials, currently in clinical use for surgical sutures,

orthopedic implants, and as drug delivery systems[46].

The main concern of these materials is related to the hydrolytic degradation mechanism,

which propagates by ester bond cleavage and generates acidic by-products. These lower the

local pH, increasing the rate of bulk degradation via autocatalysis, limiting cell viability in the

proximity of the implant, and in some cases eliciting inflammatory reactions[47].

Furthermore, PGA and PLA have a tendency to crumble during degradation, creating large

particles that may remain in the site for years. Such uncontrolled breaking of implants may

also cause damage to nascent tissue[48].

Beside the classic group of aliphatic polyesters, a large quantity of biodegradable

polymer and co-polymer systems have emerged for biomaterial systems. These include

Page 24: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

7

polyanhydrides, polyorthoesters, polyphosphazenes, and polyurethanes. Polyanhydrides have

been extensively investigated for use in drug delivery for over 25 years[49-51]. Initial

limitation due to extensive degradation was addressed by researchers and resulted in materials

with degradation rates ranging from weeks to years[7]. Vascularization of polyanhydride

implants was observed after 4 weeks in vivo[52]. Polyorthoesters are hydrophobic polymers

that are stable at physiological pH, but degrade rapidly at pH 5.5 via hydrolysis[38,53]. These

two families of biodegradable polymers have found principal use in drug-delivery systems, as

they are mechanically unsuited for scaffolding purposes[48,54]. Polyphosphazene-based

biomaterials range in stiffness from soft to hard, show great backbone flexibility, and are

being investigated for their applicability in engineering tissue[38]. Laurencin and coworkers

studied the applicability of these materials for bone regeneration, and reported

osteocompatibility as well as non-toxicity of the degradation products[55]. Composite

materials containing hydroxyapatite – a natural component of bone – were investigated for the

same application[56]. Furthermore, blending polyphosphazenes with PLGA was shown to

moderate PLGA degradation, even as the polyphophazene degradation products had a

buffering effect on the degradation milieu[57].

Polyurethanes constitute yet another class of polymers that have long been in use as

implant materials[58]. They consist of alternating “soft” segments (polyether or polyesters)

and “hard” segments (diisocyanate), and chain extenders that may be either diols or diamines.

Great flexibility in chemical composition is possible through the choice of the soft and hard

segments, and by varying the ratios of these components a wide range of mechanical

properties can be obtained. Earlier concerns regarding the putative carcinogenity of the

degradation products have been attended to mainly by replacing diisocyanate by e.g. 1,4-

diisocyanatobutane[59] and lysine diisocyanate[60], which degrade into biocompatible

components. The latter was polymerized with glucose, which produced glucose and lysine as

the principal degradation products[60].

Synthetic polymers may also be processed into hydrogels, which by definition have high

water content, ranging from 10-20 wt% and up. Synthetic hydrogels structurally resemble the

ECM of many tissues[19] and present inherently good mass transport properties[61]. The

structural stability of synthetic hydrogels can be controlled by the degree of cross-linking

between polymer chains. Typically, hydrogels used for scaffold applications can be processed

Page 25: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

8

under relatively mild conditions, allowing for the incorporation of bioactive molecules or even

cells. The disadvantages include issues of sterilization and low mechanical strength[61,62].

Synthetic polymers used to make hydrogels include poly(ethylene oxide) (PEO) and the

chemically similar poly(ethylene glycol) (PEG), poly(vinyl alcohol) (PVA), poly(acrylic

acid), and polypeptides[19]. Photo-polymerisable PVA hydrogels are injectible and found to

be elastic and strong. Methods to improve the inherently poor cell adherence in PVA and PEG

gels by modification with an RGD cell adhesion peptide have been developed[63,64].

Hydrogels have also been fabricated from PEO-containing block co-polymers with PLLA[65]

and poly(propylene fumarate)[66], although the latter exhibited certain cytotoxicity in vitro .

Finally, newly synthesized materials have also been proposed for use in TE applications.

Matrices based on PCL copolymerized with hydrophilic poly(ether ester amide)s have been

investigated for drug delivery[67], and polypyrroles, possessing modifiable electrical

properties have been suggested for nerve regeneration[68].

1.3.3 Scaffold porosity

The optimal pore size of a scaffold varies with application[46], but needs to be large

enough that cells can easily penetrate into the scaffold bulk, while at the same time optimizing

surface area, which usually increases with decreasing pore size. Highly porous scaffolds are

usually preferred, as the surface area and the capacity for large cell mass increases with void

volume. On the other hand, excessively porous structures become fragile and may not fulfill

the mechanical requirements. For the regeneration of bone, pore sizes of 100-400µm are

usually recommended[69,70], whereas they should be 20-150µm for skin regeneration[71],

and approximately 20µm for ingrowth of hepatocytes and fibroblasts[46]. In the case of

neovascularization, data must be interpreted in the context they are generated. For example,

independent microvessel invasion of membranes or thin porous scaffolds have been reported

for pore sizes of 5-15µm[72,73]. However, neovascularization of a TE scaffold (especially for

larger structures) may also take place in conjunction with tissue ingrowth, and would in such a

case be more dependent on a pore size that promotes successful invasion of the tissue in

question. For example, van Tienen and coworkers[74] observed ingrowth of vascularized

fibrous tissue in pores that measured at least 30µm. It should be noted that reported results

often depend on scaffold design and on the material chemistry that often differs from one

study to another. Induction of bone regeneration has been reported in biodegradable scaffolds

Page 26: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

9

with pore sizes as small as 16-32µm[73], while in a general manner, much larger pore sizes

are found in the literature, as mentioned.

Pore morphology and pore interconnectivity also play a significant role in regulating cell

attachment, proliferation, and matrix deposition – e.g. for bone ingrowth[70] and in the culture

of smooth muscle cells[75,76]. Subcutaneous PVA sponges, with well defined pores were

found to induce better vascularized tissue ingrowth than did PTFE implants of similar pore

size[77]. Furthermore, pore microstructure and interconnectivity strongly influence mass

transport (permeability and diffusion) and mechanical properties, as well as degradation

profiles[46,78-80]. Numerous techniques exist for the fabrication of porous scaffolds, and

these have been reviewed elsewhere[81].

1.4 Vascularization of biomaterials

A variety of in vitro assays for angiogenesis exists. Classical testing schemes such as the

chicken chorioallantoic membrane (CAM) assay, pouch assays or the study of

neovascularization of the avascular cornea are still in use, but have to some extent been

replaced by reconstituted biological materials (for reviews, see e.g. Cockerill et al.[82,83]).

These assays provide a simplified but controlled environment in which the processes and

mechanisms of tubular network formation can be studied in great detail. Given the proper 3D

environment, endothelial cells will migrate into the matrix bulk, where they will rearrange and

form a network of capillaries, which may or may not contain lumina[84]. Developments

within the field tend to consist of substrates, growth factor delivery, and culture techniques

that increasingly mimic the complex physiological environment of the cells. However,

obtaining patent networks (i.e. that are capable of withstanding the strain of a circulating

medium) still presents a challenge for researchers. Encouragingly enough, it was recently

shown that engineered endothelial networks of tube-like structures indeed have the potential

to evolve to functional blood vessels in vivo[85]. The presence of such patent networks within

TE constructs would greatly facilitate the invasion of tissue-specific cells and enhance the

viability and function of the forming tissue.

1.4.1 Blood vessels and vascularization in vivo

The structure of all veins and arteries in the body consist of three distinct layers: the

intima, the media, and the adventitia[86] (see Figure 1.2). The adventitia is the outermost

Page 27: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

10

layer of the blood vessels, and consists primarily of connective tissue and collagen fibers,

which allows the vessel wall to return to its initial shape following dilatation or contraction.

The media is composed mainly of elastic fibers and mural cells. Mural cells is the common

name referring to either smooth muscle cells, as in the case of larger vessels, or pericytes,

which are rather undifferentiated and present in newly formed and small blood vessels. The

role of these cells is to stabilize the vessel structure. Smooth muscle cells also provide the

mechanically dynamic properties of larger blood vessels. Finally, the intima makes up the part

of the blood vessel which is in actual contact with the blood stream. This layer is composed of

a monolayer of endothelial cells, supported by the basal lamina2, and thin layer of connective

tissue. Capillaries are much smaller, and consist only of a layer of endothelium, lacking both

the media and adventitia. The cells on which the focus will be placed in this text are

endothelial cells (EC) and smooth muscle cells (SMC), due to their implication in mature

vessel structures in vivo.

Figure 1.2: The structure of a large blood vessel, depicting the intima, the media, and the adventitia.

(From original image http://en.wikipedia.org/wiki/Image:Anatomy_artery.png#file)

2 A layer rich in collagens and laminin, deposited by cells.

Page 28: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

11

New blood vessels may form by one of the two following processes: 1) by angiogenesis,

which is the sprouting of new blood vessels from preexisting ones, or 2) by vasculogenesis,

which is the assembly of undifferentiated EC in situ.

Vasculogenesis proceeds through five basic steps[4]: 1) generation of differentiated EC

from bone marrow–derived precursor cells; 2) EC form aggregates and establish cell-cell

contacts; 3) EC polarize and rearrange into nascent lumen-containing tubes; 4) an immature

vessel network is formed; and 5) pericytes and SMC are recruited, which stabilizes the newly

formed vessels.

Angiogenesis involves a growth phase and a stabilization phase[87]. Four steps define

the growth phase, which is initiated by 1) cell-mediated, proteolytic degradation of the

basement membrane, allowing for endothelial cells to invade the surrounding ECM; 2)

Activated EC then proliferate and migrate into the created space, where they 3) organize into

aligned, lumen-containing cords. 4) Lastly, newly formed sprouts connect to form loops of

immature vasculature. The immature vasculature is subsequently stabilized through the

recruitment of smooth muscle cells or pericytes, accompanied by a halt in EC proliferation,

and the deposition of a basement membrane (reinforced basal lamina) around the cords.

The first reports of in vitro tube formation by capillary endothelial cells was made in

1980 by Folkman and Haudenschild[84], who observed capillary networks developing on a

gelatine substrate after several weeks of culture. Since then, methods have been refined, as the

understanding of angiogenic processes has evolved. Today it is largely recognized that the

interplay between EC and insoluble molecules in the ECM is critical in many events during

angiogenesis and vasculogenesis, including the regulation of cell growth, migration,

differentiation and tube formation[88]. Vessel growth is also influenced by cell-cell

interactions and soluble species, such as growth factors. In addition to the choice of substrate,

the source of cells will influence the outcome of a study. To some extent, endothelial cells

from rat or bovine aorta behave differently than ECs from human tissue. The latter have

higher requirements of growth factors, and may degenerate in some basic culture conditions

where animal cells spontaneously differentiate into capillary structures[89]. Further

distinctions can be made between human ECs of different origin. The commonly used human

umbilical vein ECs (HUVEC) are of macrovascular origin, whereas the cells most involved in

inflammation, wound healing and vascularization in vivo are microvascular ECs. However,

Page 29: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

12

any difference in the capacities of these two cell types to form tube-like structures is not

readily apparent[90]. Some differences may also be found in the ability to exert contractive

forces on malleable substrates, where e.g. human blood outgrowth ECs are stronger compared

to HUVEC[91]. In addition, recent publications have demonstrated capillary formation using

endothelial progenitor cells, who are implicated in post-natal neovascularization[92,93].

1.4.2 Angiogenic Factors

A number of growth factors and other angiogenic promoters are used to induce and

enhance EC sprouting and capillary network formation in cell/polymer constructs. These

factors may act directly on EC or indirectly, inducing an angiogenic response in vivo, without

eliciting a direct response in cultured EC, such as an altered migration or proliferation[4].

Vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (bFGF) are

examples of highly potent EC mitogens, stimulating angiogenesis via direct interactions with

EC. Both growth factors are widely distributed in various tissues in the body, but while bFGF

is known to act on various cell-types, VEGF is EC-specific[4]. Other angiogenic factors, such

as platelet-derived growth factor (PDGF) and transforming growth factor-beta (TGF-beta), act

indirectly to induce blood vessel formation and/or stabilization[4,94]. Table 1.1 lists these

molecules and others, their roles in angiogenesis in vivo, and examples of their use to induce

angiogenesis in biomaterials.

Many groups have explored the controlled delivery of angiogenic growth factors as a

means to improve the local vessel recruitment in implants and cell transplants[11,16,103]. The

modulation of growth factor delivery is highly important since growth factors such as VEGF

and bFGF are rapidly cleared and inactivated following administration in an unprotected

formulation[16]. Another aspect is that bolus delivery may cause a toxic or otherwise

unwanted response in the studied (or surrounding) cell population, as such administration does

not correspond to any normal situation in vivo[101,111]. Moreover, if too high concentrations

of e.g. VEGF are sustained within a tissue, the developing vasculature may become deformed

and nonfunctional[112]. Control of the release rate is therefore of utmost importance. To deal

with these concerns, the encapsulation of drugs within degradable polymers from which they

can be liberated in a controlled manner has come forth as a successful method. Biologically

derived hydrogels, including fibrin, alginate, hyaluronic acid, chitosan, and gelatine have

proved very useful as encapsulation vehicles due to their biocompatibility and inertness

Page 30: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

13

towards bioactive molecules[16,19,25]. Moreover, encouraging results have been reported

from trials with synthetic materials such as PLGA[11,36,102,103] and PEG hydrogels[98].

Growth factors may be released from encapsulating matrices via diffusion, mechanical

stimulation, or matrix degradation[19]. Control of the release rate may be obtained by varying

the rate of degradation of the capsule material, and this may in turn be achieved by altering the

molecular weights or degree of polymer cross-linking[102]. Another very interesting approach

consists of pacing the release of growth factor from the matrix to the demand of the growing

cell popluation[98,101,113]. Such designs aim at more closely mimicking the way in which

naturally sequestered growth factor is released from the ECM by proteolytic activity, and have

been shown to induce a more controlled growth of capillary networks in fibrin[101].

1.4.3 Tube formation in biological scaffolds

A few methods dominate the three-dimensional in vitro angiogenesis assays in current

use. In so-called in-gel seeding cells are mixed with the matrix solution prior to gelation. Cell-

seeded collagen type I gels prepared in this manner were not able to induce EC tube formation

in the absence of endothelial mitogen[114]. Alternatively, ECs can be cultured to confluence

or near-confluence on top of a gel (e.g. fibrin, collagen, Matrigel3), at which time a second

layer is placed on top of the cells, creating a “sandwich” consisting of two layers of gel

matrix. Trapped within the sandwich, the cells may subsequently differentiate, migrate and

form tubes, an event which is largely dependent on the interaction between the overlaying

matrix layer and the apical cell surface[116, 117]. A third approach involves the use of intact

ring sections of aortas that are embedded in collagen, fibrin, or Matrigel gels, and the

angiogenic activity is observed as radial sprouting of endothelial cells[118]. This so-called

aortic explant model has traditionally employed segments of rat aorta[119], but can be

extended to include tissues from other animals[120]. The advantage of this method compared

to cell-based assays is the realistic representation of in vivo structural cell arrangements[120].

In addition to the models mentioned, in vivo experiments are conducted in which seeded

or acellular matrices are implanted subcutaneously in mice, where a capillary network starts to

form. Subcutaneous implantation of fibrin and Matrigel lead to blood vessel ingrowth[22,

121], while collagen type I did not[22,121]. Vascularization of the implant may be evaluated

after surgical recovery[122] or, alternatively, in vivo through transparent “windows” in the 3 Matrigel: a commercial, multi-molecular matrix material extracted from Englebreth-Holm-Swarm tumors[115]

Page 31: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

14

skin of the animal[85]. (Note: Several approaches are currently being used to quantify

growing and established 3D tubular networks within a biomaterial. The lack of a standardized

method for quantification makes the direct comparison between different assays and results

complicated.)

Table 1.1: Some angiogenic factors used to promote vascularization of biomaterials.

*Coll I: collagen type I; CAM: chorioallantoic membrane; PLGA: poly(lactic-co-glycolic

acid).

Angiogenic agent Roles in vivo Examples of use in

biomaterials in vitro*

Vascular endothelial growth factor (VEGF)

Potent angiogen, required for the formation of

immature vasculature[95], promotes ECM

degradation[94]. Exists in five isoforms of different

potency. Must be well regulated. Reinforces the

angiogenic potential of bFGF[96].

Coll I[1,97], fibrin[98-

100] ,

fibrin and CAM[101],

PLGA[102,103],

PEG[98]

Basic fibroblast growth factor (bFGF)

Stimulates expression of VEGF and its receptors in

EC. Promotes ECM degradation[94], EC migration

and pericyte attraction[104]. Reinforces the

angiogenic potential of VEGF[96].

Coll I[1,97,105]

Fibrin[99,100]

PLGA in alginate[11]

Tissue growth factor-beta (TGF- beta)

Induces mural cell differentiation[3], inhibits ECM

degradation[94]. Dose-dependent stimulation of

tube formation in vitro [106]. EC proliferation and

migration are inhibited at higher

concentrations[107].

Matrigel[106]

Platelet-derived growth factor (PDGF)

Recruitment of pericytes/SMC[83]. PLGA[102]

Angiopoietins (Ang-1, Ang-2)

Ang-1 stabilizes vessels, while its antagonist Ang-2

acts as a destabilizer, promoting EC migration and

angiogenesis in presence of other factors, such as

VEGF[108,109].

Co-culture in Coll

I[109]

4 beta-phorbol 12-myristate 13-acetate

(PMA)

Known tumor promoter. Coll I[1,97,110]

Page 32: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

15

The various matrices used for endothelial cell culture each represent different

characteristics of the basement membrane. Therefore it is not surprising that the mechanism of

tubulogenesis in vitro in some aspects is matrix-specific. It has even been suggested that the

nature of the supporting material influences the angiogenic potential of growth factors[123].

The specific interactions between EC and the matrix are mediated by integrin receptors on the

cell surface[124]. Dallabrida et al.[125] observed matrix-specific tubulogenesis in a

publication where down-regulation of the expression of the

alpha v beta 3 integrin (a receptor expressed by angiogenic EC, binding both VEGF[126] and

PDGF-B[127]) in ECs reduced tubulogenesis in fibrin gels, whereas no such effect was seen

on Matrigel. Moreover, Hall et al. were able to induce angiogenesis on fibrin substrates,

covalently modified with synthetic binding sites for the alpha v beta 3 integrin[124].

Chalupowicz et al.[117] identified the interaction between HUVEC and the beta15-41

sequence of fibrinopeptide B as necessary for tube formation in fibrin sandwich.

Soluble species and matrix constituents work in concert with many cell membrane-

associated adhesion molecules during the angiogenesis process. Among these are the cell-cell

junction proteins VE-cadherin (CD144) and PECAM-1 (CD31), which have been

demonstrated to regulate lumen formation and intercellular association during EC network

formation, respectively[128]. In addition, there is evidence to suggest that VE-cadherin is

necessary for the maintenance of tubular networks in fibrin and collagen type I gels[1, 129].

Antibodies against VE-cadherin not only inhibited tubulogenesis but also disrupted already

developed networks[129].

The controlled and localized release of the growth factor bFGF has been shown to

significantly enhance vascularization of injured sites[16], and is furthermore suggested to

increase the size of growing vessels[11]. Sakakibara and coworkers[16] reported that slow

release of bFGF from gelatin microspheres in infarction areas improved cardiac function and

survival of transplanted myocytes in rats. Endothelial cells grown on gelatin-coated micro-

carriers dispersed in a fibrin gel exhibited important increases in their capacity to sprout when

bFGF or VEGF was added to the gel[100]. Ehrbar and coworkers[101] used an engineered

version of VEGF121 that was covalently incorporated into fibrin gels and subsequently

evaluated in the CAM assay. Grafted VEGF was gradually released as the matrix was

enzymatically degraded by EC, and enhanced arterial and venous network development was

Page 33: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

16

observed. In a control experiment, non-grafted soluble VEGF121 was released rapidly and

triggering chaotic vessel growth. Other examples of VEGF encapsulation include the use of

alginate[19,36] or combined alginate/PLGA matrices[102,103]. Furthermore, attempts have

been made to coordinate the controlled delivery of multiple growth factors[102]. In an

approach combining bFGF and VEGF delivery, Nillesen et al.[130] used collagen-heparin

sponges with 100µm pores to stimulate vascularization subcutaneously in rats. (Heparin

increases the amount of bFGF that may be incorporated into cross-linked collagen[105] and

alginate[131] gels.) They found that while the two growth factors alone supported vessel

formation, the combined delivery of bFGF and VEGF was superior in promoting the

formation of mature, pericyte-lined, vasculature within the scaffolds. In addition, there was a

dramatic decrease in the number of hypoxic cells in the scaffold centre when both growth

factors were present[130].

INFLUENCE OF SUBSTRATE STIFFNESS

It has long been claimed that endothelial migration and elongation take place along lines

of reorganized matrix – a result of cell forces acting on the matrix components[132]. These

forces are communicated via cell adhesion sites called focal adhesion plaques (FAP) that

consist of assemblies of structural and signaling proteins, such as integrins[133]. These sites

enable mechanical tension from the ECM to be transferred into the cell, and conversely from

the cytoskeleton to the ECM[134]. Relatively rigid substrates have been found to better

support the formation of stable focal contacts than more compliant substrates, by promoting

stronger integrin-cytoskeletal linkages[135]. However, they also bring on a more spread cell

morphology, which in turn leads to decreased cell elongation and cells organized in isolated

clusters or monolayers rather than network formations[91,132,136,137]. More malleable

substrates promote a rounded morphology that is associated with a more differentiated

phenotype in endothelial cells, although fully rounded cells will not form capillary tubes[136].

Appropriate matrix rigidity is not associated with any absolute magnitude of “stiffness”,

but needs to be put into a more elaborate picture, incorporating matrix type, matrix geometry,

conditions of constraint, and the cell type involved. The force generated by cells cultured in or

on a gel construct depends both on the cell type and on the flexibility of the matrix[91].

However, it has been clearly established that matrix contraction and tube formation within

Page 34: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

17

reconstituted gels require an active cytoskeleton[137-139]. During gel contraction the number

of stress fibers (actin filaments) present in cells can be related to the stress being imposed on

the cytoskeleton, which in turn is a reflection of the matrix resistance to deformation[91,137].

Moreover, stress fibers become oriented along the tubule axis following EC rearrangement

into capillaries[137]. Analysis of the concentration and organization of actin filaments and

proteins associated in focal adhesion plaques reveals EC-specific morphology changes in

response to changes in substrate flexibility[140].

Substrate stiffness can be altered by changing the polymer concentration, degree of

cross-linking or polymerization conditions. Ingber and Folkman[136] reported on the role of

ECM concentration on rigid surfaces, and found that intermediate coating densities supported

EC rearrangement into tubular structures. Several groups have later repeated these results in

thick substrates concluding that the onset of tubulogenesis can be controlled by altering the

stiffness of the cell support[132,137]. For example, on Matrigel or polyacrylamide gels that

were prepared with varying degree of cross-linking, HUVEC reorganized more quickly into

tube-like structures on more compliant substrates[137]. Similar results were obtained by

decreasing the density of collagen gels[141]. Furthermore, in a recent publication Sieminski

and coworkers[91] demonstrated the possibility of controlling lumen size by varying matrix

concentration. By doubling the concentration of type I collagen from 1.5mg/ml to 3mg/ml

lumen size was increased by 57% and 40% in floating and constrained gels, respectively[91].

This increase was attributed to a resistance to contraction (increasing with the density of the

substrate), bringing on a more spread cell morphology.

In order to study the role of cellular adhesion in tube formation, the inherent ability of

cells to adhere to the support material can be manipulated. Many cell types, including

endothelial cells, are dependent on the activity of the beta 1 integrin unit for adhesion on a

variety of ECM-derived substrates. Tubule formation in Matrigel is, for instance, inhibited

when the beta 1 unit is blocked by antibodies[125]. Indeed, the alpha 2 beta 1-integrin is the

main laminin and collagen receptor in EC[137], and varying the adhesion strength using anti

alpha 2 beta 1-antibodies, influences both the number and dimension of forming capillary

tubes[110]. Deroanne et al.[137] found that as the rigidity of their collagen type I substrate

decreased, the expression of the alpha 2 subunit was reduced, and EC more easily switched to

a tube-like pattern. Furthermore, they established that parallel to these effects, increased

Page 35: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

18

substrate malleability also resulted in reduced expression of actin and focal adhesion plaques

proteins, as determined by Western blot[137]. Thus, there seems to be a “window” of

favorable adhesion strength for tube formation, where cells are allowed to adhere to the

substrate without being hindered to differentiate, migrate and rearrange into network patterns.

1.4.4 Tube formation in synthetic scaffolds

Cell scaffolds fabricated from synthetic polymers present a different set of challenges for

vascularization compared to reconstituted gels. The identification of parameters crucial to

capillary formation within synthetic materials is still a work in progress, and these may prove

difficult to ascertain in general terms, given the variation in material properties. For “dry”

synthetic polymer scaffolds, the concensus is that sufficient porosity is a prerequisite for

uniform cellular invasion and long-term survival. Pore morphology may also be crucial to the

ability of synthetic scaffolds to induce cellular invasion[142]. Related to more efficient cell

invasion is the observation that porous implanted matrices become surrounded by loose and

highly vascularized connective tissue, whereas non-porous scaffolds become enveloped by a

dense fibrous capsule[143]. Vascularization around and into the implant has been observed to

be increasingly facilitated with increasing pore size, reaching a maximum at pore sizes on the

order of cellular dimensions[143,144]. Moreover, it is believed that in the absence of natural

ECM components, EC must first secrete and organize their own basement membrane in order

to trigger tube formation[84]. Alternatively, surface modification of synthetic polymers may

improve cell-material interactions and facilitate the process. However, as in the case of

biological ECM-based scaffolds, it is believed that intermediate cell adhesion is necessary for

implanted EC to be able to reorganize and actively participate in the vascularization

process[145].

Expanded polytetrafluoroethylene (ePTFE), although not a degradable biomaterial, has

been extensively studied for use in vascular grafts. Investigations of implanted ePTFE arterial

grafts, carried out by Clowes et al.[146], showed that internodal distances of 60µm provoked

the invasion of microvessels through the graft, ultimately leading to complete endothelial cell

coverage of the luminal side of the implants. Internodal distances of 30µm did not allow for

vessel invasion to the same extent, and endothelial coverage was significantly lower.

Supporting results were reported by Saltzmann et al.[147], again indicating an optimum pore

size of 60µm. Interestingly, by modifying the surface with ECM components prior to

Page 36: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

19

implantation in adipose and subcutaneous tissue, extensive vascularization could be stimulated

in similar ePTFE grafts with 30µm internodal distance[72,148]. The implants were modified

either by covalent modification of laminin type I[148] or by exposing the implants to a

preconditioning culture with a cancer cell line, which produced a coating of complex and

hypothetically angiogenic ECM[72]. No vascular infiltration was observed in unmodified

controls.

Degradable polymers are preferred for applications in tissue regeneration. Degradation

mechanisms are sometimes enzymatic[149], but in the case of the commonly used polyesters,

degradation takes place mainly through hydrolytic scission of the ester bonds[150]. This

hydrolytic action results in the production of acidic by-products which, if clearance is not

swift enough, will accumulate and lower the local pH. A degradation-related pH drop has

significant negative influence on cell survival and invasion into the matrix, as was recently

confirmed by Sung and coworkers[47] in a study comparing PLGA and PCL porous scaffolds.

PLGA degrades at a higher rate than PCL, and both polymers degrade faster in vivo than in

vitro. In vivo cell survival and migration into the scaffolds, and thus the extent of

vascularization, was directly affected by increasing acidity of the implants with degradation.

Consequently, PCL implants supported a significantly higher number of newly formed

microvessels[47]. In contrast, upon implantation of porous PLLA and PLGA, Mikos et al.[15]

found that the faster-degrading PLGA was better suited for prevascularization schemes. In this

case, PLGA void volume was filled more rapidly with vascularized fibrous tissue, and not to

same extent as in the PLLA scaffold, and was for both those reasons judged as more suitable

for subsequent cell grafting.

Material design and the method used for scaffold fabrication influence pore size,

morphology and interconnectivity. These characteristics strongly affect cell invasion and must

be considered for the engineering of hard[70]and soft tissue[75,76]as well as for scaffold

vascularization[142]. For example, in PEG-grafted PHEMA (poly(2-hydroxyethyl

methacrylate)) hydrogel sponges seeded with human micro-vascular endothelial cells

(HMVEC), Dziubla et al. observed the formation of tube-like structures with highly variable

lengths and diameters[142]. These highly porous scaffolds were extremely fragile, but

nevertheless allowed for the formation of tubules that were 102 µm long on average, with a

maximum length of 680µm and diameters of 5-10µm, although there was little evidence of

Page 37: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

20

HMVEC beyond the moderate distance of 180µm from the surface[142]. Matrices containing

larger pores (up to 15µm) and higher porosity (up to 90%) supported increasing cell invasion

and tubule formation. However, when the same experiment was conducted with ungrafted

PHEMA sponges, it did not to the same extent result in tube formation. This was proposed to

be the effect of unsuitable pore shapes and lower pore interconnectivity, as well as increased

cell adhesion in comparison with the PEG-modified sponges[142].

Two-dimensional models of micro-patterned polymers have been developed in attempts

to quantify the extent of EC spreading necessary for tube formation. EC seeded on

micropatterned, fibronectin-coated strips, 10-30µm wide, formed cell-cell contacts, spread to

an intermediate degree, and became elongated and aligned in the direction of the strips.

However, cells grown in the 10µm grooves formed lumen-containing tubes, while cells

cultured on strips that were 30µm wide continued to proliferate without any tube formation

taking place[151]. In a related PLGA model, Borenstein et al.[9] used a combination of micro-

machining and replica molding technology to fabricate surfaces containing intricate channel

networks. The resulting materials provide seeded cell populations with an elaborate “map” for

migration and capillary network formation throughout the tissue construct. Specifically, with

the aim of eventually creating patterns that mimic biological vasculature, the same group

reported that seeded EC were successfully grown within these channels for up to four weeks.

However, this technique was limited to the fabrication of channels no smaller than 30µm

wide. The optimization of cell spreading on synthetic substrates may be associated with the

importance of malleability reported for ECM gels. In support of this notion, Ingber and

Folkman found that FGF-stimulated tube formation occurred when coating concentrations of

fibronectin, gelatin, or collagen type IV were adapted to stimulate intermediate EC

spreading[136].

1.4.5 Co-culture models

Endothelial cells in blood vessels are generally surrounded and supported by either

pericytes (small vessels) or SMC (in larger vessels). Paracrine interactions in vivo between

these cell types are highly regulatory and affect the differentiation of ECs and mural

cells[109] as well as the stability and functionality of blood vessels[85]. Engineered blood

vessels often are immature and unstable[152] due to delays in the recruitment of mural cells

Page 38: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

21

from surrounding host tissue. Therefore, researchers have tried to recreate the stabilizing

effects of mural cells in vitro, resulting in the development of several 3D co-culture assays.

These assays include aortic explant models[153], and EC/mural cell co-culture models in

collagen type I[85,114], Matrigel[106], and agarose[109,122]. Trans-membrane culture has

also been used to study the influence of co-culture on cell migration, proliferation, and protein

expression[154-156].

It has been clearly established that co-culturing SMC with EC stimulates SMC to adopt a

more spindle-shaped morphology, typical of a contractile phenotype[154]. Conversely, mural

cells influence EC phenotype, but the effects of are dependent on the matrix. Co-cultures of

EC and SMC rapidly reorganized into tubular networks on Matrigel-coated dishes, whereas

the same experiment conducted on fibrin- or collagen-coated surfaces merely produced

monolayers[119]. Controversial results have been reported for ECs co-seeded with SMC in

3D collagen gels. Asakawa and Kobayashio observed co-culture-dependent EC tube formation

when cultured without external mitogens for EC in the medium[114]. Korff et al. saw that EC

take on a mature, quiescent phenotype in SMC co-culture, as verified by a down-regulation of

PDGF-B expression, a diminished responsiveness to VEGF stimulation, and reduced

apoptosis within the EC population[109]. Simultaneous stimulation with both VEGF and Ang-

2 was nevertheless able to induce endothelial sprouting. However, this model was based on

co-seeded cellulose microspheres embedded in collagen, which involved different cellular

distribution and direct cell-cell contacts[109]. In a study clearly demonstrating the potential of

co-cultures, Koike and coworkers[85] implanted fibronectin-modified collagen gels, co-

seeded with HUVEC and 10T1/2 precursor cells, and observed formation of functional,

stabilized vasculature. These successfully engineered blood vessels were able to connect to

host circulation and to withstand blood flow[85]. Implantation of gels seeded with HUVEC

alone, however, resulted in formation of noticeably fewer and regressing vessels[85].

Co-cultures have helped produce evidence that mural cell precursor recruitment is

stimulated by EC-derived PDGF-BB and that their subsequent differentiation into SMCs or

pericytes is mediated by TGF-β[106,122]. Levels of TGF-β have moreover been found to be

significantly higher in human arterial EC/SMC co-cultures than monocultures of each cell

type[155]. Thus, these heterotypic cell-cell interactions seem to mutually direct the

phenotypes of ECs and mural cells as well as favor the survival and stabilization of newly

Page 39: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

22

formed blood vessels. Additionally, a related phenomenon was observed in a representation of

the dermis in a composite skin substitute where the presence of ECM-producing fibroblasts

improved EC survival, morphogenesis and tube formation[89].

Co-culture of EC with other cell-types is also the dominating strategy employed in the

engineering of small-diameter vascular grafts, from both biological and synthetic scaffolding

materials[157]. In these constructs, SMC are employed to provide mechanical integrity,

something which is achieved in vivo by the cells themselves, as well as the ECM components

they synthesize (mainly collagens). EC are subsequently seeded onto the luminal side of the

tubular constructs to create a confluent non-thrombogenic monolayer corresponding. Auger

and collegues[158] have gone one step further in their development of an entirely biological

vessel substitute. Working with cell sheets wrapped around a porous mandrel, an SMC sheet

representing the media was co-cultured in direct contact with an outer layer of fibroblasts. The

two cell layers eventually fused following several weeks of culture, and were subsequently

seeded with EC. This resulted in a well-defined multilayer organization similar to native tissue

although circumferential ECM fiber orientation was not observed, and the constructs were less

compliant[157,158].

With respect to co-culture in synthetic matrices, increasing knowledge is building of

such techniques and their potential in biomaterial vascularization. For instance, Wu and

coworkers[92] observed microvascular network formation of EPC-derived EC on PGA/PLLA

only when they were co-seeded with SMC. In another study comparing the performance of

two implanted PCL or PLGA porous membranes (10µm), it was shown that the PCL

membrane better supported survival of seeded SMC[47]. This result was accredited in part to

slower rate of degradation and in part to a higher degree of neovascularization within this

scaffold. In more clinically relevant studies, co- and multi-cellular strategies for

prevascularization of three-dimensional scaffolds have been proposed for the engineering of

liver and muscular tissue[14,159]. Co-seeding of hepatocytes with endothelial cells on porous

PLGA with a tube-like pore morphology (inner diameter: 500µm), resulted in endothelial cells

lining the channels and the hepatocytes attached to the polymer walls[160]. However, these

scaffolds were fabricated by a 3D printing method, and it was later stated that techniques

enabling the incorporation of much smaller features are necessary for the engineering of

capillary networks[10]. In a recent publication by Levenberg et al., the successful in vitro

Page 40: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

23

vascularization of PLLA/PLGA sponges for regeneration of skeletal muscle was reported[14].

The scaffolds, which were fabricated using particulate leaching (pore size: 225-500µm), were

seeded with myoblasts, HUVEC, and embryonic fibroblasts in vitro and maintained in culture

to allow for prevascularization to occur prior to subcutaneous and intra-muscular implantation

in mice. The presence of embryonic fibroblasts strongly promoted vascularization of the

constructs, and these cells were found to differentiate into smooth muscle cells in the co-

culture milieu, confirming their stabilizing role. Following implantation, prevascularized

scaffolds better supported further vascularization, blood perfusion and survival of the

engineered muscle tissue[14].

It can be concluded that a large number of factors influence cell behavior in scaffolds,

many of which are related to intrinsic properties of the matrix material itself. In general,

biologically derived matrices present ECM-mimicking motifs to optimize cell-material

interactions, whereas synthetic materials offer enhanced control of material properties. Still,

despite considerable advances and increased insights into the processes involved in EC

network formation, we are still some distance away from successfully engineering vascular

beds to support 3D tissue growth on a larger scale. Furthermore, the examples discussed

clearly illustrate the relevance of co-culture strategies for tissue engineering. Co-cultures may

not only be used for comprehensive studies of paracrine interactions between different cell

types, but they offer the possibility of more autonomous and self-regulating cell populations in

TE constructs. The co-culture approach may help to overcome various difficulties encountered

when dealing solely with ECs. Cues provided, e.g. by direct or indirect cell-cell contact, a

more balanced supply of growth factors, as well as additional and more qualitatively diverse

ECM deposition are all factors that may allow for stable and functional vascular networks to

form with greater ease, thus meeting a crucial requirement of engineering of functional thick

tissue.

1.5 Bioreactor culture and mechanical conditioning

The term “bioreactor” in the context of tissue engineering is used to describe a dynamic

environment specifically designed to enhance culture conditions for a developing tissue.

Within the confines of the bioreactor, the exchange of gases and nutrients (e.g. CO2, O2, and

glucose) may be continuously monitored by sensors, and augmented by a constant medium

Page 41: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

24

turnover. Moreover, tissue-specific mechanical forces may be applied, such as shear stress,

stretching, and compression. The controlled application of external forces has emerged as an

efficient means of regulating cell phenotype, and may therefore be used to enhance

mechanical strength of TE constructs and to more successfully mimic biological

function[157].

1.5.1 Flow conditioning

Native blood vessel endothelium is constantly subject to fluid shear stress. These flow-

induced forces affect many levels of cell function, including the mechanical stiffness and

orientation of the cell[134,161], cell proliferation, and synthesis and secretion of

proteins[162]. Shear stress affects gene transcription and expression of several signaling

molecules, such as TGF-beta, as well as integrins and adhesion molecules[108]. The native

vessel wall responds to changes in flow in a manner that regulates shear stress and maintains

homeostasis, and it has been proposed that these changes are endothelial-dependent[162,163],

and mediated by bFGF[164]. Consequently, vessels will respond to a decrease in blood flow

rate by decreasing the luminal diameter and increasing wall thickness, and conversely for a

flow rate increase[165,166]. Repeated studies of EC monolayers and endothelium lining the

vessel lumen have shown that endothelial orientation reflects the direction of the

flow[93,162,167]. In regions of high shear stress, endothelial cells are highly elongated owing

to transmission of shear forces through the cytoskeleton, resulting in an alignment of actin

stress fibers[162] and reorganization of focal adhesion contacts[168]. Finally, it is now known

that in many cases the same second messengers that are activated by chemical agonists can

also be stimulated by fluid shear stresses[162].

The role of laminar shear stress in vascular remodeling, is reflected in its enhancing

effects on endothelial network formation in vitro[169]. These effects include an integrin-

mediated increase in cytoskeleton stiffness in reply to increased applied stress[161], and may

be compared to high-stiffness substrates giving rise to additional stress fibers. In a study of

bovine EC cultures on collagen gels, Ueda et al.[169] showed the effects of fluid shear stress

(3 dyne/cm2) on EC migration and the morphology of the forming networks. Following

angiogenic stimulation with bFGF, it could be seen that shear-conditioned cultures exhibited

significantly faster rates of elongation and bifurcation in forming networks as compared to

static cultures.

Page 42: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

25

Laminar shear forces influence both endothelial release of some growth factors, and the

responsiveness to these factors. In fact, some studies suggest the importance of flow-induced

shear stress as being more related to vascular development than the chemical stimulation of

certain growth factors. For example, increased VEGF expression in ECs has been reported

both at high (“4-fold increase”)[170] and low (1.5 dyne/cm2) [171] shear stresses relative to

physiological levels. Endothelial release of PDGF is likewise modulated by flow-induced

stress[108,165,172]. Palumbo and coworkers[173] investigated the chemotactic effect of EC-

conditioned culture media on SMC migration. PDGF-BB expression was enhanced under flow

conditions compared to static culture, but its effects were dependent on the magnitude of

laminar flow. At the higher shear stress level (15 dyne/cm2) EC produced predominantly the

PDGF-BB isoform, whereas low shear stress (5 dyne/cm2) also induced significant PDGF-AA

release. These two isomers bind the same SMC membrane receptor, and whereas PDGF-BB

alone exerted a strong chemotactic effect on SMC, the presence of PDGF-AA inhibited this

effect[173]. These observations support the notion that higher shear stress corresponds to

increased vessel wall thickness, achieved by recruitment of mural cells. Conversely, Redmond

et al.[174] demonstrated that EC subjected to shear stress produced plasminogen activator

inhibitor-1, which is inhibitory to flow-induced SMC migration.

Additionally, shear stress also functions as a regulator of the permeability of the vessel

wall to fluid and macromolecules, mediated via complex endothelial cell signaling pathways.

In vitro studies have shown that permeability can be reversibly increased 4- to 10-fold by

applying shear stresses of 1 dyne/cm2 and 10 dyne/cm2, respectively[175].

1.5.2 Cyclic strain

Cyclic strain is mainly applied to condition cells that play a role in the mechanical

integrity of tissues, such as smooth muscle cells[176] and cardiac fibroblasts[177]. This type

of mechanical stimulation has also been used extensively in the culture of artificial blood

vessels[157,178-180].

In addition to shear stress, native blood vessels are also subject to circumferential and

longitudinal stretch through the distension and relaxation of the vessel during the cardiac

cycle. These forces constitute the key mechanical stimuli of smooth muscle cells in the media,

since these cells are not normally exposed to the shear stresses exerted by blood flow. Studies

based on SMC cultured on elastic substrata or in reconstituted gels, and subjected to cyclic

Page 43: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

26

stretching have demonstrated effects on such diverse cellular functions as SMC

phenotype[181], orientation[178,181,182], proliferation[176], growth-factor release[183,184],

and ECM production and remodelling[176,179,185]. As this research has been focused mainly

on SMC behavior under stretch, much less is known about stretch-induced effects in

endothelial cells. However, stress fiber formation[186] and increased PDGF-B

expression[172] have been reported.

Kanda et al. observed that several cell types tend to organize in a direction perpendicular

to the axis of applied cyclic stretch[182]. The same group also found increased contractile

elements, such as myofilaments, in bovine SMC cultured in a collagen-based scaffold and

subjected to 10% strain amplitude at 1Hz for up to 4 weeks[187]. Niklason et al. observed

corresponding upregulation of contractile markers in SMC grown on tubular PGA mesh[188].

These results indicate the induction of a more contractile SMC phenotype in response to

cyclic mechanical stretching.

Kim et al.[176] studied the effects of 7% stretch at 1Hz on SMC cultured on PGA/PLLA

or collagen sponges, and found strain-stimulated increase in elastin and collagen production as

well as in cell proliferation and alignment. Similar results were found for SMCs grown on

P(LA-co-CL) following 8 weeks of culture in pulsatile flow[189]. However, Kim et al.

noticed that collagen sponges did not evoke these responses when cultured in serum-free

medium, indicating that adhesion molecules such as fibronectin and vitronectin, which may

adhere from serum during culture, were necessary for the translation of stretch of the

underlying substrate into a cellular response[176].

An important aspect in tissue engineering is the common lack of mechanical strength of

the TE construct in comparison to the native tissue being replaced. Cyclic stretching has been

shown to significantly increase the mechanical strength of reconstructed artificial vessels

fabricated from both biologically derived[176] and synthetic materials[188]. It seems likely

that increasing mechanical strength is tightly connected to proliferation, cellular alignment,

phenotype (contractile vs. non-contractile), and ECM production. SMC grown on collagen

type I sponges (7% strain, 1Hz) showed dramatically higher tensile modulus and ultimate

strength after 20 weeks of culture compared to static controls[176]. At this time a clear

increase in cellular alignment on the scaffold surface could also be observed. In a series of

investigations carried out by Seliktar et al., SMC-seeded collagen gels placed under 10%

Page 44: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

27

strain at 1Hz exhibited increased gel contraction and mechanical strength as compared to

static controls over 8 days[178]. Furthermore, the capacity of cells to remodel collagen

scaffolds was shown to be affected by mechanical stretching, through increasing expression of

matrix metalloproteinases (MMP)[179,180]. Modulation of MMP expression and/or cellular

responsiveness is important, especially in ECM–based scaffolds, to avoid excessive

degradation of the scaffolds and subsequent loss of mechanical integrity. These investigations,

which were performed using aortic SMC from rat or human, and human dermal fibroblasts,

showed great inconsistency in cell response between cell types, with rat SMC responding

more clearly to cyclic stretch and mediating the greatest increase in scaffold strength[180].

In conclusion, different mechanical stimuli, such as shear and cyclic stretching, are

capable of evoking a wide range of responses from endothelial and smooth muscle cells. The

results generated through mechanical conditioning are very promising, and these methods

represent tools that should be further explored for the promotion of scaffold vsacularization.

1.6 Conclusions

As knowledge of cellular interactions with various scaffolds, soluble factors, and other

cells, the complexity of artificial systems is developing to slowly approach conditions that will

one day allow for the successful engineering of three-dimensional tissues. However, the

promotion of vascularization in 3D tissue constructs remains the most important obstacle for

the generation of large tissue mass. The notion of prevascularized TE scaffolds seems a highly

promising approach, although such attempts have had little real success.

Currently, biodegradable synthetic scaffolds are preferred to ECM-based ones, mainly

due to the superior mechanical properties and controllable aspects of the former. Nevertheless,

more work towards the generation of blood vessel-like networks in porous synthetic materials

that mimic circulation conditions in vivo is required. For the efficient and functional

vascularization of biomaterials, the incorporation and synchronization of several factors

known to greatly influence vessel formation in vitro must be achieved. As discussed in this

review, these factors are to be found in the inherent characteristics of the chosen biomaterial;

the methods by which the scaffold may be modified; the choice of the endothelial cell source,

and whether these are cultured alone or in co-culture; the stimulation by regulatory molecules;

and the mechanical culture conditions to which the cell/polymer construct is subjected.

Page 45: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

28

Meeting the complex requirements of scaffolds and culture conditions will be critical for the

generation of functional vascular networks, critical to successful tissue generation.

Page 46: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

29

1.7 References

(1) Ilan N, Mahooti S, Madri JA. Distinct signal transduction pathways are utilized during

the tube formation and survival phases of in vitro angiogenesis. J. Cell. Sci. 1998; 111:3621-3631.

(2) Kannan RY, Salacinski HJ, Sales K, Butler P, Seifalian AM. The roles of tissue engineering and vascularisation in the development of micro-vascular networks: a review. Biomaterials 2005; 26:1857-1875.

(3) Folkman J, Damore PA. Blood vessel formation: What is its molecular basis? Cell 1996; 87:1153-1155.

(4) Soker S, Machado M, Atala A. Systems for therapeutic angiogenesis in tissue engineering. World J. Urol. 2000; 18:10-18.

(5) Risau W, Flamme I. Vasculogenesis. Annu. Rev. Cell Dev. Biol. 1995; 11:73-91.

(6) Stock U, Vacanti JP. Tissue engineering: current state and prospects. Annu. Rev. Med. 2001; 52:443-451.

(7) Middleton JC, Tipton AJ. Synthetic biodegradable polymers as orthopedic devices. Biomaterials 2000; 21:2335-2346.

(8) Balasubramani M, Kumar TR, Babu M. Skin substitutes: a review. Burns 2001; 27:534-544.

(9) Borenstein JT, Terai H, King KR, Weinberg EJ, Kaazempur-Mofrad MR, Vacanti JP. Microfabrication technology for vascularized tissue engineering. Biomed. Microdev. 2002; 4:167-175.

(10) Kulig KM, Vacanti JR. Hepatic tissue engineering. Transpl. Immunol. 2004; 12:303-310.

(11) Perets A, Baruch Y, Weisbuch F, Shoshany G, Nuefeld G, Smadar C. Enhancing the vascularization of three-dimensional porous alginate scaffolds by incorporating controlled release basic fibroblast growth factor microspheres. J. Biomed. Mat. Res., Part A 2003; 65A:489-497.

(12) Young D, Greulich K, Weier H. Species-specific in situ hybridation with fluorochrome-labeled DNA probes to study vascularization of human skin grafts on athymic mice. J. Burn Care Rehab. 1996; 17:305-310.

(13) Colton CK. Implantable biohybrid artificial organs. Cell Transplant. 1995; 4:415-436.

(14) Levenberg S, Rouwkema J, Macdonald M, Garfein ES, Kohane DS, Darland DC, et al. Engineering vascularized skeletal muscle tissue. Nat. Biotechnol. 2005; 23:879-884.

Page 47: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

30

(15) Mikos AG, Sarakinos G, Lyman MD, Ingber DE, Vacanti JP, Langer R. Prevascularization of porous biodegradable polymers. Biotechnol. Bioeng. 1993; 42:716-723.

(16) Sakakibara Y, Nishimura K, Tambara K, Yamamoto M, Lu FL, Tabata Y, et al. Prevascularization with gelatin microspheres containing basic fibroblast growth factor enhances the benefits of cardiomyocyte transplantation. J. Thorac. Cardiovasc. Surg. 2002; 124:50-56.

(17) Martin Y, Vermette P. Bioreactors for tissue mass culture: Design, characterization, and recent advances. Biomaterials 2005; 26:7481-7503.

(18) Yannas IV. Tissue regeneration templates based on collagen-glycosaminoglycan copolymers. In Biopolymers II: Advances in biopolymers. Heidelberg, Germany: Springer-Verlag; 1995. p.219-244.

(19) Drury JL, Mooney DJ. Hydrogels for tissue engineering: scaffold design variables and applications. Biomaterials 2003; 24:4337-4351.

(20) Hubbell JA. Materials as morphogenetic guides in tissue engineering. Curr. Opin. Biotechnol. 2003; 14:551-558.

(21) Toman PD, Chisholm G, McMullin H, Gieren LM, Olsen DR, Kovach RJ, et al. Production of recombinant human type I procollagen trimers using a four-gene expression system in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 2000; 275:23303-23309.

(22) Dvorak H, Harvey V, Estrella P, Brown L, McDonaugh J, Dvorak A. Fibrin containing gels induce angiogenesis: implications for tumor stroma generation and wound healing. Lab. Invest. 1987; 57:673-686.

(23) Dvorak H, Nagy J, Berse B, Brown L, Yeu K, Yeo T, et al. Vascular permeability factor, fibrin, and the pathogenesis of tumor stroma formation. Ann. N.Y. Acad. Sci. 1992; 667:101-111.

(24) Xue L, Greisler HP. Angiogenic effect of fibroblast growth factor-1 and vascular endothelial growth factor and their synergism in a novel in vitro quantitative fibrin-based 3-dimensional angiogenesis system. Surgery 2002; 132:259-267.

(25) Mackenzie DJ, Sipe R, Buck D, Burgess W, Hollinger J. Recombinant human acidic fibroblast growth factor and fibrin carrier regenerates bone. Plast. Reconstr. Surg. 2001; 107:989-996.

(26) Tassiopoulos AK, Greisler H. Angiogenic mechanisms of endothelialization of cardiovascular implants: a review of recent investigative strategies. J. Biomat. Sci., Pol. Ed. 2000; 11:1275-1284.

Page 48: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

31

(27) Alexander H, Brunski JB, Cooper S, Hench L, Hergenrother R, Hoffmann A, et al. Classes of materials used in medicine. In: Ratner B, Hoffmann A, Schoen F, Lemons J, editors. Biomaterials Science: an introduction to materials in medicine.San Diego: Academic Press; 1996. p. 37-130.

(28) Yannas IV. Biologically active analogues of the extracellular matrix: Artificial akin and nerves. Angew. Chem. Int. Ed. Engl. 1990; 29:20-35.

(29) Whitworth IH, Brown RA, Dore C, Green CJ, Terenghi G. Orientated mats of fibronectin as a conduit material for use in peripheral-nerve repair. J. Hand Surg.:Brit. Eur. Vol. 1995; 20B:429-436.

(30) Harding SI, Afoke A, Brown RA, MacLeod A, Shamlou PA, Dunnill P. Engineering and cell attachment properties of human fibronectin-fibrinogen scaffolds for use in tissue engineered blood vessels. Bioprocess Biosyst. Eng 2002; 25:53-59.

(31) Pachence J, Kohn J. Biodegradable polymers for tissue engineering. In: Lanza R, Langer R, Chick W, editors. Principles of tissue engineering. 1 ed. Austin, Texas, USA: R.G. Landes Company; 1997. p. 273-293.

(32) Di Martino A, Sittinger M, Risbud MV. Chitosan: A versatile biopolymer for orthopaedic tissue-engineering. Biomaterials 2005; 26:5983-5990.

(33) Prabaharan M, Rodriguez-Perez MA, de Saja JA, Mano JF. Preparation and characterization of poly(L-lactic acid)-chitosan hybrid scaffolds with drug release capability. J. Biomed. Mat. Res., Part B 2007; 81:427-434.

(34) Zhang YF, Cheng XR, Chen Y, Shi B, Chen XH, Xu DX, et al. Three-dimensional nanohydroxyapatite/chitosan scaffolds as potential tissue engineered periodontal tissue. J. Biomat. Appl. 2007; 21:333-349.

(35) Smidsröd O, Skjåk-Braek G. Alginate as immobilization matrix for cells. Trends Biotechnol. 1990; 8:71-78.

(36) Sheridan MH, Shea LD, Peters MC, Mooney DJ. Bioadsorbable polymer scaffolds for tissue engineering capable of sustained growth factor delivery. J. Controlled Release 2000; 64:91-102.

(37) Augst AD, Kong HJ, Mooney DJ. Alginate hydrogels as biomaterials. Macromol. Biosci. 2006; 6:623-633.

(38) Martina M, Hutmacher DW. Biodegradable polymers applied in tissue engineering research: a review. Polym. Int. 2007; 56:145-157.

(39) Lee JW, Gardella JA. Surface perspectives in the biomedical applications of poly(alpha-hydroxy acid)s and their associated copolymers. Analyt. Bioanalyt. Chem. 2002; 373:526-537.

Page 49: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

32

(40) Domb A, Kumar N, Sheskin T, Bentolila A, Slager J, Teomim D. Biodegradable polymers as drug carrier systems. In: Dumitriu S, editor. Polymeric biomaterials. 2 ed. Marcel Dekker, Inc.; 2002. p. 91-121.

(41) Cohn D, Stern T, Gonzalez M, Epstein J. Biodegradable poly(ethylene oxide)/poly(ε-caprolactone) multiblock copolymers. J. Biomed. Mat. Res. 2002; 59:273-281.

(42) Huang MH, Li SM, Vert M. Synthesis and degradation of PLA-PCL-PLA triblock copolymer prepared by successive polymerization of ε-caprolactone and (DL)-lactide. Polymer 2004; 45:8675-8681.

(43) Huang MH, Li SM, Hutmacher DW, Schantz JT, Vacanti CA, Braud C, et al. Degradation and cell culture studies on block copolymers prepared by ring opening polymerization of ε-caprolactone in the presence of poly(ethylene glycol). J. Biomed. Mat. Res., Part A 2004; 69:417-427.

(44) Li SM, Garreau H, Vert M, Petrova T, Manolova N, Rashkov I. Hydrolytic degradation of poly(oxyethylene)-poly-(ε-caprolactone) multiblock copolymers. J. Appl. Pol. Sci. 1998; 68:989-998.

(45) Huang MH, Suming LM, Coudane J, Vert M. Synthesis and characterization of block copolymers of ε-caprolactone and DL-lactide initiated by ethylene glycol or poly(ethylene glycol). Macromol. Chem. Phys. 2003; 204:1994-2001.

(46) Maquet V, Jerome R. Design of macroporous biodegradable polymer scaffolds for cell transplantation. Mat. Sci. Forum 1997; 250:15-42.

(47) Sung HJ, Meredith C, Johnson C, Galis ZS. The effect of scaffold degradation rate on three-dimensional cell growth and angiogenesis. Biomaterials 2004; 25:5735-5742.

(48) Griffith L. Emerging design principles in biomaterials and scaffolds for tissue engineering. Ann. N.Y. Acad. Sci. 2002; 961:83-95.

(49) Kumar N, Langer RS, Domb AJ. Polyanhydrides: an overview. Adv. Drug Delivery Rev. 2002; 54:889-910.

(50) Domb AJ. Synthesis and characterization of biodegradable aromatic anhydride copolymers. Macromolecules 1992; 25:12-17.

(51) Shieh L, Tamada J, Tabata Y, Domb A, Langer R. Drug-release from a new family of biodegradable polyanhydrides. J. Controlled Release 1994; 29:73-82.

(52) Laurencin CT, Pierrejacques HM, Langer R. Toxicology and biocompatibility considerations in the evaluation of polymeric materials for biomedical applications. Clin. Lab. Med. 1990; 10:549-570.

(53) Heller J, Maa YF, Wuthrich P, Ng SY, Duncan R. Recent developments in the synthesis and utilization of poly (ortho esters). J. Controlled Release 1991; 16:3-13.

Page 50: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

33

(54) Wang C, Ge Q, Ting D, Nguyen D, Shen HR, Chen JZ, et al. Molecularly engineered poly(ortho ester) microspheres for enhanced delivery of DNA vaccines. Nat. Mater. 2004; 3:190-196.

(55) Laurencin CT, Norman ME, Elgendy HM, Elamin SF, Allcock HR, Pucher SR, et al. Use of polyphosphazenes for skeletal tissue regeneration. J. Biomed. Mat. Res. 1993; 27:963-973.

(56) Greish YE, Bender JD, Lakshmi S, Brown PW, Allcock HR, Laurencin CT. Composite formation from hydroxyapatite with sodium and potassium salts of polyphosphazene. J. Mater. Sci.: Mater. Med. 2005; 16:613-620.

(57) Ambrosio AMA, Allcock HR, Katti DS, Laurencin CT. Degradable polyphosphazene/poly(alpha-hydroxyester) blends: degradation studies. Biomaterials 2002; 23:1667-1672.

(58) Vermette P, Griesser H, Laroche G, Guidoin R. Biomedical applications of polyurethanes. Georgetown, TX, USA: Landes Biosciences, Eurekah.com; 2001.

(59) Guan JJ, Fujimoto KL, Sacks MS, Wagner WR. Preparation and characterization of highly porous, biodegradable polyurethane scaffolds for soft tissue applications. Biomaterials 2005; 26:3961-3971.

(60) Zhang JY, Beckman EJ, Hu J, Yang GG, Agarwal S, Hollinger JO. Synthesis, biodegradability, and biocompatibility of lysine diisocyanate-glucose polymers. Tissue Eng. 2002; 8:771-785.

(61) Hoffmann A. Hydrogels for biomedical applications. Adv. Drug Delivery Rev. 2002; 43:3-12.

(62) Hutmacher DW. Scaffold design and fabrication technologies for engineering tissues - state of the art and future perspectives. J. Biomater. Sci.: Polym. Ed. 2001; 12:107-124.

(63) Mann BK, Schmedlen RH, West JL. Tethered-TGF-beta increases extracellular matrix production of vascular smooth muscle cells. Biomaterials 2001; 22:439-444.

(64) Schmedlen KH, Masters KS, West JL. Photocrosslinkable polyvinyl alcohol hydrogels that can be modified with cell adhesion peptides for use in tissue engineering. Biomaterials 2002; 23:4325-4332.

(65) Huh KM, Bae YH. Synthesis and characterization of poly(ethylene glycol)/poly(L-lactic acid) alternating multiblock copolymers. Polymer 1999; 40:6147-6155.

(66) Suggs LJ, Shive MS, Garcia CA, Anderson JM, Mikos AG. In vitro cytotoxicity and in vivo biocompatibility of poly(propylene fumarate-co-ethylene glycol) hydrogels. J. Biomed. Mat. Res. 1999; 46:22-32.

Page 51: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

34

(67) Barbato F, La Rotonda MI, Maglio G, Palumbo R, Quaglia F. Biodegradable microspheres of novel segmented poly(ether-ester-amide)s based on poly(ε-caprolactone) for the delivery of bioactive compounds. Biomaterials 2001; 22:1371-1378.

(68) George PM, Lyckman AW, Lavan DA, Hegde A, Leung Y, Avasare R, et al. Fabrication and biocompatibility of polypyrrole implants suitable for neural prosthetics. Biomaterials 2005; 26:3511-3519.

(69) Karageorgiou V, Kaplan D. Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials 2005; 26:5474-5491.

(70) Otsuki B, Takemoto M, Fujibayashi S, Neo M, Kokubo T, Nakamura T. Pore throat size and connectivity determine bone and tissue ingrowth into porous implants: Three-dimensional micro-CT based structural analyses of porous bioactive titanium implants. Biomaterials 2006; 27:5892-5900.

(71) Yannas IV, Lee E, Orgill DP, Skrabut EM, Murphy GF. Synthesis and characterization of a model extracellular-matrix that induces partial regeneration of adult mammalian skin. Proc. Natl. Acad. Sci. U.S.A. 1989; 86:933-937.

(72) Kidd KR, Dal Ponte DB, Kellar RS, Williams SK. A comparative evaluation of the tissue responses associated with polymeric implants in the rat and mouse. J. Biomed. Mater. Res. 2002; 59:682-689.

(73) Whang K, Healy KE, Elenz DR, Nam EK, Tsai DC, Thomas CH, et al. Engineering bone regeneration with bioabsorbable scaffolds with novel microarchitecture. Tissue Eng. 1999;.5:35-51.

(74) van Tienen TG, Heijkants RGJC, Buma P, de Groot JH, Pennings AJ, Veth RPH. Tissue ingrowth polymers and degradation of two biodegradable porous with different porosities and pore sizes. Biomaterials 2002 Apr;23(8):1731-1738.

(75) Zeltinger J, Sherwood JK, Graham DA, Mueller R, Griffith LG. Effect of pore size and void fraction on cellular adhesion, proliferation, and matrix deposition. Tissue Eng. 2001; 7:557-572.

(76) McGlohorn JB, Holder WD, Grimes LW, Thomas CB, Burg KJL. Evaluation of smooth muscle cell response using two types of porous polylactide scaffolds with differing pore topography. Tissue Eng. 2004; 10:505-514.

(77) Sharkawy A, Klitzman B, Truskey G, Reichert M. Engineering the tissue which encapsulates subcutaneous implants. III. Effective tissue response times. J. Biomed. Mater. Res. 1997; 40:598-605.

(78) Li SH, de Wijn JR, Li JP, Layrolle P, de Groot K. Macroporous biphasic calcium phosphate scaffold with high permeability/porosity ratio. Tissue Eng. 2003; 9:535-548.

Page 52: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

35

(79) Hollister SJ. Porous scaffold design for tissue engineering. Nat. Mater. 2005; 4:518-524.

(80) Sarazin P, Virgilio N, Favis BD. Influence of the porous morphology on the in vitro degradation and mechanical properties of poly(L-lactide) disks. J. Appl. Polym. Sci. 2006 15; 100:1039-1047.

(81) Sachlos E, Czernuszka J. Making tissue engineering scaffolds work. Review on the application of solid freeform fabrication technology to the production of tissue engineering scaffolds. Eur. Cells Mater. 2003;5:29-40.

(82) Auerbach R, Auerbach W, Polakowski I. Assays for angiogenesis: A Review. Pharmac. Ther. 1991;51:1-11.

(83) Cockerill GW, Gamble J, Vadas M. Angiogenesis: models and modulators. Intl. Rev. Cytology 1995;159:113-60.

(84) Folkman J, Haudenschild C. Angiogenesis in vitro. Nature 1980;288:551-6.

(85) Koike N, Fukumura D, Gralla O, Au P, Schechner J, Jain R. Creation of long-lasting blood vessels. Nature 2004;428:138-9.

(86) Campbell NA, Mathieu R. Circulation et échange gazeux. In: Édition du renouveau pédagogique Inc., editor. Biologie. 3 ed. St-Laurent, QC: The Benjamin/Cummings Publishing Company; 1995. p. 818-49.

(87) Vailhe B, Vittet D, Feige JJ. In vitro models of vasculogenesis and angiogenesis. Lab. Invest. 2001; 81:439-452.

(88) Sottile J. Regulation of angiogenesis by extracellular matrix. Biochim. Biophys. Acta, Rev. Cancer 2004; 1654:13-22.

(89) Black AF, Berthod F, L'Heureux N, Germain L, Auger FA. In vitro reconstruction of a human capillary-like network in a tissue-engineered skin equivalent. FASEB J. 1998; 12:1331-1340.

(90) Unger R, Peters K, Wolf M, Motta A, Migliaresi C, Kirkpatrick C. Endothelialization of a non-woven silk fibroin net for use in tissue engineering: growth and gene reulation of human endothelial cells. Biomaterials 2004; 25:5137-5146.

(91) Siegbahn A, Hebbel RP, Gooch KJ. The relative magnitudes of endothelial force generation and matrix stiffness modulate capillary morphogenesis in vitro. Exp. Cell Res. 2004; 297:574-584.

(92) Wu X, Rabkin-Aikawa E, Guleserian K, Perry T, Masuda Y, Sutherland F, et al. Tissue-engingeering microvessels on three-dimensional biodegradable scaffolds using human endothelial progenitor cells. Am. J. Physiol. Heart Circ. Physiol. 2004; 287:H480-H487.

Page 53: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

36

(93) Yamamoto K, Takahashi T, Asahara T, Ohura N, Sokabe T, Kamiya A, et al. Proliferation, differentiation, and tube formation by endothelial progenitor cells in response to shear stress. J. Appl. Phys. 2003; 95:2081-2088.

(94) Saksela O, Moscatelli D, Rifkin DB. The Opposing Effects of Basic Fibroblast Growth-Factor and Transforming Growth-Factor-Beta on the Regulation of Plasminogen-Activator Activity in Capillary Endothelial Cells. J. Cell Biol. 1987; 105:957-963.

(95) Yancopoulos GD, Davis S, Gale NW, Rudge JS, Wiegand SJ, Holash J. Vascular-specific growth factors and blood vessel formation. Nature 2000; 407:242-248.

(96) Neufeld G, Cohen S, Gengrinovitch S, Poltorak Z. Vascular endothelial growth factor (VEGF) and its receptors. FASEB J. 1999; 13:9-22.

(97) Satake S, Kuzuya M, Ramos MA, Kanda S, Iguchi A. Angiogenic stimuli are essential for survival of vascular endothelial cells in three-dimensional collagen lattice. Biochem. Biophys. Res. Commun. 1998; 244:642-646.

(98) Zisch AH, Lutolf MP, Ehrbar M, Raeber GP, Rizzi SC, Davies N, et al. Cell-demanded release of VEGF from synthetic, biointeractive cell-ingrowth matrices for vascularized tissue growth. FASEB J. 2003; 17.

(99) Lafleur MA, Handsley MM, Knauper V, Murphy G, Edwards DR. Endothelial tubulogenesis within fibrin gels specifically requires the activity of membrane-type-matrix metalloproteinases (MT-MMPs). J. Cell Sci. 2002; 115:3427-3438.

(100) Nehls V, Drenckhahn D. A novel, microcarrier-based in-vitro assay for rapid and reliable quantification of 3-dimensional cell-migration and angiogenesis. Microvasc. Res. 1995; 50:311-322.

(101) Ehrbar M, Djonov VG, Schnell C, Tschanz SA, Martiny-Baron G, Schenk U, et al. Cell-demanded liberation of VEGF(121) from fibrin implants induces local and controlled blood vessel growth. Circ. Res. 2004; 94:1124-1132.

(102) Richardson TP, Peters MC, Ennett AB, Mooney DJ. Polymeric system for dual growth factor delivery. Nat. Biotechnol. 2001; 19:1029-1034.

(103) Peters MC, Polverini PJ, Mooney DJ. Engineering vascular networks in porous polymer matrices. J. Biomed. Mater. Res. 2002; 60:668-678.

(104) Presta M, Dell'Era P, Mitola S, Moroni E, Ronca R, Rusnati M. Fibroblast growth factor/fibroblast growth factor receptor system in angiogenesis. Cytokine Growth Factor Rev. 2005; 16:159-178.

(105) Wissink M, Beernink R, Pieper A, Poot A, Engbers G, Beugeling T, et al. Binding and release of basic fibroblast growth factor from heparinized collagen matrices. Biomaterials 2001; 22:2291-2299.

Page 54: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

37

(106) Darland DC, D'Amore PA. TGF-beta is required for the formation of capillary-like structures in three-dimensional cocultures of 10T1/2 and endothelial cells. Angiogenesis 2001; 4:11-20.

(107) Szekanecz Z, Halloran M, Haskell C, Shah M, Polverini PJ, Koch A. Mediators of angiogenesis: The role of cellular adhesion molecules. Trends Glycosci. Glycotechnol. 1999; 11:73-93.

(108) Conway EM, Collen D, Carmeliet P. Molecular mechanisms of blood vessel growth. Cardiovasc. Res. 2001; 49:507-521.

(109) Korff T, Kimmina S, Martiny-Baron G, Augustin HG. Blood vessel maturation in a 3-dimensional spheroidal coculture model: direct contact with smooth muscle cells regulates endothelial cell quiescence and abrogates VEGF responsiveness. FASEB J. 2001; 15:447-457.

(110) Gamble JR, Matthias LJ, Meyer G, Kaur P, Russ G, Faull R, et al. Regulation of in vitro capillary-tube formation by anti-integrin antibodies. J. Cell Biol. 1993; 121:931-943.

(111) Mazue G, Bertolero F, Garofano L, Brughera M, Carminati P. Experience with the preclinical assessment of basic fibroblast growth factor (bFGF). Toxicol. Lett. 1992; 64-65:329-338.

(112) Springer ML, Chen AS, Kraft PE, Bednarski M, Blau HM. VEGF gene delivery to muscle: Potential role for vasculogenesis in adults. Mol. Cell 1998;2:549-558.

(113) Zisch AH, Schenk U, Schense JC, Sakiyama-Elbert SE, Hubbell JA. Covalently conjugated VEGF-fibrin matrices for endothelialization. J. Controlled Rel. 2001;72:101-113.

(114) Asakawa H, Kobayashi T. The effect of coculture with human smooth muscle cells on the proliferation, the IL-1 beta secretion, the PDGF production and tube formation of human aortic endothelial cells. Cell Biochem. Funct. 1999; 17:123-130.

(115) Kleinman HK, McGarvey ML, Liotta LA, Robey PG, Tryggvasson K, Martin GR. Isolation and characterisation of type IV procollagen, laminin, and heparin sulfate proteoglycan from the EHS sarcoma. Biochemistry. 1982; 21:6188-6193.

(116) Jackson CJ, Giles I, Knop A, Nethery A, Schrieber L. Sulfated polysaccharides are required for collagen-induced vascular tube formation. Exp. Cell Res. 1994; 215:294-302.

(117) Chalupowicz DG, Chowdhury ZA, Bach TL, Barsigian C, Martinez J. Fibrin-II induces endothelial-cell capillary-tube formation. J. Cell Biol. 1995; 130:207-215.

Page 55: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

38

(118) Hoying JB, Boswell C, Williams SK. Angiogenic potential of microvessel fragments established in three-dimensional collagen gels. In Vitro Cell. Dev. Biol. 1996; 32:409-419.

(119) Nicosia RF, Ottinetti A. Modulation of microvascular growth and morphogenesis by reconstituted basement-membrane gel in 3-dimensional cultures of rat aorta - A comparative study of angiogenesis in matrigel, collagen, fibrin, and plasma clot. In Vitro Cell. Dev. Biol. 1990; 26:119-128.

(120) Stiffey-Wilusz J, Boice J, Ronan J, Fletcher A, Anderson M. An ex vivo angiogenesis assay utilizing commercial porcine carotidartery: Modification of the rat aortic ring assay. Angiogenesis 2001; 4:3-9.

(121) Haralabopoulos GC, Grant DS, Kleinman HK, Maragoudakis ME. Thrombin promotes endothelial cell alignment in Matrigel in vitro and angiogenesis in vivo. Am. J. Physiol.: Cell Physiol. 1997; 273:C239-C245.

(122) Hirschi KK, Rohovsky SA, D'Amore PA. PDGF, TGF-beta, and heterotypic cell-cell interactions mediate endothelial cell-induced recruitment of 10T1/2 cells and their differentiation to a smooth muscle fate. J. Cell Biol. 1998; 141:805-814.

(123) Fournier N, Doillon CJ. In-vitro effects of extracellular-matrix and growth-factors on endothelial-cell migration and vessel formation. Cells Mater. 1994; 4:399-408.

(124) Hall H, Baechi T, Hubbell JA. Molecular properties of fibrin-based matrices for promotion of angiogenesis in vitro. Microvasc. Res. 2001;62(3):315-26.

(125) Dallabrida SM, De Sousa MA, Farrell DH. Expression of antisense to integrin subunit beta(3) inhibits microvascular endothelial cell capillary tube formation in fibrin. J. Biol. Chem. 2000; 275:32281-32288.

(126) Schneller M, Vuori K, Ruoslahti E. alpha v beta 3 integrin associates with activated insulin and PDGF beta receptors and potentiates the biological activity of PDGF. Embo J. 1997; 16:5600-5607.

(127) Soldi R, Mitola S, Strasly M, Defilippi P, Tarone G, Bussolino F. Role of alpha(v)beta(3) integrin in the activation of vascular endothelial growth factor receptor-2. Embo J. 1999; 18:882-892.

(128) Yang S, Graham J, Kahn JW, Schwartz EA, Gerritsen ME. Functional roles for PECAM-1 (CD31) and VE-cadherin (CD144) in tube assembly and lumen formation in three-dimensional collagen gels. Am. J. Pathol. 1999; 155:887-895.

(129) Bach TL, Barsigian C, Chalupowicz DG, Busler D, Yaen CH, Grant DS, et al. VE-cadherin mediates endothelial cell capillary tube formation in fibrin and collagen gels. Exp. Cell Res. 1998; 238:324-334.

Page 56: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

39

(130) Nillesen STM, Geutjes PJ, Wismans R, Schalkwijk J, Daamen WF, Van Kuppevelt TH. Increased angiogenesis and blood vessel maturation in acellular collagen-heparin scaffolds containing both FGF2 and VEGF. Biomaterials 2007; 28:1123-1131.

(131) Tanihara M, Suzuki Y, Yamamoto E, Noguchi A, Mizushima Y. Sustained release of basic fibroblast growth factor and angiogenesis in a novel covalently crosslinked gel of heparin and alginate. J. Biomed. Mater. Res. 2001; 56:216-221.

(132) Vernon R, Angello J, Iruela-Arispe L, Lane T, Sage H. Reorganization of basement membrane matrices by cellular traction promotes the formation of cellular networks in

vitro. Lab. Invest. 1992; 66:536-547.

(133) Yamada KM, Geiger B. Molecular interactions in cell adhesion complexes. Curr. Opin. Cell Biol. 1997; 9:76-85.

(134) Wang N, Butler J, Ingber D. Mechanotransduction across the cell surface and through the cytoskeleton. Science 1993; 260:1124-1127.

(135) Maheswari G, Brown G, Lauffenburger D, Wells A, Griffith L. Cell adhesion and motility depend on nanoscale RGD clustering. J. Cell Sci. 2000; 113:1677-1686.

(136) Ingber DE, Folkman J. Mechanochemical switching between growth and differentiation during fibroblast growth factor-stimulated angiogenesis in vitro: Role of extracellular matrix. J. Cell Biol. 1989; 109:317-330.

(137) Deroanne CF, Lapiere CM, Nusgens BV. In vitro tubulogenesis of endothelial cells by relaxation of the coupling extracellular matrix-cytoskeleton. Cardiovasc. Res. 2001; 49:647-658.

(138) Davis GE, Allen KAP, Salazar R, Maxwell SA. Matrix metalloproteinase-1 and -9 activation by plasmin regulates a novel endothelial cell-mediated mechanism of collagen gel contraction and capillary tube regression in three-dimensional collagen matrices. J. Cell Sci. 2001; 114:917-930.

(139) Numaguchi Y, Huang S, Polte T, Eichler G, Wang N, Ingber D. Caldesmoson-dependent switching between capillary endothelial cell growth and apoptosis through modulation of cell shape and contractility. Angiogenesis 2003; 6:55-64.

(140) Bayless KJ, Davis GE. Microtubule depolymerization rapidly collapses capillary tube networks in vitro and angiogenic vessels in vivo through the small GTPase Rho. J. Biol. Chem. 2004; 279:11686-11695.

(141) Kanzawa S, Endo H, Shioya N. Improved in vitro angiogenesis model by collagen density reduction and the use of type-III collagen. Ann. Plast. Surg. 1993; 30:244-251.

(142) Dziubla TD, Lowman AM. Vascularization of PEG-grafted macroporous hydrogel sponges: A three-dimensional in vitro angiogenesis model using human microvascular endothelial cells. J. Biomed. Mater. Res., Part A 2004; 68:603-614.

Page 57: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

40

(143) Sharkawy A, Klitzman B, Truskey G, Reichert M. Engineering the tissue which encapsulates subcutaneous implants. II. Plasma-tissue exchange properties. J. Biomed. Mater. Res. 1997; 40:586-597.

(144) Sharkawy A, Klitzman B, Truskey G, Reichert M. Engineering the tissue which encapsulates subcutaneous implants. I. Diffusion properties. J. Biomed. Mater. Res. 1996; 37:401-412.

(145) Sieminski A, Gooch KJ. Biomaterials-microvascular interactions. Biomaterials 2000; 21:2233-2241.

(146) Clowes AW, Kirkman TR, Reidy MA. Mechanisms of arterial graft healing - rapid transmural capillary ingrowth provides a source of intimal endothelium and smooth-muscle in porous PTFE prostheses. Am. J. Pathol. 1986; 123:220-230.

(147) Salzmann DL, Kleinert LB, Berman SS, Williams SK. The effects of porosity on endothelialization of ePTFE implanted in subcutaneous and adipose tissue. Journal J. Biomed. Mater. Res. 1997; 34:463-476.

(148) Williams SK, Kleinert LB, Hagen KM, Clapper DL. Covalent modification of porous implants using extracellular matrix proteins to accelerate neovascularization. J. Biomed. Mater. Res., Part A 2006; 78:59-65.

(149) Li SM, Molina I, Martinez MB, Vert M. Hydrolytic and enzymatic degradations of physically crosslinked hydrogels prepared from PLA/PEO/PLA triblock copolymers. J. Mater. Sci.: Mater. Med. 2002; 13:81-86.

(150) Li S, Anjard S, Rashkov I, Vert M. Hydrolytic degradation of PLA/PEO/PLA triblock copolymers prepared in the presence of Zn metal or CaH2. Polymer 1997; 39:5421-5430.

(151) Dike LE, Chen CS, Mrksich M, Tien J, Whitesides GM, Ingber DE. Geometric control of switching between growth, apoptosis, and differentiation during angiogenesis using micropatterned substrates. In Vitro Cell. Dev. Biol. Anim. 1999; 35:441-448.

(152) Jain R. Molecular regulation of vessel maturation. Nat. Med. 2003;.9:685-693.

(153) Nicosia RF, Ottinetti A. Growth of microvessels in serum-free matrix culture of rat aorta. Lab. Invest. 1990; 63:115-122.

(154) Powell RJ, Cronenwett JL, Fillinger MF, Wagner RJ, Sampson LN. Endothelial cell modulation of smooth muscle cell morphology and organizational growth pattern. Ann. Vasc. Surg. 1996; 10:4-10.

(155) Axel D, Riessen R, Athanasiadis A, Runge H, Koveker G, Karsch K. Growth factor expression of human arterial smooth muscle cells and endothelial cells in a transfilter coculture system. J. Mol. Cell Cardiol. 1997; 29:2967-2978.

Page 58: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

41

(156) Axel DI, Brehm BR, WolburgBuchholz K, Betz EL, Koveker G, Karsch KR. Induction of cell-rich and lipid-rich plaques in a transfilter coculture system with human vascular cells. J. Vasc. Res. 1996;.33:327-339.

(157) Isenberg BC, Williams C, Tranquillo RT. Small-diameter artificial arteries engineered in vitro. Circ. Res. 2006; 98:25-35.

(158) L'Heureux N, Pâquet S, Labré R, Germain L, Auger FA. A completely biological tissue-engineering human blood vessel. FASEB J. 2004; 12:47-56.

(159) Marler JJ, Upton J, Langer R, Vacanti JP. Transplantation of cells in matrices for tissue regeneration. Adv. Drug Delivery Rev. 1998; 33:165-182.

(160) Chaignaud B, Wu B, Lee HB, Utsunomiya H, Cima LG, Vacanti JP. Complex three-dimensional tissues with a vascular network using a new technology of solid free form fabrication. Am. Ped. Surgical Assoc. Annual Meeting, San Diego. 1996.

(161) Ingber D. Tensegrity: The architectural basis of cellular mechanotransduction. Annu. Rev. Physiol. 1997; 59:575-599.

(162) Ziegler T, Nerem RM. Tissue engineering of a blood vessel: Regulation of vascular biology by mechanical stress. J. Cell. Biochem. 1994; 56:204-209.

(163) Langille B. Remodelling of developing and mature arteries: endothelium, smooth muscle cells, and matrix. J. Cardiovasc. Pharmacol. 1993;21:S11-S17.

(164) Shane R, Bjercke R, Erichsen D, Rege A, Lindner V. Vascuar remodeling in response to altered blood flow is mediated by fibroblast growth factor-2. Circ. Res. 1999; 84:323-328.

(165) Aromatario C, Sterpetti A, Palumbo R, Patrizi A, Di Carlo A, Proietti P, et al. Fluid shear stress increases the release of platelet derived growth factor BB (PDGF BB) by aortic endothelial cells. Minerva Cardioangiol. 1997; 45:1-7.

(166) Greenwald SE, Berry CL. Improving vascular grafts: the importance of mechanical and haemodynamic properties. J. Pathol. 2000; 190:292-299.

(167) Nerem RM, Levesque MJ, Cornhill JF. Vascular endothelial morphology as an indicator of blood flow. ASME J. Biomech. Eng. 2004; 103:172-176.

(168) Chicurel M, Chen C, Ingber D. Cellular Control lies in the balance of forces. Curr. Opin. Cell Biol. 1998; 10:232-239.

(169) Ueda A, Koga M, Ikeda M, Kudo S, Tanishita K. Effect of shear stress on microvessel network formation of endothelial cells with in vitro three-dimensional model. Am. J. Physiol. Heart. Circ. Physiol. 2004; 287:H994-H1002.

Page 59: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

42

(170) Milkiewicz M, Brown MD, Egginton S, Hudlicka O. Association between shear stress, angiogenesis, and VEGF in skeletal muscles in vivo. Microcirculation 2001; 8:229-241.

(171) Conklin BS, Zhong DS, Zhao WD, Lin PH, Chen CY. Shear stress regulates occludin and VEGF expression in porcine arterial endothelial cells. J. Surg. Res. 2002; 102:13-21.

(172) Sumpio B, Du W, Galagher G, Wang X, Khashigian M, Collins T, et al. Regulation of PDGF-B in endothelial cells exposed to cyclic strain. Arterioscler.,Thromb., Vasc. Biol. 1998; 18:349-355.

(173) Palumbo R, Gaetano C, Antonini A, Pompilio G, Bracco E, Rönnstrand L, et al. Different effects of high and low shear stress on platelet-derived growth factor isoform release by endothelial cells. Arterioscler.,Thromb., Vasc. Biol. 2002; 22:405-411.

(174) Redmond E, Cullen J, Cahill P, Sitzmann J, Stefansson S, Lawrence D, et al. Endothelial cells inhibit flow-induced smooth muscle cell migration: Role of plasminogen activator inhibitor-1. Circulation 2001; 103:597-603.

(175) Jo H, Dull RO, Hollis TM, Tarbell JM. Endothelial albumin permeability is shear dependent, time-dependent, and reversible. Am. J. Physiol. 1991; 260:H1992-H1996.

(176) Kim BS, Nikolovski J, Bonadio J, Mooney DJ. Cyclic mechanical strain regulates the development of engineered smooth muscle tissue. Nat. Biotechnol. 1999; 17:979-983.

(177) Wang J, Chen H, Seth A, McCulloch CA. Mechanical force regulation of myofibroblast differentiation in cardiac fibroblasts. Am. J. Physiol. Heart. Circ. Physiol. 2003; 285:H1871-H1881.

(178) Seliktar D, Black RA, Vito RP, Nerem RM. Dynamic mechanical conditioning of collagen-gel blood vessel constructs induces remodeling in vitro. Annals Ann. Biomed. Eng. 2000; 28:351-362.

(179) Seliktar D, Nerem RM, Galis ZS. The role of matrix metalloproteinase-2 in the remodeling of cell-seeded vascular constructs subjected to cyclic strain. Ann. Biomed. Eng. 2001; 29:923-934.

(180) Seliktar D, Nerem RM, Galis ZS. Mechanical strain-stimulated remodeling of tissue-engineered blood vessel constructs. Tissue Eng. 2003; 9:657-666.

(181) Cha JM, Park TN, Noh TH, Suh T. Time-dependent modulation of alignment and differentiation of smooth muscle cells seeded on a porous substrate undergoing cyclic mechanical strain. Artificial Organs 2006; 30:250-258.

(182) Kanda K, Matsuda T. Behavior of arterial-wall cells cultured on periodically stretched substrates. Cell Transplant. 1993; 2:475-484.

Page 60: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

43

(183) Cheng GC, Briggs WH, Gerson DS, Libby P, Grodzinsky AJ, Gray ML, et al. Mechanical strain tightly controls fibroblast growth factor-2 release from cultured human vascular smooth muscle cells. Circ. Res. 199; 80:28-36.

(184) Wilson E, Mai Q, Sudhir K, Weiss RH, Ives HE. Mechanical strain induces growth of vascular smooth-muscle cells via autocrine action of PDGF. J. Cell Biol. 1993; 123:741-747.

(185) Sumpio BE, Banes AJ, Link WG, Johnson G. Enhanced collagen production by smooth-muscle cells during repetitive mechanical stretching. Arch. Surg. 1988; 123:1233-1236.

(186) Luo SS, Sugimoto K, Fujii S, Takemasa T, Fu SB, Yamashita K. Role of heat shock protein 70 in induction of stress fiber formation in rat arterial endothelial cells in response to stretch stress. Acta Histochem. Cytochem. 2007; 40:9-17.

(187) Kanda K, Matsuda T, Oka T. Mechanical stress induced cellular orientation and phenotypic modulation of 3-D cultured smooth muscle cells. ASAIO J 1993; 39:M690.

(188) Niklason LE, Gao J, Abbott WM, Hirschi KK, Houser S, Marini R, et al. Functional arteries grown in vitro. Science 1999; 284:489-493.

(189) Jeong SI, Kwon JH, Lim JI, Cho SW, Jung YM, Sung WJ, et al. Mechano-active tissue engineering of vascular smooth muscle using pulsatile perfusion bioreactors and elastic PLCL scaffolds. Biomaterials 2005; 26:1405-1411.

Page 61: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

44

PREFACE TO CHAPTER 2. As a first step towards the fabrication of a three-

dimensional scaffold, it was judged important to have a series of biocompatible

polymers available, whose material properties could be fine-tuned to respond to

the demands of a specific application. Among these properties are mechanical

properties and the rate of degradation. The compositions chosen, the synthesis

and characterization of those polymeric materials are described in this chapter.

Page 62: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

45

CHAPTER 2:

CHARACTERISATION, DEGRADATION AND MECHANICAL STRENGTH OF POLY(D,L-LACTIDE-CO-ε-CAPROLACTONE)-POLY(ETHYLENE GLYCOL)-POLY(D,L-LACTIDE-

CO-ε-CAPROLACTONE)

2.1 Abstract

A series of three biocompatible P(CL-co-LA)-PEG-P(CL-co-LA) copolymers were

synthesized using ring-opening polymerization and characterized by 1H-NMR, GPC, DSC,

DMA, and X-ray diffraction. The number of monomer units was kept constant, while the D,L-

LA fraction was varied so as to constitute 0, 30, or 70% of the end segments. The molecular

weights were sufficiently high to eventually permit 3D scaffold preparation. A degradation

study was carried out over 26 weeks, and the effect of monomer composition on the rate of

degradation as well as on changes in mechanical strength, was investigated. Pure PCL-PEG-

PCL copolymer, P(100/0), was a crystalline material displaying no measurable mass loss, a

30% reduction in mean molecular weight (Mn), and only very slight changes in tensile

strength. The random incorporation of 30% and 70% D,L-LA, into the end sections of the

polymer chain, produced more and more amorphous materials, exhibiting increasingly high

rates of degradation, mass loss and loss of tensile strength. Compared to random P(CL-co-

LA), the presence of the PEG block was found to improve hydrophilicity and thus the rate of

degradation. The tested copolymers range from materials exhibiting low mechanical strength

and high rate of degradation to slow-degrading materials with high mechanical strength

suitable, e.g., for three-dimensional scaffolding.

2.2 Introduction

Among the polymers most intensively studied for biomedical applications in the past

decade are polycaprolactone (PCL) and poly(lactic acids) (PLA). These resorbable polymers

belong to the group of aliphatic polyesters, and their copolymers present a wide range of

tuneable physicochemical properties depending on molecular weight, molecular structure and

monomer composition[1-3]. Thus, material properties may be tailored to suit a particular

application, with respect to e.g., mechanical strength, degradability, and drug-release profile.

Some characteristics set PCL apart from other aliphatic polyesters. Its low glass-transition

Page 63: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

46

temperature (-60°C), makes it flexible at physiological temperature and contributes to its high

permeability for many therapeutic drugs[4]. Poly(lactic acid) is available in two isoforms,

namely the crystalline PLLA and the amorphous PDLLA. Whereas the hydrolysis of PLLA is

a very slow process, PDLLA will degrade in a matter of weeks, and even the inclusion of

very small amounts of D,L-LA in the polymer chain will dramatically accelerate

degradation[5,6]. In a similar manner it is possible to control the rate of degradation of the

highly crystalline and hydrophobic PCL by varying the CL-to-D,L-LA monomer ratio of the

corresponding co-polymer. Studied separately as well as in various copolymer formulations

and blends, CL/LA copolymers have proved useful in a number of biomedical applications

including surgical sutures[7], controlled-release drug delivery[8], and nerve guide

channels[9]. However, PDLLA-PCL-PDLLA triblock copolymers have been known to

exhibit burst-like degradation behaviour during hydrolysis[10], due to poor initial water

absorption followed by removal of the rapidly degrading PDLLA blocks. Such features

would be highly undesirable in a polymer implant that is designed to degrade in a controlled

manner in order to be gradually replaced by invading tissues. Random P(CL-co-DLLA)

copolymers, however, show stable but slow water uptake and degradation profiles[11].

Poly(ethylene glycol) (PEG) is a polymer well known for its low-fowling properties,

biocompatibility and solubility both in water and organic solvents. PEG is not susceptible to

hydrolysis, although its incorporation into the polymer backbone has been shown to enhance

the rate of degradation by increasing hydrophilicity and, thus, water uptake[12]. The effect of

PEG on the degradation of amorphous PDLLA blocks is reportedly modest[13]. In contrast,

increasing the weight fraction of PEG from 23% to 62% in PEG-PCL multiblock polymers

increased mass loss due to hydrolysis from 20% to 96%, respectively, in a study conducted

over 70 weeks[14]. PEG-containing co-polymers have been proposed for such applications as

tissue engineering[15] and drug-delivery systems[16,17]. For example, successful bone

formation was observed in vivo during a three-week study of a cytokine-liberating PDLLA-

PEG-PDLLA scaffold[16].

In the present study we chose to combine these two measures to generate P(CL-co-

DLLA)-PEG-P(CL-co-DLLA) copolymers of sufficiently high molecular weights for these

materials to be considered for scaffold fabrication. In such a ternary copolymer,

hydrophilicity is improved by including a PEG block in the polymer backbone, while at the

Page 64: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

47

same time using a random distribution of CL and DLLA units rather than a pure block

structure, producing a more amorphous structure, whose degradation profile should be easier

to control. Importantly, although some previous reports on this type of copolymer can be

found in literature[18,19], previously investigated materials all had modest molecular weights

(MW = 2 500 – 13 500 Da) and were designed for drug-delivery applications[19]. These

materials would therefore be unsuitable for use in large, three-dimensional tissue engineering

scaffolds, which require good mechanical integrity – not only initially but also during the

course of degradation. Here, we describe and evaluate the influence of the monomer

composition (CL-to-DLLA ratio) on various aspects of hydrolytic degradation and

crystallinity. Additionally, the influence of degradation on the tensile strength of the three

test materials are investigated, which has not yet been done for this type of polymer.

2.3 Materials and Methods

2.3.1 Materials

PEG diols (MW=7500, cat. # 06103) and D,L-lactide (D,L-LA, cat. # 16640) were

purchased from Polysciences Inc. (Warrington, PA, USA). PEG diols were vacuum-dried at

least 48h at 50°C and D,L-LA was recrystallized from ethyl acetate prior to use. Stannous

octoate (SnOct, cat. # S3252) and ε-caprolactone (ε-CL, cat. # 21510) were purchased from

SigmaAldrich (Oakville, ON, Canada) and used as received. Penicillin and streptomycin were

purchased from Invitrogen (Burlington, ON, Canada; cat. # 15140-122).

2.3.2 Polymer synthesis

All copolymers were synthesized according to Bogdanov et al.[20] by SnOct-catalyzed

ring-opening polymerization of ε-CL and D,L-LA in the presence of PEG diols. The polymers

were named P(x/y) according to the molar percent content of ε-CL (x) and D,L-LA (y) in the

end segments. Briefly, for the synthesis of P(100/0), PEG diols and ε-CL were introduced

into a three-neck round-bottom flask and melted at 140°C. The molar ratio of EG-to-

monomer was 1:4.2. The temperature was then lowered to 120°C and the melted mixture

degassed and purged with N2. Stannous octoate was added at a monomer-to-catalyst ratio of

1000:1, and the polymerization was allowed to continue for 24h at 120°C under constant

stirring.

Page 65: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

48

To obtain P(70/30) and P(30/70) copolymers, PEG diols and monomers were introduced

into a three-neck round-bottom flask at a molar ratio of monomer:EG of 4.2:1. The molar

ratio of ε-caprolactone:D,L-lactide was either 70:30 or 30:70. Following degassing and N2-

purging at 90°C, the temperature was briefly raised to 140°C to complete the melting of D,L-

LA and subsequently lowered to 120°C at which temperature polymerization was carried out

for 24h under constant stirring.

All polymers were dissolved in dichloromethane and precipitated in cold ethanol,

followed by filtration and drying in a vacuum oven at 30°C until constant weight.

2.3.3 Film preparation

Films were fabricated between Teflon sheets in a Carver Laboratory hot melt press at

150°C, using a stainless steel mold. The entire mold was subsequently cooled under running

water. The obtained film thickness was either 0.25mm or 0.75mm. Following drying of the

films, 0.75-mm thick samples were punched out using a 1-cm diameter punch, to be used for

the degradation study. Rectangular shapes (10x60mm) were cut from the 0.25-mm films for

tensile testing.

2.3.4 Characterization

The polymers were characterized by 1H-NMR (Brüker AC-300F, 300.13MHz) using

deuterated chloroform as solvent.

Gel permeation chromatography (GPC) was performed with THF as the mobile phase and

0.5ml/min flow rate, using a Waters system equipped with a differential refractometer. The

system was calibrated with polystyrene standards.

Differential scanning calorimetry (Perkin-Elmer DSC7) measurements were carried out

on 0.75-mm film samples over a temperature range of -80°C to 100°C at a rate of 10°C/min.

The apparatus was calibrated with both Hg and In. Thermal transitions were identified as the

peak values of the enthalpy changes. The percentage of CL that was able to crystallize within

the copolymer films was calculated from the thermograms, using the weight fractions of PCL

based on 1H-NMR results. The crystallinity (Xc) of the CL fraction was then calculated as:

Xc = (∆Hexo + ∆Hendo)/∆HPCL (2.1)

Page 66: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

49

where ∆Hexo and ∆Hendo, respectively, are sample melting and crystallizing enthalpies of

the same heating run, and ∆HPCL is the reference heat of melting of pure PCL. As glass

transition temperatures (Tg) were not clearly visible in the DSC thermograms, they were

identified by dynamic-mechanical analysis (DMA), performed on a Q800 model from TA

Instruments. Measurements were carried out from -80°C to 10°C at a heating rate of

5°C/min, using the three-point flexion clamp at a frequency of 1Hz.

Wide-angle X-ray diffraction (WAXD) spectra were recorded in a Panalytical X’pert Pro

MPD (Almelo, The Netherlands) apparatus using a Cu Kα radiation source (45 kV, 40 mA)

and a scan rate of 0.44°/min.

2.3.5 Hydrolytic degradation

Polymer degradation by hydrolysis was studied by incubating the 0.75-mm-thick samples

(1-cm diameter) in a buffer solution for up to 26 weeks. Weighed samples were sterilized by

soaking in 70% ethanol for 30 minutes. Following rinsing in sterile water, samples were

immersed in 2mL PBS at pH 7.4 and 37°C. The buffer was supplemented with 100U/mL

penicillin and 100mg/mL streptomycin to prevent bacterial growth. The medium was

changed every 3 weeks for the first 6 weeks, every 2 weeks for the following three months,

and from week 18 the medium was changed every week until the end of the study. At

predetermined intervals (week 3, 6, 10, 14, 18, and 26; n=3) samples were removed from the

degradation medium, washed 3x5min in deionized water and gently blotted on filter paper

before being weighed to determine PBS absorption. Samples were then dried under vacuum,

weighed again, and prepared for analysis by 1H-NMR and GPC. Mass loss due to degradation

was determined from Equation 2.2, and the absorption of degradation medium from Equation

2.3:

Mass loss (%) = 100*(w0-wt)/w0 (2.2)

Absorption (%) = 100*(wwet-wt)/wt , (2.3)

where w0 is the initial weight of the sample, wt is the mass of the degraded sample after

drying, and wwet is the wet weight of a retrieved sample.

Page 67: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

50

2.3.6 Tensile testing

Mechanical testing to determine tensile strength of initial (dry) and partially degraded

(wet) films was carried out according to the ASTM 882-02 standard at room temperature.

Five samples were evaluated for each material and condition. All measurements were carried

out on an Instron 5565 using a 50-N load cell and a cross-head speed of 40 mm/min. Partially

degraded films were tested while wet. Pure PCL films (CAPA 6500, Solvay Caprolactones,

Houston, USA; MW=50 kDa) were used as a reference material.

2.4 Results and Discussion

2.4.1 Characterization of PEG-containing co-polymers

Polymers prepared by the method described were characterized by 1H-NMR and GPC to

obtain the monomer composition, the molecular weight and the polydispersity index. The

results of these analyses showed a good agreement between the monomer ratios in the feed

and in the final product (Table 2.1). However, some loss of monomer in the P(30/70) material

was observed during synthesis, although the CL-to-LA ratio remained precisely the same as

in the feed. Additionally, there was a small quantity of unreacted PEG and D,L-LA during the

polymerization of P(100/0) and P(70/30), respectively. The incomplete incorporation of D,L-

LA has already been reported[1]. Moreover, we may note that the deviations between

monomer ratios in the feed and in the final polymers are well in line with the differences

between the theoretical molecular weights and those derived from 1H-NMR spectroscopy.

Finally, polydispersity indices (PI) of 1.7 were found for all of the synthesized polymers.

Table 2.1: Characterization of the three copolymers synthesized by ring-opening

polymerization.

Monomer content Molecular weight (Mn)

Polymer Feed

(CL+LA):EG Finala

Feed CL:LA

Finala Theoretical 1H-NMR GPC PIb

P(100/0) 4.2:1 4.5:1 88 900 94 800 60 900 1.7

P(70/30) 4.2:1 4.2:1 70:30 74:26 79 900 81 100 46 900 1.7

P(30/70) 4.2:1 3.1:1 30:70 30:70 67 900 52 400 35 500 1.7

Values derived from a) 1H-NMR spectra and b) GPC measurements. The thermodynamic properties were characterized by DSC and DMA, and the values of

the transition temperatures for the copolymer films are presented in Table 2.2. DMA

Page 68: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

51

measurements revealed significantly increasing glass transition temperatures (Tg) with

increasing D,L-LA content. The single melting peak (Tm) of the first heating of P(100/0) was

found at 57°C, close to the Tm of a pure PCL. A single crystallization peak was also recorded

at 26°C, showing that P(100/0) crystallizes at temperatures close to room temperature.

P(30/70) only exhibited one melting transition, during the first heating (Tm = 40°C), and was

then unable to recrystallize under the experimental conditions. A similar absence of

crystallinity in P(CL-co-D,L-LA) of comparable composition has been previously

reported[21]. Nevertheless, the clearly semi-crystalline state of the melt- pressed film

signifies that this material crystallized either during the cooling of the mold or during storage

at room temperature.

Table 2.2: Thermodynamic transitions, fusion energy, and % of crystallinity

of P(100/0), P(70/30), and P(30/70) films. 1st heatinga 1st coolinga 2nd heatinga

Polymer Tg

b

(°C) Tm

(°C)

∆Hm

(J/g)

Tc

(°C)

∆Hc

(J/g)

Tc

(°C)

∆Hc

(J/g)

Tm

(°C)

∆Hm

(J/g)

Xc

(%CL)

P(100/0) -59 57 73 26 51 53 46 43

P(70/30) -45 40/43 39 -9 26 -22 5 40/44 39 38

P(30/70) -21 40 18 0

Data obtained from: a: DSC; b: DMA, using the peak of the loss modulus (E’’). Xc was calculated from the

2nd heating.

Furthermore, in P(70/30) but not in P(30/70), phase separation took place as indicated by

the presence of a bimodal melting peak (Tm = 40/43°C) in the DSC thermogram of the

former. Although PCL is a highly crystalline polymer, PEG blocks are sometimes able to

crystallize within copolymers provided that the block length is sufficiently large, and that

surrounding components do not hinder their crystallization[20,22]. In this case, however, the

weight fraction of PEG blocks for either material does not exceed 15%, which, according to

Gan et al.[23], is insufficient for PEG to crystallize in PEG-PCL diblock copolymers. Thus, it

was hypothesized that the bimodal melting endotherm reflected different crystallization

conditions of the PCL fraction, as previously described[23], according to e.g. their position

relative to surrounding blocks and the presence of impurities. P(70/30) exhibited

crystallization peaks (Tc) both during heating (Tc = -22°C) and cooling (Tc = -9°C), where the

Page 69: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

52

latter is in agreement with values reported for short CL oligomers[18]. P(70/30) and P(30/70)

both exhibited lower Tm and melting enthalpies (∆Hm) compared to P(100/0). This was

expected seeing as the CL fraction in the polymer chains is smaller, but may also reflect the

influence from the amorphous D,L-LA, which strongly increase the mobility of the chains and

thereby their ability to crystallize[2].

To further investigate the nature of copolymer crystallinity, samples were analyzed by

wide-angle X-ray diffraction (WAXD). The WAXD results showed two dominant peaks for

crystalline PCL at 2θ = 21.4 and 23.7°, and two peaks for PEG at 2θ = 19.1 and 23.3° (Fig.

2.1). The spectra of P(100/0) and P(70/30) revealed that crystallinity was due entirely to the

PCL blocks. Since P(70/30) contains relatively less CL than P(100/0), this explains the lower

intensity observed, although the presence of D,L-LA in P(70/30) may also play a role in

reducing crystallinity. Finally, the P(30/70) spectrum reflects a largely amorphous material,

where the intensity of the peak (at 2θ = 21.4°) corresponding to crystalline CL-rich regions is

clearly diminished. It is very interesting to note that PEG is able to crystallize in this material.

2 Thetao16 18 20 22 24 26

Inte

nsi

ty

PEG

PCL

P(70/30)

P(100/0)

P(30/70)

Figure 2.1: WAXD curves of films of PEG-containing copolymers compared to the homo-polymers PEG (powder) and PCL (film).

The percentage of CL that was able to crystallize within the copolymer films was

calculated, using the weight fractions of PCL based on 1H-NMR results and a value of the

fusion enthalpy of 100% crystalline PCL of 135 J/g[25]. For the second heating, Xc of PCL

Page 70: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

53

were 43, 38, and 0% in P(100/0), P(70/30), and P(30/70), respectively. Thus, we can

conclude that the presence of D,L-LA indeed has a reducing effect on PCL crystallinity,

especially in the case of non-isothermal crystallization of P(30/70). Additionally, the value of

PCL crystallinity calculated for the first heating of P(100/0) is somewhat higher, 57%, which

indicates that significant isothermal crystallization of P(100/0) films seems to take place

during storage at room temperature.

2.4.2 Hydrolytic degradation

Polymer samples were exposed to hydrolytic conditions for up to 26 weeks. Samples

were recovered at set time intervals and water absorption (Fig. 2.2a) and mass loss (Fig. 2.2b)

were determined along with changes in molecular weight and monomer composition (Fig.

2.3). Several factors come into play during degradation, including the presence of PEG

blocks that facilitate water uptake into the bulk, as does the increasingly amorphous nature

caused by higher D,L-LA content[2]. The degradation process involves two stages. It

proceeds, first, through random scission at the ester bonds along the polymer chain of the

amorphous regions. Once the amorphous regions are almost entirely degraded, the second

stage begins, involving stepwise scission of chain segments in the crystalline zones[26]. This

reaction produces carboxyl acids that lower the pH locally and are known to catalyze the

degradation process[27].

Page 71: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

54

Degradation time (weeks)

0 5 10 15 20 25

Abs

orpt

ion

(wt%

)

0

10

20

30

40

50P(100/0)P(70/30)P(30/70)

0 5 10 15 20 25

Mas

s lo

ss (

%)

0

5

10

15

20

P(100/0)P(70/30)P(30/70)

a) b)Degradation time (weeks)

Figure 2.2: a) Water absorption of P(100/0), P(70/30), and P(30/70) during degradation in PBS (37°C, pH 7,4) for up to 26 weeks. b) Mass loss of P(100/0), P(70/30), and P(30/70) samples as degradation in PBS (37°C, pH 7,4) proceeds for up to 26 weeks. (P(30/70) were not possible to handle correctly after 3 weeks of degradation.)

Page 72: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

55

10 12 14 16 18 20

a)0 w

10 w

26 w PI=1.9

PI=1.8

PI=1.7

10 12 14 16 18 20

b)

Elution time (min)

10 12 14 16 18 20

c)

26 w

10 w

0 w

0 w

26 w

10 w

PI=2.2

PI=2.1

PI=1.7

PI=1.4*

PI=2.1*

PI=1.7

Figure 2.3: Gel permeation chromatograms (signal intensity vs. elution time) showing the effect of degradation on the polydispersity and molecular weights of a) P(100/0), b) P(70/30), and c) P(30/70) samples after 0 (0w), 10 (10w) and 26 (26w) weeks of degradation. *Note that baseline separation was not obtained for P(30/70) samples after 10 and 26 weeks of degradation.

Page 73: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

56

The rate of degradation was strongly influenced by D,L-LA content. P(100/0) remained

seemingly unaffected by hydrolysis within the time frame of this study, retaining its initial

shape and appearance. In sharp contrast, the P(30/70) samples, which initially were

transparent, turned white and sticky, and barely retained their shape past the 3rd week of

degradation. Beyond this time point, these samples were too difficult to handle in order to

correctly determine mass loss and water absorption. In addition to the initial molecular

weight of this copolymer being lower, 1H-NMR and GPC measurements as well as early data

(Fig. 2.2) clearly point out P(30/70) as the fastest degrading material, owing to the larger

amount of D,L-LA. Furthermore, we must take into account the physical state of the LA-

containing copolymers, which may be more prone to degradation at 37°C. In fact, as the Tm

onset is observed at 35°C (data not shown), the material is slightly more amorphous at 37°C

and should therefore be more susceptible to hydrolysis.

The water absorption of the comparatively crystalline P(100/0) remained at a constant

level of approximately 5% of the sample dry weight, and this material seemed practically

unchanged throughout the investigated time period, with the exception of a 31% total drop in

mean molecular weight (Fig. 2.4a). The water content in P(70/30) samples, immediately

exceeded 10% and began to increase steadily after 6 weeks of immersion until a level of 50%

of the dry weight was reached at 26 weeks. From Figure 2.2b, it is also clear that the polymer

had not yet reached saturation at this time point. From our results, higher LA-content was

consistent with higher water absorption (Fig. 2.2a), and increasing mass loss from the

material (Fig. 2.2b). Cho et al.[18], although studying copolymers of considerably lower

molecular weights, reported similar trends of mass loss, while they observed lower rates of

water absorption, possibly due to the shorter PEG segment. Furthermore, in a study of the

hydrolytic degradation of random P(CL-co-LA) of comparable molecular weights, Malin et

al.[11] reported drastically lower water absorption and mass loss. This further emphasizes the

significant role played by the central PEG segment.

Page 74: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

57

Degradation time (weeks)

0 5 10 15 20 25

Mn

0

10000

20000

30000

40000

50000

60000

70000

P(100/0)P(70/30)P(30/70)

Degradation time (weeks)

0 5 10 15 20 25

(CL

+L

A):

EG

rat

io

0

2

4

6

8

P(100/0)P(70/30)P(30/70)

Degradation time (weeks)

0 5 10 15 20 25

% C

L in

end

seg

men

ts

20

30

40

50

60

70

80

90

100

P(70/30)P(30/70)

a)

b)

c)

Figure 2.4: a) The effect of degradation on the average molecular weight, Mn, as determined by GPC (n=3). Past week 10, baseline separation was no longer achieved for P(30/70) samples. The effect of degradation on monomer composition as calculated from 1H-NMR spectra: b) the (CL+LA):EG ratio; and c) the CL content in the end segments of P(70/30) and P(30/70) as compared to initial percentage. Insufficient amounts of P(30/70) remained to perform GPC at the last time point (26 weeks).

Page 75: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

58

The changes in polydispersity and molecular weight were followed with the help of gel

permeation chromatography (Fig. 2.3) and 1H-NMR (Fig. 2.4a). The results indicate several

differences in the way degradation occurs in the three polymers. Firstly, while P(100/0) is

subject to random hydrolytic scission in the PCL chain, the polydispersity index (PI) did not

change a great deal, and the total decrease in mean molecular weight (Mn) was

approximately 30% (see Fig. 2.4a). Meanwhile, the hydrolysis of P(70/30) generates CL-LA

blocks of varying sizes depending on the distribution of D,L-LA along the macromolecular

chain. A clear attenuation in Mn was noticed (86% over 26 weeks) along with an increase in

PI as degradation proceeds. In P(30/70), Mn diminishes drastically (96% overall). However,

the changes in polydispersity were difficult to quantify as baseline separation for P(30/70)

was lost after 10 weeks. It seemed, however, that all materials showed a more or less

pronounced trend of increasing polydispersity until 14 weeks (10 weeks for P(30/70)) (data

not shown), after which it gradually decreased. Similar behaviour has been previously

reported for PDLLA-PCL-PDLLA[10] and PLLA-PEG-PLLA[12] block copolymers.

Contrary to those two reports, our polymers displayed an initially bimodal molar mass

distribution, which for the LA-containing materials eventually disappeared with increased

degradation.

After 26 weeks of degradation of P(70/30), the ratio of (CL+LA)-to-EG increased

markedly (Fig. 2.4b). This abrupt change suggests the loss of PEG blocks from the

copolymer, as reported previously by Li et al.[14]. This may be explained by the fact that the

PEG unit, although in itself unsusceptible to hydrolysis, is water soluble, and so, may be

leached from the sample once it has been separated from surrounding P(CL-co-LA)

components as the result of extensive chain cleavage. No such event was observed in

P(30/70). This fact is most likely due to extensive mass loss from the sample, including PEG

blocks as well as D,L-LA and CL; hardly anything remained of P(30/70) at the end of the

degradation period.

Additional analysis of 1H-NMR spectra revealed that the fraction of CL rose from 30 to

53% in the CL-LA portion of P(30/70) between weeks 10 and 18 of degradation, and from 74

to 87% in P(70/30) samples between weeks 14 and 26 (Fig. 2.4c). However, clear evidence

of mass loss appeared much earlier during the study (weeks 3 and 10, respectively). As the

degrading films were of sufficient thickness (0.75mm), it is suggested that this may be a

Page 76: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

59

consequence of heterogeneous degradation mechanisms[28]. The results imply that

solubilised mono- and oligomers on the surface of the films immediately leach from the

sample, resulting in equal rates of loss of CL and LA (and PEG) due to random scission.

Meanwhile, degradation in the bulk can proceed under the additional action of autocatalysis

before the cleaved species may leave the sample[26]. The onset of leaching from the bulk

depends on the copolymer composition, and our results suggest that this appears earlier with

larger D,L-LA fraction (Fig. 2.4c). The D,L-LA units in the bulk also seem to be cleaved to a

greater extent than the CL fraction in the bulk, and are thus lost at a higher rate from the

samples at later stages of the degradation process (Fig. 2.4c).

2.4.3 Tensile strength

Tensile testing of P(100/0) and P(70/30) was performed on non-degraded samples as well

as partially degraded samples after 2-10 weeks of degradation. However, for P(30/70), given

the virtual absence of mechanical resistance, and the rapid loss of structural integrity during

degradation, only non-degraded specimens were tested. We noted that the tensile properties

of P(100/0) (Mn=95kDa) matched the tensile properties of pure PCL (Mw=50kDa). Thus, to

offer a point of reference, this data is included in Table 2.3.

Table 2.3: Changes in mechanical properties with time of degradation.

Degradation

Time (weeks)

Stress at Yield

(MPa)

Stress at Break

(MPa)

Elongation at Break

(%)

Young’s Modulus

(MPa)

PCL 0 11.1±0.6 14.6±1.7 750±56 184±7

0 13.9±1.2 18.3±2.1 747±44 254±21

2 13.0±0.8 14.7±0.8 665±68 211±11

4 14.1±1.0 15.2±1.0 601±80 223±4 P(100/0)

10 14.4±0.8 16.9±2.6 731±64 220±11

0 4.1±0.3 6.6±0.9 938±155 74±4

2 3.5±0.4 4.7±0.2 378±106 51±4

4 3.5±0.1 3.7±0.3 90±61 51±2 P(70/30)

10 - 1.8±0.2 5±1 38±3

P(30/70) 0 0.7±0.1 - > 1000 14±1

Partially degraded samples (2, 4, and 10 weeks) were tested while wet. No values for the maximum extension

were obtained for P(30/70), which did not break within the limits of the testing apparatus.

Page 77: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

60

When comparing the non-degraded co-polymers (Table 2.3), the mechanical strength

dropped quickly with increasing LA-content. Higher LA-content also coincides with lower

molecular weight and lower degree of crystallinity, which are both factors that lower the

mechanical resistance in a polymeric material. For example, the Young’s modulus (or initial

modulus) for P(100/0) was found to be 3.5 and 18 times higher than those of P(70/30) and

P(30/70), respectively. Conversely, elongation at break increased significantly with

increasing LA-content, whereas for P(100/0) and P(70/30), the corresponding values were

747% and 938%. Clearly, out of the synthesized co-polymers, P(100/0) was the material

possessing the highest mechanical strength, both in the dry state and after 10 weeks of

incubation in aqueous solution at 37°C.

In full agreement with the above reported observations of degradation, P(100/0) did not

appear to be significantly affected by 10 weeks of hydrolytic degradation, except a slightly

decreased Young’s Modulus, which falls from an average of 254±21 MPa for dry samples to

a rather centred medium value of 218±10 MPa for all partially degraded P(100/0) samples

(weeks 2, 6, and 10 inclusively). This may be the result of a certain degree of water

absorption in the material bulk (samples were wet when tested), and/or a small drop in mean

molecular weight. In comparison, the same period of hydrolysis resulted in a steady decrease

in mechanical strength of P(70/30). At the end of the study, a decrease in ultimate stress from

6.6±0.9MPa to 1.8±0.2MPa was observed, as well as a 51% decrease in Young’s modulus.

Notably, at 10 weeks, at which time the mean molecular weight of P(70/30) had dropped

below 20kDa (Fig. 2.4a), it became brittle and no yield could be observed. Drastically

decreasing values of the elongation at break were observed, beginning with a 60% relative

drop already after 2 weeks of degradation and finally yielding a poor value of 5% for

maximum elongation at 10 weeks. Thus, within the 10-week timeframe of this study, the loss

of tensile strength due to hydrolytic degradation may be principally attributed to the presence

of LA.

2.5 Conclusions

In this study we characterize three PEG-containing copolymers with varying ratios of D,L-

lactide (D,L-LA) and ε-caprolactone (ε-CL). As crystallinity in all samples was due mainly to

the CL fraction, the pure block copolymer P(100/0) was, not surprisingly, a highly crystalline

material, and hydrolytic degradation of this material was very slow. As a consequence, the

Page 78: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

61

tensile strength of this material was marginally affected by degradation. The incorporation of

increasing amounts of D,L-LA into the chain, rendered the copolymer increasingly

amorphous and thus, more apt to water uptake and subsequent degradation. Furthermore,

higher D,L-LA content did not only reduce the toughness of non-degraded copolymers, but

also resulted in a faster loss of mechanical resistance and integrity with degradation time.

We conclude that these copolymers are materials whose mechanical properties and

stability in vitro may be effectively altered by changing the content of D,L-LA. Based on the

findings of this study, we propose copolymers such as P(100/0) and P(70/30) as suitable

materials for large, three-dimensional scaffolds for tissue engineering.

2.6 Acknowledgments

This work was supported by the NSERC through a Discovery grant (PV) and a Strategic

grant (PV) and a post-doctorate fellowship (PS).

Page 79: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

62

2.7 References

(1) Helminen AO, Korhonen H, Seppälä JV. Cross-linked poly(ε-caprolactone/D,L-lactide) copolymers with elastic properties. Macromol. Chem. Phys. 2002; 203:2630-2639.

(2) Huang MH, Suming LM, Coudane J, Vert M. Synthesis and characterization of block copolymers of ε-caprolactone and DL-lactide initiated by ethylene glycol or poly(ethylene glycol). Macromol. Chem. Phys. 2003; 204:1994-2001.

(3) Hiljanen-Vainio MP, Orava PA, Seppälä V. Properties of ε-caprolactone/D,L-LA (ε-CL/D,L-LA) copolymers with a minor ε-CL content. J. Biomed. Mater. Res. 1997; 34:39-46.

(4) Pachence J, Kohn J. Biodegradable polymers for tissue engineering. In: Lanza R, Langer R, Chick W, editors. Principles of tissue engineering. Austin, Texas, USA: R.G. Landes Company, 1997: p.273-293.

(5) Bergsma JE, Rozema FR, Bos RRM, Boering G, Joziasse CAP, Pennings AJ. In vitro predegradation at elevated temperatures of poly(lactide). J. Mat Sci.: Mat. Med. 1995; 6:642-646.

(6) Pitt CG. Poly-ε-caprolactone and its copolymers. In: Chasin M, Langer R, editors. Biodegradable Polymers as Drug Delivery Systems. New York: Marcel Dekker, 1990: p.71.

(7) Baimark Y, Molloy R, Molloy N, Siripitayananon J, Punyodom W, Sriyai M. Synthesis, characterization and melt spinning of a block copolymer of L-lactide and ε-caprolactone for potential use as an absorbable monofilament surgical suture. J. Mat Sci.: Mat. Med. 2005; 16:699-707.

(8) Ge H, Hu H, Yang S, Jiang X, Yang C. Preparation, characterization, and drug release behaviors of drug-loaded ε-caprolactone/L-lactide copolymer nanoparticles. J. Appl. Polym. Sci. 1999; 75:874-882.

(9) Meek M, Jansen K, Steendam R, van Oeveren W, van Wachem P, Luyn M. In vitro degradation and biocompatibiliaty of poly(D,L-lactide-ε-caprolactone) nerve guides. J. Biomed. Mater. Res. 2004; 68:43-51.

(10) Huang MH, Li SM, Vert M. Synthesis and degradation of PLA-PCL-PLA triblock copolymer prepared by successive polymerization of ε-caprolactone and (DL)-lactide. Polymer 2004; 45:8675-8681.

(11) Malin M, Hiljanen-Vainio M, Karjalainen T, Seppala J. Biodegradable Lactone Copolymers. II. Hydrolytic Study of ε-Caprolactone and Lactide Copolymers. J. Appl. Pol. Sci. 1996; 59:1289-1298.

Page 80: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

63

(12) Li S, Anjard S, Rashkov I, Vert M. Hydrolytic degradation of PLA/PEO/PLA triblock copolymers prepared in the presence of Zn metal or CaH2. Polymer 1997; 39:5421-5430.

(13) von Burkersroda F, Gref R, Gopferich A. Erosion of biodegradable block copolymers made of poly(D,L-lactic acid) and poly(ethylene glycol). Biomaterials 1997; 18:1599-1607.

(14) Li SM, Garreau H, Vert M, Petrova T, Manolova N, Rashkov I. Hydrolytic degradation of poly(oxyethylene)-poly-(ε-caprolactone) multiblock copolymers. J. Appl. Polym. Sci. 1998; 68:989-998.

(15) Gorlatchouk A, Freier T, Shoichet M. Synthesis of degradable poly(L-lactide-co-ethylene glycol) porous tubes by liquid-liquid centrifugal casting for use as nerve guidance channels. Biomaterials 2005; 26:7555-7563.

(16) Saito N, Okada T, Horiuchi H, Murakami N, Takahashi J, Nawata M et al. A biodegradable polymer as a cytokine delivery system for inducing bone formation. Nat. Biotechnol. 2001; 19:332-335.

(17) Hu Y, Jiang X, Ding Y, Zhang L, Yang C, Zhang J et al. Preparation and drug release behaviors of nimodipine-loaded poly(caprolactone)-poly(ethylene oxide)-polylactide amphiphilic copolymer nanoparticles. Biomaterials 2003; 24:2395-2404.

(18) Cho H, An J. The effect of ε-caproyl/D,L-lactyl unit composition on the hydrolytic degradation of poly(D,L-lactide-ran-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-ran-ε-caprolactone). Biomaterials 2006; 27:544-552.

(19) Cho H, Chung D, Jeongho A. Poly(D,L-lactide-ran-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-ran- ε-caprolactone) as parenteral drug delivery systems. Biomaterials 2004; 25:3733-3742.

(20) Bogdanov B, Vidts A, Van Den Bulcke A, Verbeeck R, Schacht E. Synthesis and thermal properties of poly(ethylene glycol)-poly(ε-caprolactone) copolymers. Polymer 1998; 39:1631-1636.

(21) Rich J, Karjalainen T, Ahjopalo L, Seppala J. Model compound release from DL-lactide/ε-caprolactone copolymers and evaluation of specific interactions by molecular modeling. J. Appl. Polym. Sci. 2002; 86:1-9.

(22) Piao L, Dai Z, Deng M, Chen X, Jing X. Synthesis and characterization of PCL/PEG/PCL triblock copolymers by using calcium catalyst. Polymer 2003; 44:2025-2031.

(23) Gan ZH, Zhang J, Jiang BZ. Poly(ε-capralactone)/poly(ethylene oxide) diblock copolymer. 2. Nonisothermal crystallization and melting behavior. J. Appl. Polym. Sci. 1997; 63:1793-1804.

Page 81: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

64

(24) Sosnik A, Cohn D. Poly(ethylene glycol)-poly(ε-caprolactone) block oligomers as injectable materials. Polymer 2003; 44:7033-7042.

(25) Cresceni V, Manzini G, Calzolari G, Borri C. Thermodynamics of fusion of poly-β-propiolactone and poly-ε-caprolactone. Comparative analysis of the melting of aliphatic polylactone and polyester chains. Eur. Polym. J. 1987; 8:449-463.

(26) Li S. Hydrolytic degradation characteristics of aliphatic polyesters derived from lactic and glycolic acids. J. Biomed. Mater. Res., Part B 1999; 48:342-353.

(27) Pitt CG, Gratzel M, Kimmel G, Surles J, Schindler A. Aliphatic polyesters. 2. The degradation of poly(DL-lactide), poly(ε-caprolactone) and their copolymers in vivo. Biomaterials 1981; 2:215-220.

(28) Grizzi I, Garreau H, Li S. Hydrolytic degradation of devices based on poly(DL-lactic acid) size-dependence. Biomaterials 1995; 16:305-311.

Page 82: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

65

PREFACE TO CHAPTER 3. Following the reults of Chapter 2, it was decided –out

of pure curiosity – to look into wether blending of the co-polymers could be an

easy and efficient way to obtain intermediate rates of degradation. It was also

hypothesized that blending of slow- and fast-degrading co-polymers may result

in porous materials during the course of degradation.

Page 83: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

66

CHAPTER 3:

BLENDS AS A STRATEGY TOWARDS TAILORED HYDROLYTIC DEGRADATION OF

P(CL-CO-D,L-LA)-PEG-P(CL-CO-D,L-LA) COPOLYMERS

3.1 Abstract

Binary blends were prepared from poly(ε-caprolactone) (PCL), and P(CL-co-D,L-lactic

acid)-P(ethylene glycol)-P(CL-co-D,L-lactic acid) copolymers, where the D,L-LA content in

the side segments varied from 0 to 70 mol%. Blend discs were fabricated by melt-molding,

and the effect of blend composition on hydrolytic degradation was studied. Variations in

medium pH were monitored, and morphological changes were observed using scanning

electron microscopy. Blending of these copolymers was found to constitute a simple means by

which intermediate rates of water absorption and mass loss were obtained, compared to those

observed in pure copolymer preparations. In one of the blends, prepared from the two

components containing 70 or 0 mol% D,L-LA in the side segments, and thereby exhibiting

large differences in degradation rate, hydrolysis resulted in the formation of a porous material

over time. Furthermore, all blend samples maintained their initial shape throughout the study.

Such materials may be interesting for further investigations for applications in cellular therapy

and controlled release.

3.2 Introduction

Polymer blends offer a simple means of combining the properties of two materials to

achieve a third. Its simplicity makes it preferable, in some cases, to copolymerisation, which

is laborious and time consuming. In the field of tissue engineering, blending two polymers

with different degradation profiles may serve to tune physical and mechanical properties[1],

or to achieve a desirable rate of degradation[2]. Parameters affecting the controlled

degradation of blends, apart from the properties of the components, include the composition,

preparation method, compatibility and miscibility of the components[3,4]. Several blend pairs

of biodegradable homo- and co-polymers have been studied, including materials composed of

poly(ε-caprolactone) (PCL), polyanhydrides, poly(ethylene glycol) (PEG), and poly(D,L-lactic

Page 84: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

67

acid) (PDLLA)[1,2,4]. Morphological and thermodynamic analyses of polymer blends have

shown that PCL are immiscible but compatible with PEG[2,5] and PDLLA[6], whereas

PDLLA and PEG have been reported to be miscible[7]. However, in two co-

polymer/homopolymer blends, random P(CL-co-D,L-LA) co-polymers were found to be

partially compatible with PDLLA as well as with the amorphous part of the semicrystalline

PCL[4]. Furthermore, blends of PCL and PEG-PCL-PLA co-polymers exhibited properties

indicative of immiscibility of the PEG blocks with PCL[2].

The mechanism of hydrolytic degradation in large devices fabricated from aliphatic

polyesters has been described by Pitt et al.[8] and Grizzi et al.[9]. According to Grizzi et

al.[9], in the case of materials in the poly(glycolic acid)/polylactide family, initial water

uptake is very rapid and essentially homogeneous. However, after the onset of hydrolysis,

degradation proceeds at a faster rate in the sample interior due to the autocatalytic effect of

generated carboxylic acids and the difficulty of soluble degradation products to diffuse out of

the sample. As a result, a thick sample will eventually be completely degraded internally,

while a “skin” layer of less degraded polymer forms around the device, a phenomenon that

has been observed both in vitro and in vivo[10,11]. For semi-crystalline polymers, such as

PCL, the way a sample may be expected to degrade over time is thus thickness dependent.

In a recent publication, P(CL-co-D,L-LA)-PEG-P(CL-co-D,L-LA) co-polymers, containing

either 100, 70 or 30 mol% of CL in the side segments, were characterized and their

degradation profiles studied[12]. The resulting materials are referred to here as P100, P70, and

P30, respectively. Using melt-molded films, it was found that by varying the LA-content in

the co-polymers, mechanical properties and the rate of hydrolytic degradation could be

controlled. However, the degradation profiles of the three tested materials covered a very

large range, and in the case of P30, rapid degradation resulted in early loss of sample

integrity[12]. In view of previous results, this study was designed to answer two main

questions: 1) Can intermediate degradation profiles be obtained by way of co-polymer

blending? And 2) does the blending of a slow-degrading material with a fast-degrading one

yield a porous matrix during the course of degradation? In order to answer these questions

three blend pairs were prepared; PCL/P70, P100/P70, and P100/P30. Based on previous

reports on the morphology and miscibility of similar co-polymers and blends, it is reasonable

to assume that phase separation would result in PCL/P70 blends, due to the reported

Page 85: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

68

immiscibility of PCL with both PDLLA and PEG. The same would be true for P100/P30 and

P100/P70, due to the D,L-LA content of P30 and P70, although these two blends may be

expected to exhibit a higher degree of compatibility.

3.3 Materials and Methods

3.3.1 Materials

PEG diols (MW 7 500, cat. # 06103) and D,L-LA dimers (cat. # 16640) were purchased

from Polysciences Inc (Warrington, PA, USA). SnOct (cat. # S3252) and ε-CL (cat. # 21510)

were obtained from Sigma-Aldrich (Oakville, ON, Canada). PCL pellets (CAPA® 6500, MW

50 000) were provided by Solvay Caprolactones (Houston, TX, USA).

3.3.2 Copolymers, blend preparation and sample fabrication

All copolymers were synthesized according to Bogdanov et al.[13] as described

earlier[12]. The resulting co-polymers were composed of a central PEG block with two side

segments in which the molar ratio of ε-CL:D,L-LA was either 100:0, 70:30 or 30:70. The

corresponding copolymers are referred to as P100, P70, and P30, respectively, according to

the ε-CL-content in the side segments. The obtained molecular weights for P100, P70, and

P30 were 94 800, 81 100, and 52 400 Da, respectively, as previously determined[12].

Copolymer blends were prepared from a 10% solution (wt/wt) of two (co)polymers

(weight ratio of 1:1) in dichloromethane in a round-bottom flask, where they were stirred for

one hour. Polymer blends were recovered by precipitation in cold ethanol, filtration and

drying to constant weight. The blends thus prepared were PCL/P70, P100/P70, and P100/P30.

Films were fabricated using a hot melt press at 120°C, using a stainless steel mold. From each

blend film, discs-shaped specimens of two thicknesses (1.25mm and 0.75mm) were prepared.

Samples were punched out using a 1-cm diameter biopsy punch and dried under vaccum prior

to use.

3.3.3 Hydrolytic degradation

MASS LOSS, BUFFER ABSORPTION, AND PH OF THE DEGRADATION MEDIUM

Polymer degradation by hydrolysis was studied by incubating the circular discs in a

phosphate buffer solution for up to 26 weeks. Samples were dried in vacuum overnight,

weighed, and sterilized by soaking in 70% ethanol for 30 minutes. (The average weight of

thin and thick samples was 97 and 126mg, respectively.) Each sample was then immersed in

Page 86: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

69

2mL PBS, pH 7.4, and incubated at 37°C. The buffer was supplemented with 100U/mL

penicillin and 100mg/mL streptomycin to prevent bacterial growth. The medium was initially

changed every 3 weeks for the first 6 weeks, every 2 weeks for the following three months,

and finally every week from week 18 on. At predetermined intervals (weeks 1, 3, 6, 10/12, 14,

18, and 26; n=3), samples were removed from the degradation medium, washed 3x5min in

deionized water and gently blotted on filter paper before being weighed to determine PBS

absorption. Samples were then dried under vacuum, weighed again and the mass loss and PBS

absorption were determined from Equation 3.1 and 3.2, respectively:

Mass loss (%) = 0

0100w

ww t−× (3.1)

Absorbtion (%) = t

twet

w

ww −×100 (3.2)

where w0 is the initial weight of the sample, wt is the mass of the degraded sample after

drying, and wwet is the wet weight of a retrieved sample. In addition, changes in the pH of the

degradation buffer were monitored throughout the 26-week study.

MORPHOLOGY

Scanning electron microscopy (SEM) was used to follow any morphological changes, at

the surface and in cross-sections of thick samples, that might take place during the course of

degradation. Cross-sections were obtained by breaking the discs following quenching at

-80°C. Discs were sputter-coated with Pd/Au, and imaged at 5kV using a JEOL JSM-840-A

scanning electron microscope.

3.34 Statistical Analysis

Experimental data were evaluated for statistical significance using the Student’s t-test. A

value of p < 0.05 was considered significant.

Page 87: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

70

3.4 Results and Discussion

3.4.1 Mass loss and buffer absorption

Figure 3.1A shows the water absorption in thin samples as a function of time and blend

composition. Absorption curves for pure co-polymer films of identical geometry were also

included for comparison [12]. At the initial absorption level (reached after one week),

absorption increased in the following order: PCL/P70 < P100/P70 < P100/P30. This order was

then conserved throughout the duration of the study. For P100/P30 blends it was observed that

the rate of absorption in thin samples increased until week 18, whereas a marked decrease was

observed in this rate between weeks 18 and 26. Similar profiles around the six-month mark

were previously reported by Huang et al.[2] for thinner (0.4mm) films fabricated from PLA-

PCL-PLA and PLA-PCL-PEG block co-polymers, and this was accredited to behaviour

“typical” of PEG and PLA blocks.

Regarding mass loss in the three blends, similar trends were observed, as may be seen in

Figure 3.1B. Twenty-six weeks of degradation resulted in 4.5-7% total mass loss in PCL/P70

and P100/P70. As expected, P100/P30 exhibited significantly faster mass loss compared to the

other blend compositions. As in the case of absorption, the rate of mass loss in thin P100/P30

samples decreased towards the end of the study. However, in this binary blend, composed of

slow-degrading P100 and fast-degrading P30, it would be expected to see the slower

degradation profile of P100 appearing towards the end of the degradation of the P30 phase.

Similar mass loss and absorption profiles were obtained for thick samples (data not shown).

Compared to pure P100, P70, and P30, intermediate rates of water absorption and mass

loss were indeed achieved for blends (see Fig. 3.1). The large differences found for pure P70

and P30 films, due to disparities in hydrophilicity, molecular weight and crystallinity[12],

induced considerable differences also in this study, as shown by elevated rates of mass loss in

P100/P30 blends compared to P100/P70. It is interesting to observe that pure P30 samples,

which were impossible to handle past three weeks of hydrolysis, contributed to faster rates of

sample degradation in this study, while the integrity of the discs was maintained by the P100

phase in the blend preparation. In fact, all blend samples studied preserved their initial shape

throughout the study. In addition, we may conclude that the PEG block in P100 facilitates

water uptake in P100/P70 blends compared to PCL/P70 (p<0.05 from weeks 1 and 6 in thin

Page 88: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

71

and thick samples, respectively). The effect of the PEG block on mass loss was less

pronounced, but significantly higher mass loss was observed after 26 weeks (p<0.05).

Fig. 3.1: Mass loss (% of initial weight) and water absorption (% of dry weight) presented as a function of degradation time over 26 weeks. The graphs show water absorption (A) and mass loss (B) in thin samples as a function of blend composition. The profiles of pure co-polymers discs of the same thickness have been incorporated for comparison [9]. The pure co-polymers are represented by filled symbols and the blends by open symbols.

3.4.2 Evolution of pH of degradation media

The changes in pH over time of degradation are shown in Figure 3.2. The blends

composed of relatively slow-degrading (co)polymers, i.e. PCL/P100 and P100/P70, did not

induce any large fluctuation in pH over 26 weeks. P100/P30 samples provoked significantly

lower pH between eight and 20 weeks of degradation, with thin samples reaching a minimum

of 3.8 at 12 weeks and thick P100/P30 samples reaching a minimum of 3.0 at 14 weeks. The

decrease in pH observed for P100/P30 blend discs corresponds to the degradation and loss of

amorphous regions[14]. Thus, at the end of the study, the more resistant crystalline regions

remain and pH is maintained stable. The pH for thick samples behaved similarly, but was

slightly shifted towards the end of the study. Moreover, while the pH suddenly increased at 18

weeks, mass loss in thick samples continued at a nearly constant rate from week 10 to 26.

This suggests that the acidic degradation products were retained longer within the bulk of the

thicker samples, thus continuing to catalyze the hydrolytic process, and resulting in sustained

mass loss, in agreement with previous reports on thinner films[15].

Time of degradation (weeks)

0 3 6 9 12 15 18 21 24 27

Ab

s (%

)

0

20

40

60

80

100

120

P100PCL/P70P70P100/P70P30P100/P30

Time of degradation (weeks)

0 3 6 9 12 15 18 21 24 27

Mass

lo

ss (

%)

0

5

10

15

20

25

30

35

40

A) B)

Page 89: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

72

We may also note that the pH for degrading thick samples exhibited a sharp increase after

week 18, at the same time as medium changes become more frequent. This is hardly a

coincidence, and may serve to illustrate the importance of frequent medium changes for fast-

degrading polyesters, in support of previous findings[16].

Time of degradation (weeks)

0 4 8 12 16 20 24

pH

2

3

4

5

6

7

8

Fig. 3.2: Effect of sample thickness and blend composition on the pH of the degradation media. Changes in pH of the degradation media are presented for (○) thick PCL/P70, (●) thin PCL/P70, ( ∇ ) thick P100/P70, (▼) thin P100/P70, (□) thick P100/P30, and (■) thin P100/P30. The vertical lines indicate the times at which the frequency of the change of media was changed from every three weeks to every two weeks (left line), and from every two weeks to every week (right line).

Page 90: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

73

3.4.3 Morphology

Surfaces of P100/P30 changed considerably with continuing degradation, as can be seen

in Figure 3.3, while no changes in PCL/P70 and P100/P70 blend discs were observed, apart

from a slight cracking of the dried sample surfaces. Dispersed, micrometer-sized pores could

be seen already in non-degraded samples, and pores of a slightly larger size were visible by 10

weeks of degradation. However, rather than the initial pores simply enlarging, there seems to

be a top layer containing tiny pores, which is gradually removed, revealing an underlying

layer exhibiting fewer pores.

Fig 3.3: Scanning electron micrographs of partially degraded P100/P30 copolymer blends. Surfaces of thick samples (1,25mm) are shown un-degraded (A), and after 10 weeks (B), and 26 weeks (C). The white scale bars indicate 10µm.

Cross-sections of thick discs showed gradual changes in all blends as a result of

progressing hydrolysis (Figure 3.4). For P100/P30, signs of degradation were clearly visible

already at 6 weeks (not shown), and samples were evidently porous at weeks 18 and 26, with

homogenously dispersed pores throughout the bulk in the size range of 1-10µm. These pores

are likely the result of the degradation and solubilisation of the P30-rich phase, and their

appearnce coincide with the lower pH levels observed in the middle of this study.

Page 91: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

74

Fig 3.4: Scanning electron micrographs of partially degraded PCL/P70, P100/P70, and P100/P30 co-polymer blends. Images of cross-sections of thick samples (1.25mm) were taken in the region close to the surface of the discs, and shown un-degraded, and after 18 weeks and 26 weeks. The white scale bars indicate 10µm.

3.5 Conclusions

The present study illustrates how the use of co-polymer blends may constitute a simple

and straightforward technique to fine tune rates of hydrolytic degradation in these materials.

The presence of a PEG block in the P100 component significantly increased water uptake in

P100/P70 blends as compared to PCL/P70. P100/P30 exhibited the highest water uptake and

mass loss, due to the LA-rich P30 component, and degradation resulted in a porous material

over time. The pH in the degradation media remained stable for PCL/P100 and P100/P70

blends, while rapid degradation in P100/P30 samples resulted in very low pH levels (pH 3-5)

2

PCL/P70

P100/P70

P100/P30

Undegraded 18 weeks 26 weeks

2

PCL/P70

P100/P70

P100/P30

Undegraded 18 weeks 26 weeks

Page 92: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

75

in the middle of the study, reflecting the release of acidic degradation products from the

samples.

It would be of interest to optimize blending conditions, possibly using other blending

techniques, in order to investigate the development of the morphology in these degrading

blends in more detail. The objective of such a study would be to fabricate porous materials

from initially non-porous ones, as such materials may find interesting use in cellular therapy

and controlled release applications.

3.6 Acknowledgments

This work was supported by the NSERC through a Discovery grant (PV) and a Strategic

grant (PV) and a post-doctorate fellowship (PS).

Page 93: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

76

3.7 References

(1) Tsuji H, Ikada Y. Blends of aliphatic polyesters .1. Physical properties and

morphologies of solution-cast blends from poly(DL-lactide) and poly(ε-

caprolactone). J. Appl. Polym. Sci. 1996; 60:2367-2375.

(2) Huang MH, Li SM, Hutmacher DW, Coudane J, Vert M. Degradation characteristics of poly(ε-caprolactone)-based copolymers and blends. J. Appl. Polym. Sci. 2006; 102:1681-1687.

(3) Domb AJ. Degradable polymer blends .1. Screening of miscible polymers. J. Polym. Sci., Part A: Polym. Chem. 1993; 31:1973-1981.

(4) Shen YQ, Sun WL, Zhu KJ, Shen ZQ. Regulation of biodegradability and drug release behavior of aliphatic polyesters by blending. J. Biomed. Mater. Res. 2000; 50:528-535.

(5) Tsuji H, Smith R, Bonfield W, Ikada Y. Porous biodegradable polyesters. I. Preparation of porous poly(L-lactide) films by extraction of poly(ethylene oxide) from their blends. J Appl Polym Sci 2000; 75:629-637.

(6) Aslan S, Calandrelli L, Laurienzo P, Malinconico M, Migliaresi C. Poly(D,L-lactic acid)/poly(e-caprolactone) blend membranes: preparation and morphological characterisation. J Mater Sci 2000; 35: 1615-1622.

(7) Na YH, He Y, Shuai X, Kikkawa Y, Doi Y, Inoue Y. Compatibilization effect of poly(e-caprolactone)-b-poly(ethylene glycol) block copolymers and phase morphology analysis in immiscible poly(lactide)/poly(e-caprolactone) blends.

(8) Pitt CG, Gratzel M, Kimmel G, Surles J, Schindler A. Aliphatic polyesters II. The degradation of poly(DL-lactide), poly(e-caprolactone), and their copolymers in vivo. Biomaterials 1981; 2(4): 215-220.

(9) Grizzi I, Garreau H, Li S, Vert M. Hydrolytic degradation of devices based on poly(DL-lactic acid) size-dependence. Biomaterials 1995; 16:305-311.

(10) Li SM, Garreau H, Vert M. Structure property relationships in the case of the degradation of massive poly(alpha-hydroxy acids) in aqueous-media .2. Degradation of lactide-glycolide copolymers - PLA37.5GA25 and PLA75GA25. J. Mater. Sci.: Mater. Med. 1990; 1:131-139.

(11) Therin M, Christel P, Li SM, Garreau H, Vert M. In vivo degradation of massive poly(Alpha-hydroxy acids) - validation of in vitro findings. Biomaterials 1992; 13:594-600.

(12) Bramfeldt H, Sarazin P, Vermette P. Characterization, degradation, and mechanical strength of poly(D,L-lactide-co-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-co-ε-caprolactone). J. Biomed. Mater. Res., Part A 2007; 83(2): 503-511.

Page 94: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

77

(13) Bogdanov B, Vidts A, Van Den Bulcke A, Verbeeck R, Schacht E. Synthesis and thermal properties of poly(ethylene glycol)-poly(ε-caprolactone) copolymers. Polymer 1998; 39:1631-1636.

(14) Hu D, Luh J. Role of crystallin emorphology in degradation kinetics of poly(D,L-lactide), poly(L-lactide, and poly(ε-caprolactone) for medical implants. Biomed Eng Applic Basis Commun 1994; 6:88-93.

(15) Lu L, Garcia CA, Mikos AG. In vitro degradation of thin poly(DL-lactic-co-glycolic acid) films. J. Biomed. Mater. Res. 1999; 46:236-244.

(16) Grayson ACR, Cima MJ, Langer R. Size and temperature effects on poly(lactic-co-glycolic acid) degradation and microreservoir device performance. Biomaterials 2005; 26:2137-2145.

Page 95: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

78

PREFACE TO CHAPTER 4. As stated in the conclusion of Chapter 2, two of the synthesized co-

polymers were considered suitable for use in large, porous scaffolds. However, initial tests

showed very poor cell adhesion on films fabricated from these co-polymers, which indicated a

need for surface modification.

Co-cultures of smooth muscle cells and endothelial cells are being used as a technique to

generate stable and functional vascular networks in biomaterials, and such strategies was also

discussed in Chapter 1. As these cells constitute a natural and essential part of blood vessels, it

was chosen to optimize cell-material interactions using smooth muscle cells. In Chapter 4, the

effect of a protein-based surface modification technique on smooth muscle cell adhesion and

proliferation is evaluated.

Page 96: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

79

CHAPTER 4:

ENHANCED SMOOTH MUSCLE CELL ADHESION AND PROLIFERATION ON

PROTEIN-MODIFIED POLYCAPROLACTONE-BASED COPOLYMERS

4.1 Abstract

Smooth muscle cells (SMC) were cultured for up to six days on copolymer films

fabricated from a PCL-PEG-PCL block co-polymer or P(ε-CL-co-D,L-LA)-PEG-P(ε-CL-co-

D,L-LA), named P(100/0) and P(70/30), respectively. The films were modified by aminolysis

using 1,6-hexanediamine, and fibronectin, fibrinogen, or fibrin layers were subsequently

immobilized by physisorption or by covalent coupling using imidoester chemistry.

Immobilization of all the tested proteins resulted in significantly enhanced cell adhesion on

these polymers. Moreover, we found that covalently immobilized proteins supported

significantly greater cell proliferation than physisorbed proteins over six days. SMC cultured

on P(100/0) films modified by covalently attached fibronectin or fibrin layers proliferated at a

rate comparable to that observed on control tissue culture polystyrene. The proposed surface

modification schemes were much less efficient in improving cell attachment and proliferation

on P(70/30) films. However, pre-wetting P(70/30) with a phosphate buffer prior to aminolysis

significantly improved cell numbers following immobilization of fibronectin. Immuno-

staining of smooth muscle-specific α-actin of SMC grown on protein-modified P(100/0) 8h

and 48h after cell seeding, confirmed preserved SMC phenotype on all modified surfaces.

4.2 Introduction

The performance of scaffolds for tissue engineering applications is governed not only by

material properties, biocompatibility and degradability, but also by the way they facilitate and

guide cellular interactions with their surfaces. Scaffold surfaces need to be conceived in a

manner that permits adequate cell adhesion, growth, morphology and viability over time in

order to meet the demands of a given application. Synthetic scaffold materials in particular,

although presenting superior mechanical properties and offering a greater degree of control

over surface chemistry and material properties, lack the biochemical motifs provided by

biologically derived materials, such as collagen and fibrin gels. Chemical and/or biochemical

Page 97: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

80

modification of synthetic materials therefore becomes necessary in order to enhance cell-

material interactions. Numerous publications can be found describing the detailed effects of

modified synthetic-surface properties on cellular function e.g., in endothelial cells[1-3],

fibroblasts[4], smooth muscle cells[3,5], osteoblasts[6], and adipose stem cells[7].

Smooth muscle cells (SMC) are present in gastrointestinal, urological, and cardiovascular

tissues (e.g., in the lining of larger blood vessels), and represent a cell type of high interest for

tissue engineering applications. The above mentioned tissues are all examples of mechanically

dynamic environments, which explain the dependence of SMC phenotype on mechanical

stimuli, as previously demonstrated by several research groups[8-11]. Moreover, studies

comparing various protein coatings have shown that attachment strength and migration in

SMC may be controlled both by the type and concentration of plated proteins[5,12]. On the

same note, covalent modification of biomaterials with adhesion peptides and growth factors

has been shown to enhance SMC adhesion[13] as well as extracellular matrix production[14].

However, some researchers argue that isolated peptides sequences, although efficient in

promoting a particular cell response (such as the RGD motif supporting cell adhesion), are

rather poor at reproducing the diverse functions of key proteins, such as fibronectin[15].

Consequently, it becomes clear that optimizing the strength and nature of SMC adhesion to a

biomaterial would be crucial to controlling phenotype while achieving effective culture

conditions under flow and/or post-implantation.

Polyesters, including polycaprolactone (PCL), polyglycolide, polylactides (PLA) and their

co-polymers, can be activated through the covalent attachment of amino groups to PCL films

by reacting the ester groups with a diamine, such as 1,6-hexanediamine. The aminolyzed PCL

films, may then be further modified by the immobilization of proteins or adhesion peptides

using, e.g., glutaraldehyde[16] or carbodiimide chemistry[7]. These approaches were both

shown to effectively improve cell attachment and proliferation compared to unmodified and

aminolyzed PCL[7,16]. Dimethyl pimelimidate (DMP) is a homobifunctional, cross-linking

agent that may be used to couple proteins to aminated surfaces via reaction with primary

amines on proteins, while introducing a 7-carbon spacer. This water soluble molecule is

readily available and holds potential for biomaterial applications, as shown by previous

studies in which it was used for the effective immobilization of enzymes, while preserving

enzyme activity and stability[17,18].

Page 98: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

81

The present study reports the adhesion and proliferation of smooth muscle cells grown on

surface-modified co-polymers consisting of PEG, D,L-LA, and/or ε-CL (for detailed

characterization, see Bramfeldt et al.[19]). Following aminolysis, DMP chemistry was used to

investigate the effects of immobilized proteins and thin fibrin layers on SMC adhesion and

growth. It is demonstrated that this immobilization method represents a highly efficient means

of creating bioactive materials from polyether-esters. Cell adhesion and proliferation were

both significantly improved compared to unmodified polymer films, while preserving smooth

muscle cell phenotype, as shown by smooth muscle α-actin labeling.

4.3 Materials and Methods

4.3.1 Materials

PEG diols (MW 7500, cat. no. 06103) and D,L-LA monomers (cat. no. 16640) were

purchased from Polysciences Inc. Stannous octoate (SnOct, cat. no. 3252) and ε-caprolactone

(ε-CL, cat. no. 21510) were obtained from Sigma-Aldrich. For surface modification, bovine

plasma thrombin (cat. no. T4265) and fibrinogen (cat. no. F4753), as well as fibronectin (cat.

no. F4759), 1,6-hexanediamine (cat. no.H11696), and dimethyl pimelimidate (DMP, cat. no.

D8388) were purchased from Sigma-Aldrich.

4.3.2 Co-polymer synthesis

Two co-polymers were synthesized as described earlier[19]. Briefly, macroinitiator (PEG

diols) and monomers (ε-CL or a mix of 70mol% ε-CL and 30mol% D,L-LA) were introduced

into a round-bottom flask, where they were melted and degassed under constant stirring. The

monomer:EG ratio was kept at 4.2:1 The ring-opening polymerization reaction took place at

120˚C for 24h, using SnOct as a catalyst (monomer:catalyst = 1:1000). The co-polymers were

recovered by dissolution in dichloromethane and precipitation in cold ethanol. The resulting

co-polymers were a pure block co-polymer, PCL-PEG-PCL, and a poly(ε-caprolactone-co-

D,L-lactide)-poly(ethylene glycol)-poly(ε-caprolactone-co- D,L-lactide), referred to here as

P(100/0) and P(70/30), respectively. The molecular weights of P(100/0) and P(70/30) were

95kDa and 81kDa , respectively, with polydispersity indices of 1.7. The glass transition

temperatures, Tg, of P(100/0) and P(70/30) were -59˚C and -45˚C, respectivley as previously

determined19.

Page 99: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

82

4.3.3 Film preparation

Films were fabricated between Teflon sheets in a Carver Laboratory hot melt press at

120°C, using a stainless steel mold. The entire mold was subsequently cooled under cold,

running water. Following drying of the films, samples were punched out using a 1-cm

diameter biopsy punch. For this study, films of 0.25-mm thickness were prepared.

4.3.4 Surface modification

AMINOLYSIS

Samples (prepared as described above) were immersed in a 10wt% solution of 1,6-

hexanediamine in isopropanol, and the incubation time varied from 0-2h. The samples were

subsequently washed three times in Milli-Q water to remove excess of 1,6-hexanediamine.

P(100/0) samples were further treated in 5mM HCl to protonize the amine groups. Three

additional rinsing steps were carried out and the aminolyzed films were then dried under

vacuum and stored in a desiccator until use. P(70/30) samples were always used freshly

aminolyzed (i.e., same day).

PROTEIN IMMOBILIZATION

Aminolyzed samples were incubated in 10mM DMP in PBS, pH 8, for 1h at room

temperature. Following a rinsing step (PBS, pH 8), fibronectin and fibrinogen solutions of

predetermined concentration (10µg/mL) were added and allowed to react for 4h at 4˚C

followed by 1h at room temperature. Fibrin-mimicking surfaces were prepared by adsorbing

fibrinogen from a 100µg/ml fibrinogen (Fgn) solution onto aminated polymer films for 1h at

room temperature without previous treatment with DMP. The excess Fgn was aspirated and a

fresh mixture of 25 NIH units/ml thrombin and 10mM DMP (pH 8) was immediately added.

An immobilized fibrin-like gel layer was allowed to form for 3h at 37˚C. Control surfaces

carrying physisorbed proteins were prepared in a similar fashion, but without DMP. Following

the completed immobilization or adsorption of proteins, all modified surfaces were rinsed in

PBS, and sterilized overnight at 4˚C in PBS containing 500U/mL penicillin and 500mg/mL

streptomycin. Sterilization using antibiotics was previously shown to be an efficient and mild

alternative for aliphatic polyesters20. Finally, the modified polymer films were thoroughly

rinsed in sterile PBS containing no antibiotics and transferred to new, sterile 48-well plates,

after which they were immediately used for cell culture experiments.

Page 100: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

83

4.3.5 Characterization of amine content

Polymer films were prepared by melt press moulding and aminolyzed, as described

above. The presence of amine groups on the films was confirmed and quantified by a

ninhydrin assay. Ninhydrin reacts with available amino groups on the surface, forming a

purple compound, which can be detected by a spectrophotometer. A 1.0M ninhydrin solution

was prepared in ethanol and the aminolyzed co-polymer discs were immersed for one minute.

The discs were then separately transferred to clean glass vials and heated in a water bath at

70˚C for 15min to accelerate the reaction. The reacted samples were dissolved in 3ml THF

and 2ml isopropanol was added to the solution. Finally, the absorbance at 570nm was

measured in a spectrophotometer. Each condition was analyzed in triplicate and identical, but

untreated polymer discs were used as controls. Pure PCL was included in the experiment as a

reference material. A standard curve was prepared using known concentrations of 1,6-

hexanediamine.

4.3.6 Cell culture

Human Umbilical Artery Smooth Muscle Cells (SMC) (Promocell, Heidelberg, Germany)

were cultured in basic growth medium recommended by the manufacturer, supplemented with

basic fibroblast growth factor (2ng/mL), epithelial growth factor (0.5ng/mL), insulin

(5µg/mL), 5% fetal bovine serum, and antibiotics (100U/mL penicillin and 100mg/mL

streptomycin). Medium was changed every two to three days, and cells were used for

experiments between passages 4-6.

4.3.7 Cytotoxicity

The cytotoxicity of the degradation media collected from partially degraded co-polymer

films was tested using a mitochondrial activity MTT assay. P(100/0) and P(70/30) film

samples (0.75 mm thick, 10 mm Ø, 100mg medium sample weight) were allowed to degrade

individually in 2mL phosphate buffer, pH 7.4, at 37˚C, and medium from the degrading

samples was collected at 2 weeks, 6 weeks, and 10 weeks and stored at 4˚C until use.

SMC were seeded into 24-well plates and allowed to proliferate. When the cells reached

near-confluence, growth medium containing 10% (v./v.) filter-sterilized degradation medium

was added to each well. Following a 24h incubation period, the medium was aspirated and the

wells rinsed in PBS. Medium containing MTT-formazan substrate (Thiazolyl Blue

Tetrazolium Bromide, cat. no. M5565, Sigma-Aldrich) at a final concentration of 5mg/ml was

Page 101: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

84

added and allowed to react for 6h at 37˚C, whereupon the medium was removed. The formed

formazan crystals were subsequently solubilized in DMSO and a glycine buffer was added to

stabilize the product. Each condition was tested in 4 separate wells and absorption was

measured in a plate reader at λ=560nm. In addition, in order to quantify any Sn content in the

degradation media, atomic absorption spectroscopy was carried out using a Varian SpectrAA

55 apparatus at an absorption wavelength λ=235.5nm.

4.3.8 Cell adhesion and proliferation

Human umbilical artery smooth muscle cells (SMC) were detached using trypsin/EDTA

and seeded at 10,000 cells/well (7,800 cells/cm2) onto modified and unmodified polymer

samples, prepared as described in Section 4.3.4.

Cellular proliferation was followed by stopping the culture at given time intervals (24h, 3

days, and 6 days). The cells were trypsinized, stained in trypan blue, and the live cells were

subsequently counted in a haemacytometer. Prior to trypsinization, the films were carefully

transferred to new multiwell plates and rinsed in PBS to avoid counting loosely adhering cells

and cells adhering to underlying tissue culture polystyrene (TCPS). Three discs of each

surface type were counted together (considered as n=1), and the cell number/cm2 was

calculated. Cell numbers were then normalized to the number of adhering cells on TCPS after

24h. For each experiment, n=3-4 (i.e., 9-12 discs). Aminolyzed and unmodified co-polymers

were used as controls. In addition, SMC cultured on Fn-immobilized films as well as on TCPS

were trypsinized at the end of the experiment, reseeded, and labeled for α-actin to confirm

differentiated phenotype.

The morphology of cells cultured on modified co-polymers was studied using scanning

electron microscopy (SEM) and fluorescent staining of α-actin. As P(70/30) did not support

the fixation procedure, only P(100/0) films could be analyzed. For SEM imaging, cells were

fixed in glutaraldehyde (1,4% for 30min at room temperature and 2,5% overnight at 4˚C ),

post-fixed in OsO4 (1h at room temperature), followed by dehydration in serial dilutions of

ethanol and drying under vacuum. (Neither co-polymer resists critical point drying due to low

Tm[19]). Samples were sputter-coated in Au/Pd and imaging was carried out in a JSM-840-A

JEOL microscope (V=5kV, I=6×10-8A). P(70/30) samples did not resist the chemical

treatment used to fix the cells for SEM imaging and therefore only P(100/0) films could be

analyzed.

Page 102: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

85

Further evaluation of the cell shape during the adhesion and the subsequent spreading of

the cells on various surfaces, was carried out by fluorescent labeling of smooth muscle cell-

specific α-actin using a mouse monoclonal antibody. A secondary FITC-conjugated goat-anti-

mouse antibody was used to visualize the labeling. Cells were observed using a Nikon TE200-

S inverted microscope equipped with a high power mercury lamp.

4.3.9 Statistical analysis

Experimental data from MTT and ninhydrin assays were evaluated for statistical

significance using the Student’s t-test, and significance was considered for p<0.05. Cell

adhesion and proliferation data were analyzed using ANOVA. When overall significance

(P<0.05) was indicated, data pairs within a group were compared using calculated values for

critical difference (CD). Data pairs between groups were compared using Student’s t-test.

Page 103: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

86

4.4 Results and Discussion

4.4.1 Cytotoxicity

The cytocompatibility of the two co-polymers P(100/0) and P(70/30) was evaluated using

the MTT assay. (For the results of the degradation study, the reader is referred to Chapter 2.)

The cell viability after 24h incubation in culture medium containing 10% degradation medium

is shown in Figure 4.1. No significant difference in viability compared to TCPS cultures was

observed for degrading P(100/0) or for P(70/30) at two and six weeks of degradation.

However, at 10 weeks of degradation, the P(70/30) degradation medium caused a significant

decrease in cell viability (64±10%) compared to the TCPS control (p<0.05).

Fig 4.1: MTT mitochondrial activity assay of the cytocompatibility of degradation medium. The degradation medium was collected after 2, 6, and 10 weeks of hydrolytic degradation.

In an attempt to identify a possible cause of the observed cytotoxic effect of the 10-week

P(70/30) sample, the pH of the degradation media was monitored. The pH of the P(70/30)

degradation medium was observed to decrease from an initial value of 7.4 to 6.3 following 10

weeks of degradation (data not shown). However, at the concentration tested (10% v./v.), the

pH did not deviate more than 0.1 units from 7.4 when measured immediately following

mixing, and had in all cases returned to physiological pH after 30 minutes in room

temperature. It was also thought that residual Sn from the co-polymer synthesis may affect

cell viability, and a verification of the Sn content in the degradation media was made using

Control P(100/0) P(70/30)

Cel

l v

iab

ilit

y (

%)

0

20

40

60

80

100

120

Control2 weeks6 weeks10 weeks

Page 104: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

87

atomic absorption spectroscopy. All samples were tested and they were all found to contain 2-

3ppm Sn. This is approximately 10 times smaller than the IC50 of Sn previously reported for

fibroblasts[22]. Therefore, in the final concentration (10%) used for the MTT assay, it is

unlikely that Sn would have caused the observed decrease in cell viability.

4.4.2 Surface modification

Co-polymer films, prepared as described, were aminolyzed for 0, 0.5, 1, or 2h and

analyzed for free amino group content using a colorimetric ninhydrin assay. The results were

compared to a reference pure PCL film, and can be seen in Figure 4.2. No significant increase

in amine groups between 0.5 and 2h was observed on PCL or P(70/30). The comparable

amine content on PCL and P(100/0) films following 0.5h and 1h of aminolysis, indicates that

the presence of unreactive PEG-blocks does not interfere with the reaction. Seeing as the PEG

content of P(100/0) decreases the degree of crystallization compared to PCL[19], the

amorphous regions would possibly allow for the diamine molecules to more effectively

penetrate the film surface with longer incubation times, which may be a possible explanation

for the rapid and significant increase in amine content between 1h and 2h (p<0.05). Moreover,

it has been reported that amorphous regions more aptly react with diamines as they increase

surface wettability[22]. This was supported by the observations made of amine group

incorporation onto P(70/30). This material is even more amorphous in nature[19], and rapidly

reached amine densities higher than those of both P(100/0) and PCL (Fig. 4.2). However, only

small increases in amine group content were achieved with prolonged times of incubation.

The incubation times of the two co-polymers were selected to achieve a similar degree of

surface activation, while minimizing the negative impacts of prolonged incubation times on

mechanical strength[16,22]. Therefore, for protein immobilization schemes, P(100/0) films

were incubated for 2h, while P(70/30) aminolysis was limited to 30 minutes.

In view of the poor results obtained in previous studies[16] as well as in our laboratory,

using surface characterization techniques such as X-ray photoelectron spectroscopy (XPS) and

Fourier-transform infrared (FT-IR) spectroscopy, no additional characterization is reported

here. Nonetheless, cell behavior was considered to constitute a reliable, although non-

quantitative, indication of successful surface modification.

Page 105: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

88

Fig 4.2: (A) Absorbance at 570nm of aminolyzed copolymer films, showing the dependence of incubation time on the amount of 1,6-hexanediamine incorporated in the copolymer film surfaces. A pure PCL reference is included for comparison.

The significance towards a final engineered tissue composite of the three protein

compositions chosen for surface modification here may be illustrated by the integrin

interactions they trigger. These integrins, in turn, activate distinct intracellular signaling

pathways that play an important role in regulating cell phenotype[23-26]. SMC adhesion to Fn

stimulates expression of several integrins, including the main Fn receptor α5β1, as well as

integrins α8β1, and αvβ3[26]. The α8β1 integrin is also a known differentiation marker in

vascular smooth muscle cells, and it was recently shown that expression of the α8 subunit is

required to maintain a differentiated phenotype[24]. Fibronectin is, in fact, the only adhesion

protein recognized by the α5β1 integrin, which plays a key role in the formation of fibrillar

adhesions[15]. Moreover, both α5β1 and αvβ3 receptors are implicated in the regulation of

angiogenesis[15] and Fn has been suggested as a suitable matrix component in scaffolds

designed to regenerate blood vessels[27,28]. The αvβ3 integrin is likewise expressed in cells

adhering to fibrin and fibrinogen[26], and is believed to enhance SMC migration and

proliferation in vivo[23]. And although these are not always desirable events in vivo, they may

be profitable when promoting cellular scaffold invasion in vitro.

A) Time of aminolysis (h)

0,0 0,5 1,0 1,5 2,0

A.U

. ( λλ λλ

=5

75

nm

)

0

30

60

90

120

150

180

NH

2 g

rou

ps

( µµ µµm

ol/

cm2)

0

2

4

6

8

10

PCLP(100/0)P(70/30)

*

A) Time of aminolysis (h)

0,0 0,5 1,0 1,5 2,0

A.U

. ( λλ λλ

=5

75

nm

)

0

30

60

90

120

150

180

NH

2 g

rou

ps

( µµ µµm

ol/

cm2)

0

2

4

6

8

10

PCLP(100/0)P(70/30)

*

Page 106: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

89

4.4.3 Smooth muscle cell culture

SMC ADHESION AND PROLIFERATION ON P(100/0)

Cell adhesion was evaluated 24h after seeding, and cell proliferation was assessed

following three and six days in culture. Cell adhesion was improved on all protein-modified

surfaces, compared to unmodified and aminolyzed P(100/0) (see Fig. 4.3). No significant

difference in cell adhesion on P(100/0) was found between the two latter groups. Cell

proliferation showed a stronger dependence on protein type than did cell adhesion. Notably,

surfaces bearing covalently immobilized Fn and fibrin showed cell numbers comparable to

those on TCPS surfaces after six days. Fibrinogen alone stimulated strong proliferative

activity in SMC, although cell numbers were clearly lower on Fgn (229 ± 27%) than on fibrin

(304 ± 16%) on day six.

In addition, covalently immobilized proteins evoked different cell responses than

physisorbed proteins, indicating that the degree of substrate anchorage was a determining

factor. Adsorbed Fn promoted more efficient SMC adhesion (78±15%) than immobilized Fn

(59±10%) (p<0.01). Moreover, a large difference was found in proliferation on physisorbed

versus covalently coupled Fn, where cell numbers on physisorbed Fn were significantly lower

after six days in culture (p<0.01). A similar but more pronounced trend was observed for

fibrin-mimicking surfaces. Despite good initial cell adhesion, the physisorbed fibrin did not

support any significant proliferation over six days. The dependence of early cell adhesion

mechanisms on Fn surface anchorage was previously shown[2], and such mechanisms may

play a role in subsequent proliferation rates. However, as for the differences observed between

the adsorbed fibrin and Fn, the displacement of proteins from the surface may play an

important role. It can be hypothesized that the lower cell numbers observed on physisorbed

fibrin compared to that on Fn, may be explained by a more rapid and complete detachment of

the gel-like fibrin layer compared to individually adsorbed Fn molecules. Further

investigations will be needed to address this hypothesis.

Finally, in order to verify SMC phenotype, cells that were grown on Fn-immobilized

P(100/0) and TCPS were trypsinized after 6 days of culture, plated separately on TCPS and

stained for α-actin. Both cell groups exhibited strong α-actin labeling (see Appendix, Fig A3),

Page 107: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

90

and it was concluded that differentiated phenotype was conserved in these samples throughout

the duration of this experiment.

Fig 4.3 : Cell adhesion and proliferation on aminolyzed and protein-immobilized P(100/0) films. The P(100/0) surfaces were aminolyzed (NH2), and modified with covalently immobilized (“imm”) fibronectin (Fn), fibrin (Fib), and fibrinogen (Fgn) or with adsorbed (“ads”) Fn and Fib. Unmodified P(100/0) and TCPS were included for comparison. All cell numbers are normalized to TCPS at 24h. * indicates significantly higher cell adhesion numbers after 24h compared to the aminated surface (NH2) at 24h. ** indicates significant increase in cell number from 24h to 6 days within one group, p<0.05.

SMC ADHESION AND PROLIFERATION ON P(70/30)

Cell-surface interactions on P(70/30) clearly were not optimized using the surface

modification method described here. The observed cell numbers (Fig. 4.4) indicated

unsuccessful protein immobilization, in turn signifying different surface chemistry as

compared to P(100/0). This was somewhat unexpected seeing as both co-polymers

investigated were poly(ether-esters) with much the same monomer content. However,

although partial crystallization takes place in both materials, P(70/30) is a much more

amorphous material compared to the highly crystalline P(100/0)[19]. Additionally, the onset

of water absorption and decreasing molecular weights in P(70/30) was observed to occur

much earlier during hydrolytic degradation. It was thus hypothesized that unsuccessful protein

TCPS P(100/0) NH Fn ads Fn imm Fib ads Fib imm Fgn imm

Cel

l n

um

ber

(%

TC

PS

at

24

h)

0

100

200

300

400

24h3 days6 days

*

* **

*

**

****

**

2TCPS P(100/0) NH Fn ads Fn imm Fib ads Fib imm Fgn imm

Cel

l n

um

ber

(%

TC

PS

at

24

h)

0

100

200

300

400

24h3 days6 days

*

* **

*

**

****

**

TCPS P(100/0) NH Fn ads Fn imm Fib ads Fib imm Fgn imm

Cel

l n

um

ber

(%

TC

PS

at

24

h)

0

100

200

300

400

24h3 days6 days

*

* **

*

**

****

**

2

Page 108: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

91

immobilization may have been the result of polymer chain reorganization. In both cases, the

“hiding” of incorporated amine groups from the surface layer, would drastically diminish the

capacity for protein immobilization using DMP. In order to investigate the plausibility of this

explanation, dry samples as well as P(70/30) samples that were pre-wet for 24h in PBS (pH

7.4) at room temperature were used for aminolysis and subsequent protein modifications, and

the cell response was compared. The results can be seen in Figure 4.4.

From Figure 4.4, it is clear that pre-wet samples were superior to dry-treated films in terms

of cell numbers. Aminated surfaces without proteins did not improve cell adhesion, regardless

of whether samples were pre-wet or not. However, at six days, the pre-wet aminolyzed

P(70/30) supported significantly higher cell numbers (p<0.05) than their dry-treated

counterparts. Among the surfaces tested, only Fn immobilized on pre-wet P(70/30) lead to a

significant increase in adhering cell numbers compared to unmodified P(70/30) (p<0.01).

Moreover, the surfaces tested, only pre-wet P(70/30) bearing covalently coupled Fn supported

weak, but non-significant cell proliferation between days 1 and 6 (27±6% and 44±13%,

respectively). This stands in sharp contrast to the poor cell adhesion observed on its dry-

treated counterpart. No significant changes in cell numbers between days 1 and 3 on any

surface were observed. Adsorbed Fn on pre-wet surfaces also lead to markedly improved cell

adhesion compared to unmodified and aminated co-polymers, although in contrast to P(100/0)

results, this surface treatment did not support cell proliferation. The more hydrophilic nature

of P(70/30) compared to P(100/0), while as yet unmeasured, is clearly visible to the eye.

Displacement of physisorbed Fn was previously shown to take place to a greater extent from

hydrophilic polymer surfaces, whereas hydrophobic polymer substrates are better able to

retain the proteins[2]. Differences in Fn displacement from the two surfaces may thus offer a

partial explanation as to why cell numbers continue to multiply over six days of culture on

physisorbed Fn on P(100/0), but not on the corresponding P(70/30) surface.

Page 109: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

92

Fig 4.4 : SMC growth was compared on Fn-modified P(70/30), previously aminated from either pre-wet (marked with ♦ in group labels) or dry films. The effect of amination alone, covalently immobilised Fn (“imm”) on dry and pre-wet films as well as absorbed Fn (“abs”) were investigated. Unmodified P(70/30) was included for comparison. The 100% level corresponds to the cell number on TCPS after 24h of culture. ** in the graph indicates significantly higher cell adhesion numbers after 24h compared to the NH2* surface at 24h, p<0.05.

SMC MORPHOLOGY

SMC in vivo are characterized by phenotypic plasticity. Highly responsive to

environmental cues, divergent populations exhibiting different phenotypes may be present

within one tissue at the same time[29]. Via reversible mechanisms, SMC may switch from an

immature, proliferative, synthetic phenotype to a quiescent, contractile state and back

again[29,30]. These phenotypic conditions are characterized by specific cell shapes. Synthetic

SMC have a more hypertrophic and rectangular morphology while the contractile phenotype

appears bipolar and spindle-shaped[31].

Cell morphology and spreading were observed using scanning electron microscopy (SEM)

and fluorescent α-actin labeling. Unfortunately, the poor resistance of P(70/30) samples to the

fixation procedures limited SEM imaging to include P(100/0) surfaces only (Fig. 4.5). Firstly,

it can be seen that SMC attached on aminolyzed P(100/0) (Fig. 4.5D) are much smaller (60

µm) than cells on protein-modified surfaces (100 µm) (Fig. 4.5A-C). All cells showed distinct

filopodia 8h after seeding. SMC grown on protein-modified surfaces became more elongated

with time, with some extremely elongated features (100 µm) appearing at 48h. Additionally,

parallel grooves, resulting from irregularities in the Teflon sheets used during melt press film

fabrication, are clearly visible in the SEM images. However, there was no preferential cellular

alignment along these surface features.

P(70/30) NH NH Fn ads Fn imm Fn imm*

Cel

l n

um

ber

(%

of

TC

PS

at

24h

)

0

10

20

30

40

50

60

24h3 days6 days

2 2

**

**

♦ ♦ ♦P(70/30) NH NH Fn ads Fn imm Fn imm*

Cel

l n

um

ber

(%

of

TC

PS

at

24h

)

0

10

20

30

40

50

60

24h3 days6 days

2 2

**

**

♦ ♦ ♦♦ ♦ ♦

Page 110: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

93

8h 48h

A)

B)

C)

D)

8h 48h

A)

B)

C)

D)

Fig 4.5: α-actin staining of SMC on modified P(100/0), 8h and 48h after seeding. Surfaces were treated by aminolysis (A), followed by physisorption of Fn (B), or by covalent immobilization of Fn (C) and Fgn (D). The white bar indicates 100µm in all images but for (A) 8h, where the white bar indicates 10µm.

Page 111: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

94

8h 48h

A)

B)

C)

D)

E)

8h 48h

A)

B)

C)

D)

E)

Fig 4.6 : Alpha-actin staining of SMC on modified P(100/0), 8h and 48h after seeding. Surfaces were treated by aminolysis (A), followed by physisorption of Fn (B), or by covalent immobilization of Fn (C), Fgn (D), and fibrin (E). Original magnification: 40x.

Page 112: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

95

Fig 4.7 : Alpha-actin staining of SMC on modified P(70/30), 8h and 48h after seeding. Surfaces were treated by aminolysis on pre-wet films (A), followed by adsorption of Fn on pre-wet films (B), and covalent immobilization of Fn on dry (C) or pre-wet (D) films. Original magnification: 10x. Where higher magnification shows interesting features, inserts at 60x magnification have been included.

8h 48h

A)

B)

C)

D)

8h 48h

A)

B)

C)

D)

8h 48h

A)

B)

C)

D)

Page 113: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

96

Staining of the SMC-specific cytoskeleton component α-actin further supported the

morphological observations made in the SEM. In addition, positive α-actin immuno-staining

confirmed SMC phenotype. Cells grown on aminated surfaces were irregular and retracted.

On modified P(100/0) surfaces (Fig. 4.6), a well-defined, rather homogeneously distributed

cytoskeleton was observed in all cells following 8h of adhesion. The adhered cells exhibited

rounded, dendritic cell shapes, characteristic of the synthetic phenotype. Cells were clearly

elongated on all protein-modified surfaces after 48h, although cells growing on physisorbed

fibronectin (Fig. 4.6B) seemed to rearrange in a more pronounced aligned fashion than cells

on covalently immobilized proteins.

SMC cytoskeleton staining on P(70/30) films modified with fibronectin again revealed

large differences in morphology as compared to that of cells on P(100/0). Partial and extensive

spreading was observed already after 8h on pre-wet aminated or Fn-modified films (Fig. 4.7).

With the exception of dry-treated P(70/30), cells were more bipolar and slightly more

elongated following 48h incubation, although the populations were heterogeneous in this

regard. Strong α-actin staining appeared around the nucleus after 8h on prewet, aminolyzed

P(70/30) with or without subsequent Fn physisorption, but this feature disappeared following

48h in culture (7A, B). Films that had been aminolyzed from the dry state and subsequently

Fn-modified (Fig. 7C) did not support cell spreading, which is most probably a result of

unsuccessful fibronectin immobilization as discussed above, but this requires further

investigation.

It should be noted that all cell cultures were carried out in medium containing fetal bovine

serum (FBS). It was deemed necessary to perform cultures with FBS to promote proliferation

and avoid cell death within the six days of culture. Although this implies that all surfaces may

have been modified by adsorbing proteins from the medium, which may have aided cell

adhesion, significant differences in cell numbers were still observed between unmodified,

aminolyzed, and protein-modified polymers.

4.5 Conclusions

In conclusion, we observed that immobilization of fibronectin, fibrinogen and fibrin using

imidoester chemistry greatly enhanced the ability of P(100/0) and P(70/30) to support SMC

adhesion and proliferation. Cell adhesion was equally improved on physisorbed fibronectin

Page 114: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

97

and fibrin, although proliferation rates on these surfaces were hampered, demonstrating a clear

dependence on the degree of protein anchorage. Improved cell adhesion and proliferation were

obtained on all protein-modified surfaces. On modified P(100/0) covalently immobilized

fibronectin and fibrin supported SMC proliferation comparable to that on TCPS control

surfaces following six days in culture. Protein immobilization on P(70/30) was not as efficient

in improving cell-material interactions. Pre-wetting this co-polymer in PBS prior to

aminolysis and subsequent modification steps greatly improved cell numbers, in all

probability as a result of more efficient protein immobilization. However, cell numbers on

protein-modified surfaces were constantly and considerably lower on P(70/30) than on

P(100/0), and no significant cell proliferation was observed on any of the P(70/30)-based

surfaces. Further investigation will be required to establish why.

Morphological observations of cells grown on aminated copolymer films showed irregular

and retracted cell shapes. SMC grown on protein-modified P(100/0) revealed rectangular,

dendritic cell morphology 8h after seeding, while cells became bipolar and elongated after

48h. A less evident elongation with time of culture was observed on modified P(70/30),

although cells were comparatively more spread already at 8h.

4.6 Acknowledgements

This work was supported by the NSERC through a Discovery grant (PV) and a Strategic

grant (PV).

Page 115: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

98

4.7 References

(1) Baguneid M, Murray D, Salacinski HJ, Fuller B, Hamilton G, Walker M, et al. Shear-

stress preconditioning and tissue-engineering-based paradigms for generating arterial substitutes. Biotechnol. Appl. Biochem. 2004;39:151-157.

(2) Pompe T, Kobe F, Salchert K, Jorgensen B, Oswald J, Werner C. Fibronectin anchorage to polymer substrates controls the initial phase of endothelial cell adhesion. J. Biomed. Mat. Res., Part A 2003;67A:647-657.

(3) Serrano MC, Pagani R, Manzano M, Comas JV, Portoles MT. Mitochondrial membrane potential and reactive oxygen species content of endothelial and smooth muscle cells cultured on poly(ε-caprolactone) films. Biomaterials 2006;27:4706-4714.

(4) Faucheux N, Schweiss R, Lutzow K, Werner C, Groth T. Self-assembled monolayers with different terminating groups as model substrates for cell adhesion studies. Biomaterials 2004;25:2721-2730.

(5) Dimilla PA, Stone JA, Quinn JA, Albelda SM, Lauffenburger DA. Maximal migration of human smooth-muscle cells on fibronectin and type-IV collagen occurs at an intermediate attachment strength. J. Cell Biol. 1993;122:729-737.

(6) Lieb E, Hacker M, Tessmar J, Kunz-Schughart LA, Fiedler J, Dahmen C, et al. Mediating specific cell adhesion to low-adhesive diblock copolymers by instant modification with cyclic RGD peptides. Biomaterials 2005;26:2333-2341.

(7) Santiago LY, Nowak RW, Rubin JP, Marra KG. Peptide-surface modification of poly(caprolactone) with laminin-derived sequences for adipose-derived stem cell applications. Biomaterials 2006;27:2962-2969.

(8) Costa KA, Sumpio BE, Cerreta JM. Increased elastin synthesis by cultured bovine aortic smooth-muscle cells subjected to repetitive mechanical stretching. FASEB J. 1991;5:A1609.

(9) Kulik TJ, Alvarado SP. Effect of stretch on growth and collagen-synthesis in cultured rat and lamb pulmonary arterial smooth-muscle cells. J.Cell. Physiol. 1993;157:615-624.

(10) Kim BS, Nikolovski J, Bonadio J, Mooney DJ. Cyclic mechanical strain regulates the development of engineered smooth muscle tissue. Nat. Biotechnol. 1999; 17:979-983.

(11) Riha GM, Wang XW, Wang H, Chai H, Mu H, Lin PH, et al. Cyclic strain induces vascular smooth muscle cell differentiation from murine embryonic mesenchymal progenitor cells. Surgery 2007;141:394-402.

Page 116: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

99

(12) Goodman LV, Majack RA. Vascular smooth-muscle cells express distinct transforming growth factor-beta receptor phenotypes as a function of cell-density in culture. J. Biol. Chem. 1989;264:5241-5244.

(13) Mann B, Tsai A, Scott-Burden T, West J. Modification of surfaces with cell adhesion peptides alters extracellular matrix deposition. Biomaterials 1999;20:2281-2286.

(14) Mann BK, Schmedlen RH, West JL. Tethered-TGF-beta increases extracellular matrix production of vascular smooth muscle cells. Biomaterials 2001;22:439-444.

(15) Vogel V, Baneyx G. The tissue engineering puzzle: A molecular perspective. Annu. Rev. Biomed. Eng. 2003;5:441-463.

(16) Zhu YB, Gao CY, Liu XY, Shen JC. Surface modification of polycaprolactone membrane via aminolysis and biomacromolecule immobilization for promoting cytocompatibility of human endothelial cells. Biomacromolecules 2002;3:1312-1319.

(17) Chernukhin IV, Klenova EM. A method of immobilization on the solid support of complex and simple enzymes retaining their activity. Anal. Biochem. 2000;280:178-181.

(18) Moeschel K, Nouaimi M, Steinbrenner C, Bisswanger H. Immobilization of thermolysin to polyamide nonwoven materials. Biotechnol. Bioeng. 2003;82:190-199.

(19) Bramfeldt H, Sarazin P, Vermette P. Characterization, degradation, and mechanical strength of poly(D,L-lactide-co-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-co-ε-caprolactone). J. Biomed. Mater. Res., Part A 2007;83:502-511

(20) Shearer H, Ellis MJ, Perera SP, Chaudhuri JB. Effects of common sterilization methods on the structure and properties of poly(D,L-lactic-co-glycolic acid) scaffolds. Tissue Eng. 2006;12:2717-2727.

(21) Tanzi MC, Verderio P, Lampugnani MG, Resnati M, Dejana E, Sturani E. Cytotoxicity of some catalysts commonly used in the synthesis of copolymers for biomedical use. J. Mat. Sci.: Mat. Med. 1994;5:393-396.

(22) Fukatsu K. Mechanical-Properties of Poly(Ethylene-Terephthalate) Fibers Imparted Hydrophilicity with Aminolysis. J. Appl. Pol. Sci. 1992;45:2037-2042.

(23) Clemmons DR, Horvitz G, Engleman W, Nichols T, Moralez A, Nickols GA. Synthetic alpha V beta 3 antagonists inhibit insulin-like growth factor-I-stimulated smooth muscle cell migration and replication. Endocrinology 1999;140:4616-4621.

(24) Zargham R, Thibault G. alpha 8 Integrin expression is required for maintenance of the smooth muscle cell differentiated phenotype. Cardiovasc. Res. 2006;71:170-178.

Page 117: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

100

(25) Zargham R, Thibault G. alpha 8 integrin overexpression in the less differentiated vascular smooth muscle cells restores the differentiation properties and attenuates migratory activity. Arterioscler., Thromb., Vasc. Biol. 2006;26:E70.

(26) Moiseeva EP. Adhesion receptors of vascular smooth muscle cells and their functions. Cardiovasc. Res. 2001;52:372-386.

(27) Underwood S, Afoke A, Brown RA, MacLeod AJ, Dunnill P. The physical properties of a fibrillar fibronectin-fibrinogen material with potential use in tissue engineering. Bioprocess Eng. 1999;20:239-248.

(28) Harding SI, Afoke A, Brown RA, MacLeod A, Shamlou PA, Dunnill P. Engineering and cell attachment properties of human fibronectin-fibrinogen scaffolds for use in tissue engineered blood vessels. Bioprocess Biosyst. Eng. 2002;25:53-59.

(29) Halayko AJ, Solway J. Molecular mechanisms of phenotypic plasticity in smooth muscle cells. J. Appl. Physiol. 2001;90:358-368.

(30) Halayko AJ, Camoretti-Mercado B, Forsythe SM, Vieira JE, Mitchell RW, Wylam ME, et al. Divergent differentiation paths in airway smooth muscle culture: induction of functionally contractile myocytes. Am. J. Physiol.: Lung Cell. Mol. Physiol. 1999;276:L197-L206.

(31) Kocher O, Gabbiani F, Gabbiani G, Reidy MA, Cokay MS, Peters H, et al. Phenotypic features of smooth-muscle cells during the evolution of experimental carotid-artery intimal thickening - biochemical and morphological studies. Lab. Invest. 1991;65:459-470.

Page 118: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

101

PREFACE TO CHAPTER 5. In this final chapter, the co-polymers that were

characterized and surface-modified in previous chapters are used to fabricate

three-dimensional, porous scaffolds with controllable pore size. The surface

modification technique employed in Chapter 4, is translated to the three-

dimensional structures and an initial smooth muscle cell adhesion test is carried

out.

Page 119: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

102

CHAPTER 5 :

SMOOTH MUSCLE CELL ADHESION IN SURFACE-MODIFIED THREE-DIMENSIONAL

CO-POLYMER SCAFFOLDS PREPARED FROM CO-CONTINUOUS BLENDS

5.1 Abstract

In this article, we report on the applicability of the co-continuous blend technique for

the fabrication of tissue-engineering scaffolds from co-polymers, using polystyrene as the

porogen phase. The co-polymers consisted of P(ε-CL-co-D,L-LA)-PEG-P(ε-CL-co-D,L-LA),

referred to as P(100/0) and P(70/30), respectively, reflecting the %monomer content (CL/LA)

in the side chains. By means of static annealing of the blends for 30-120min followed by

extraction of the porogen phase, pore sizes roughly in the range of 15-350µm could be

obtained. The efficiency of in vitro seeding of smooth muscle cells was investigated in three-

dimensional, fibronectin-modified P(100/0) scaffolds of two different pore sizes. Considerably

higher numbers of adhering cells were found in the scaffolds with larger pore sizes, indicating

the importance of this parameter to promote effective cell intrusion into bulk materials.

Furthermore, the influence of scaffold pore size on compressive strength, permeability was

investigated. Compressive moduli ranged from 2.3±0.3 to 67±15 MPa and decreased with

increasing pore size. The reverse trend was found for scaffold permeability, which generated

values of κ from 8.5 × 10-16 to 6.7 × 10-11 m2. This level of permeability was comparable to

values reported for other scaffolds with higher pore sizes and void volumes, but with less open

pore morphologies. Taken together, the results obtained in this study strongly motivate further

investigation for possible future application in tissue engineering.

5.2 Introduction

The fabrication of porous polymer scaffolds has attracted much attention in the past 15

years. For tissue engineering, the importance of matching the properties of the scaffold to that

of the tissue being replaced has been repeatedly underlined by many research teams. These

properties include mechanical characteristics, permeability, pore size and porosity, and

resemblance of the cell-scaffold interface to the extracellular matrix surrounding cells in

natural tissues[1-3]. It has also been proposed that the rate of degradation in vitro and/or in

vivo should match an appropriate rate of tissue growth as it proceeds within the scaffolds[2].

Page 120: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

103

Numerous techniques have been described in the literature for the fabrication of porous

polymer structures for tissue engineering, including the solvent casting/particulate leaching

method[4-6], gas foaming[7], emulsion freeze drying[8], phase separation[9], and solid free-

form fabrication[10]. For further details on these and other techniques, the reader is referred to

the numerous reviews on the subject[11-14]. The majority of scaffolds used for the

engineering of smooth muscle-based tissues are either fabricated from non-woven

mesh[15,16] or particulate leaching[5,17] methods using members of the poly(lactic acid)

(PLA)/ poly(glycolic acid) (PGA) family. Methods combining particulate leaching with free-

form fabrication[18] and gas foaming[19] have also been reported. The drawbacks of particle

leaching methods are the use of organic solvents, and the commonly occurring heterogeneous

pore morphology (principal pores connected by smaller fenestrations), which may complicate

cell invasion. However, results indicate that synthetic biomaterials are suitable for the

engineering of smooth muscle-based tissues, as smooth muscle cells (SMC) generally show

higher proliferative potential and elastin production when seeded in synthetic scaffolds

compared to collagen-based constructs[15,19]. Furthermore, to enable effective SMC

invasion, pore sizes of at least 100µm have been recommended[18].

The use of immiscible binary polymer blends for potential applications in tissue

engineering has attracted increasing attention lately[20-23]. In immiscible blends, the minor

phase forms dispersed droplets in the main matrix. However, upon increasing the

concentration of the minor component, the dispersed phase will grow and coalesce until it

becomes the major phase. Close to this phase inversion point, both phases become fully

interconnected, and the blend morphology can be described as co-continuous. This technique

not only offers control over parameters such as porosity and pore size[20,23], but due to the

continuous nature of the blend phases, it also ensures an entirely interconnected porosity once

the porogen phase has been leached out. However, since the fully co-continuous region does

not allow for very large variations in blend composition, the porosity generated is generally

not higher than 60-70%[21,23]. Pore size has been effectively altered for several blend pairs

by static annealing of the blends at temperatures above the Tm of both components[24].

Among the blends that have been studied, and which may be relevant to tissue engineering,

are P(L-LA)/polystyrene (PS), poly(ε-caprolactone) (PCL) /poly(ethylene glycol), PLLA/PCL,

P(D,L-LA)/PCL, and P(L-LA)/PGA[20-23,25]. Pore size and porosity may also influence such

properties as permeability and mechanical strength, which may thus be tuned to meet the

Page 121: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

104

demands of a specific scaffold application. These qualities imply the suitability of these

scaffolds towards the study of cell interactions with three-dimensional porous structure. This

was recently illustrated by Washburn and co-workers[26], who showed the feasibility of using

such structures for regeneration of bone tissue.

It is generally accepted that surface modification methods are key elements in implant or

tissue engineering to improve and control the interaction of cells with synthetic polymer

materials[29-33]. However, to our knowledge, while many detailed studies of two-

dimensional models can be found in the literature, very few results of cell culture in truly

three-dimensional, covalently surface-modified, synthetic tissue-engineering scaffolds have

been reported. Schemes involving coating techniques based on physisorption of proteins from

aqueous solution may improve cell adhesion and proliferation within porous scaffolds, but it is

not clear whether such coatings would resist long-term culture or withstand mechanically

demanding environments, such as e.g. the shear stresses occurring within a perfusion

bioreactor.

In a recent publication we characterized three flexible (ε-CL-co-D,L-LA)-PEG-P(ε-CL-co-

D,L-LA) co-polymers with varying ε-CL: D,L-LA ratios in the side chains[27]. Based on the

obtained results, co-polymers containing 100 and 70 monomer% ε-CL in the side chains, were

judged to possess mechanical properties and degradation profiles suitable for scaffold

fabrication. These materials, referred to here as P(100/0) and P(70/30), respectively, were

subsequently evaluated for protein-modification and two-dimensional smooth muscle cell

culture[28]. In this work, we attempted to prepare porous scaffolds from these two co-

polymers using immiscible binary blends with polystyrene as the porogen phase. Following

selective extraction of the polystyrene phase and verification of the co-continuous nature of

the resulting porosity, the efficiency of annealing as a tool for increasing and controlling pore

size of the scaffolds was studied. In addition, the changes in compressive stiffness and

permeability with varying pore size were followed. Finally, three-dimensional scaffolds were

covalently modified with fibronectin and seeded with SMC using a dynamic cell seeding

method, to examine whether these structures may be suitable for the engineering of smooth-

muscle based tissues.

Page 122: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

105

5.3 Materials and Methods

5.3.1 Materials

PEG diols (MW 7500, cat. no. 06103) and D,L-LA monomers (cat. no. 16640) were

purchased from Polysciences Inc (Warrington, PA, USA). Stannous octoate (SnOct, cat. no.

3252) and ε-caprolactone (ε-CL, cat. no. 21510) were obtained from Sigma-Aldrich (Oakville,

ON, Canada), while polystyrene (PS, MW 192 000) was obtained from Dow (Midland, MI,

USA). For surface modification, fibronectin (cat. no. F4759), 1,6-hexanediamine (cat.

no.H11696), and dimethyl pimelimidate (DMP, cat. no. D8388) were all purchased from

Sigma-Aldrich.

5.3.2 Co-polymers

Two copolymers were synthesized as described earlier[27] using ring-opening

polymerization, stannous octoate as a catalyst, and a PEG diol as macro-initiator. By varying

the percent composition of the monomer mix, the side chains contained either 100% ε-CL, or

70% ε-CL and 30% D,L-LA in random distribution. The resulting two copolymers were thus a

pure block copolymer, PCL-PEG-PCL, MW 95 kDa, and a poly(D,L-lactide-co-ε-

caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-co-ε-caprolactone), MW 81 kDa,

referred to here as P(100/0) and P(70/30), respectively.

5.3.3 Preparation of blends

Binary melt blends were prepared by melt mixing in a Brabender internal mixer (Plasti-

corder Digi-System, C.W. Brabender Instruments Inc., NJ 07606). Vacuum-dried polystyrene

and co-polymer were introduced to the mixing chamber at a 1:1 weight ratio. The temperature

was set to 160°C for both blends. However, to enable blending of P(70/30)/PS, the

temperature was increased to reach an effective temperature of 165°C, which was maintained

for five minutes before returning to the initial level. The time of blending was 15 minutes for

all blends, and the mixing speed was kept constant at 50rpm. The torque remained nearly

stable, indicating that no significant degradation took place in either blend.

5.3.4 Annealing and scaffold fabrication

For this study, polystyrene (PS) was used as the porogen phase. P(100/0)/PS and

P(70/30)/PS blends were rapped in Teflon sheets and aluminium foil and placed between

metal slabs (without applied pressure) in an oven and annealed at 200˚C and 180˚C,

Page 123: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

106

respectively. Additionally, a set of P(100/0)/PS blends was annealed at 160˚C to verify

whether the lower annealing temperature would produce smaller pore sizes. The time of

annealing for all blends was 0, 30, 60, or 120 minutes, after which time they were quenched in

liquid nitrogen. The annealed blends were machined into cylinders (Ø=4mm) and further cut

into pieces, approximately 5mm in length. The PS phase was then extracted by selective

dissolution in cyclohexane. The extractions were performed initially in Falcon tubes and then

continued in glass vials at room temperature until the weight of the dried samples was

constant. The resulting porous, cylindrical samples were used for the experiments described in

Sections 3.6-3.8.

As a direct consequence of the co-continuity of the blends and the extraction method

employed to remove the porogen phase, the resulting porosity can be considered to be fully

interconnected. Approximate values for the porosity of the scaffolds were calculated based on

the mass change due to PS extraction:

%100% ×−

=initial

finalinitial

m

mmporosity (5.1)

where minitial and mfinal are the mass of the specimen before and after PS extraction,

respectively.

5.3.5 SEM imaging

For SEM imaging, annealed samples were cut to appropriate size using a sharp razor blade

and then microtomed to achieve a smooth, flat surface. The microtome cuts were performed

with a Leica RM 2155 in a liquid nitrogen cell (LN21) using glass knives. Samples were

sputter coated in Au/Pd and SEM imaging was carried out in a JSM-840-A JEOL (V=5kV,

unless otherwise indicated). Pore diameters were measured by image analysis using Sigma

Scan Pro software. The surface regions of cell-seeded porous scaffolds were imaged in the

same fashion.

5.3.6 Mechanical properties

Cylindrical samples of porous P(100/0) and P(70/30) were prepared as detailed above and

tested in the wet state at room temperature, following 48h immersion in distilled water. The

compression strength of the porous samples was measured using an Instron apparatus. The

Page 124: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

107

loading rate was set to 1mm/min using a 1000N load cell. For each condition, 3-5 samples

were tested. Results are presented as mean values ± standard deviation.

5.3.7 Permeability of scaffolds

The permeability of P(100/0) and P(70/30) scaffolds resulting from the annealing of the

blends at 200˚C and 180˚C, respectively, was determined by measuring the flow of distilled

water through cylindrical samples. The samples were fixed in a flexible silicone tube (inner

diameter < 4mm) connected to a water reservoir. All samples were immersed in distilled water

24h prior to the experiment to limit the effects of swelling and water absorption, which may

disturb the flow profile. Once mounted in the experimental set up, water was allowed to flow

through the samples until a uniform flow was obtained. Finally, Darcy’s law was applied to

calculate the permeability21:

PA

Ql

∆=

µκ (5.2)

where Q is the volumetric flow rate (m3/s), l is the sample length (m), µ is the fluid viscosity

(1mPas for water), ∆P is the pressure drop across the specimen (Pa) and A is the cross-

sectional area (m2) of the specimen. Three samples per copolymer and annealing condition

were used in this experiment, and the flow through each sample measured at two different ∆P.

The data is presented as mean values of κ ± standard deviation.

5.3.8 Protein immobilization and smooth muscle cell seeding

AMINOLYSIS AND PROTEIN IMMOBILIZATION

Cylindrical, porous P(100/0) was used to evaluate surface modification and cell adhesion

using a simple apparatus constructed in-house to carry out chemical activation and cell

seeding of the produced scaffolds. The device (see Fig. 5.1) consisted of a 1mL syringe

(i.d.=5mm), in which a scaffold was fixed using two tight-fitting Teflon rings (o.d.=5mm,

i.d.=3mm). Two pistons were used in each syringe to enable and to control flow through the

scaffolds. For each rinsing step and change of solution, the new solution was added and

flushed several times through the fixed samples (from side to side in the syringes), aspirated to

remove traces of old solution, and fresh solution was then added a second time. In order to

facilitate cell penetration into the scaffold interior, a small hole (approximately 2mm deep)

Page 125: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

108

was carved into one end of the cylindrical samples using a hypodermic needle. Each sample

was dried and weighed prior to modification and cell seeding.

Fig 5.1: Schematic drawing of cell-seeding device, consisting of a 1mL syringe (i.d.=5mm) and two Teflon rings (o.d.=5mm, i.d.=3mm). The scaffold is shown in gray between the two Teflon rings. The small cavity in the seeding end (large arrow) of the scaffold serves to facilitate the distribution of suspended cells into the interior of the scaffold. For clarity, the second piston has been left out.

The scaffolds were initially treated with 10mg/ml 1,6-hexanediamine in isopropanol for 2h

to introduce primary amines on the external as well as internal surfaces. The aminolyzed

samples were rinsed in large amounts of distilled, sterile-filtered water, and subsequently

incubated in 10mM dimethylpimelimidate (DMP) in PBS, pH 8, for 1h at room temperature.

Following a rinsing step in PBS, pH 8, a fibronectin solution (10µg/mL) was added and

thoroughly flushed through the samples to ensure thorough dispersion, and allowed to react

for 4h at 4˚C and 1h at room temperature. Finally, the scaffolds were sterilized in a 500U/ml

penicillin/streptomycin overnight.

Amine-reactive ninhydrin was used as a colorimetric indicator of the presence of amine

groups on exterior and interior surfaces of the scaffolds. Aminated scaffolds were cut into

slices using a sharp razor blade, and allowed to react for 1min in a 1M ninhydrin solution. The

slices were then placed in separate glass vials and heated to 70˚C in a water bath to accelerate

the reaction. The appearance and distribution of blue color on the samples were considered

qualitative proof of successful aminolysis.

CELL CULTURE AND ADHESION TEST

Human Umbilical Artery Smooth Muscle Cells (SMC) (Promocell, Heidelberg,

Geramany) were cultured in basic growth medium recommended by the manufacturer,

supplemented with basic fibroblast growth factor, epithelial growth factor, insulin, 5% fetal

Teflon rings Piston

Scaffold

Page 126: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

109

bovine serum, and antibiotics (100U/mL penicillin and 100mg/mL streptomycin). Medium

was changed every two to three days, and cells were used for experiments in passage 6.

Using the same chamber, sterilized scaffolds were rinsed twice in PBS, immersed in

complete medium, and equilibrated for 30 minutes at 37˚C. Subconfluent SMC were detached

using trypsin/EDTA and seeded at 1.2×106 cells/mL by carefully moving the cell suspension

several times from end to end in the syringe. The syringes containing the scaffolds were

incubated for 5h in 37˚C, at which time the cell suspension was carefully replaced by fresh

medium.

The number of adhering cells was verified 7h after seeding using a haemacytometer. Prior

to cell detachment, the scaffolds were transferred to new syringes and rinsed twice in PBS.

Cells were subsequently detached with trypsin/EDTA and vigorously flushed to extract all

adherent cells. The total cell number was determined from the medium value of eight counted

fields, and three scaffolds of each pore size were analyzed.

In a separate set of identical scaffolds, adhering cells were visualized by SEM imaging.

Cells were fixed in 1.4% glutaraldehyde for 30 minutes and in 2.5% glutaraldehyde at 4˚C

overnight. Following post-fixing in OsO4, the cell-scaffold constructs were dried in a vacuum

oven and finally sputter-coated in Au/Pd prior to imaging.

Page 127: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

110

5.4 Results and Discussion

5.4.1 Effects of annealing on pore size

In this study, the use of a co-continuous polymer blend technique was investigated as the

first step in the fabrication of macroporous scaffolds from two co-polymers. The resulting

materials can be seen in Figure 5.2, showing SEM micrographs of porous P(100/0), P(70/30),

and PS (with the P(100/0) phase extracted). These micrographs show interconnected pores,

regardless of which phase is extracted, signifying that these blends are immiscible.

Fig 5.2: SEM images confirming continuous porosity in non-annealed samples of P(100/0) (A) and P(70/30) (B) with the polystyrene phase extracted. In (C) is shown porous polystyrene after extraction of P(100/0) using acetic acid. The white scale bars indicate 10 µm.

Following the verification of the applicability of the melt blend technique, the effect of

static annealing on the pore sizes of the two co-polymer/PS blends was studied to determine

whether pore size could be efficiently controlled through this method. Since in these

structures, the pores may be better described as channels, the mention of “pore size” in the

discussion below will correspond to the diameter of these channels. During annealing,

coalescence of each of the phases results in larger pore sizes following the subsequent

dissolution of the porogen phase. This change in morphology during annealing is influenced

by parameters such as temperature, time, interfacial tension, and viscosity of the

components[23,24]. Yuan and Favis proposed that the driving force for this phase coarsening

during static annealing is driven by capillary instability[24]. Considering the interconnected

networks of each of the components to be composed of thin and thick, cylindrical rods, they

further described this process as one where the thinner rods merges/flows into thicker ones,

eventually resulting in the retraction or break-up of the thin rods. The coarsening of one

phase due to coalescence reduces the interfacial area, and hence the free energy of the

system[24].

Page 128: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

111

SEM micrographs of the P(100/0) scaffolds annealed for 30-120min at 160˚C and 200˚C

can be seen in Figure 5.3. The resulting pore sizes as determined by image analysis can be

found in Table 5.1. As expected, the higher temperature induced a more rapid coalescence of

each of the two phases, resulting in important pore growth over two hours. The pore size

found in non-annealed porous P(100/0) was 1.3±0.4µm, and this increased to reach

354±202µm after two hours of annealing at 200˚C. Lowering the annealing temperature to

160˚C, which is just above the critical flow temperature of PS, resulted in a mean pore

diameter of 17±7µm after two hours annealing. The pore size of non-annealed porous

P(70/30) was determined to be 3.2±1.2µm. Upon annealing for up to 120min at 180˚C, mean

pore sizes grew as shown in Fig. 5.4. Due to the method employed, only a medium value of

the pore size is obtained. However, it has been previously shown that porous structures

obtained from co-continuous blends have unimodal pore-size distribution[34,35]. Altogether,

these results show that by varying the time and temperature of annealing, efficient control of

pore size was accomplished in both copolymer/PS systems studied.

Table 5.1: Pore size and approximate porosity as a function of annealing time and temperature.

Pore diameters, obtained via SEM image analysis, are expressed as the mean value ± standard

deviation. The number of measurements made is indicated in the N0 column for each condition.

Co-polymer

Temperature (˚C)

P(70/30)

180

P(100/0)

200

P(100/0)

160

Annealing

time (min)

Pore

diameter

(µm)

No Porosity

(%)

Pore

diameter

(µm)

No Porosity

(%)

Pore

diameter

(µm)

No

0 3.2±1.2 110 49 1.3±0.4 201 48 - -

30 23±11 200 53 85±36 113 54 7±3 201

60 37±18 135 54 137±51 112 54 12±5 202

120 47±22 101 59 354±202 36 52 17±7 204

Page 129: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

112

A B C

D E F

A B CA B C

D E FD E F

Fig 5.3: Porous P(100/0): effect of annealing on pore size. The samples were annealed at 160˚C (A-C) or at 200˚C (D-F) for 30min (A, D), 60min (B, E), or 120min (C, F). The white scale bars indicate 100µm.

Fig 5.4: Porous P(70/30): effect of annealing on pore size. The samples were annealed at 180˚C for 30min (A), 60min (B), 120min (C). The white scale bars indicate 100µm.

It may be noted that compared to several other scaffold fabrication methods, co-

continuous blend technology leads to a porosity with more curved pore walls and unimodal

size distribution[34,35]. The pore morphology of the resulting scaffolds is thus qualitatively

different, and may potentially play a role in the performance of the respective scaffolds, e.g.

with respect to cell-material interactions.

Page 130: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

113

5.4.2 Compressive strength

Figure 5.5 shows representative compression curves of porous P(100/0) and P(70/30) as a

function of annealing time (annealing temperatures were 200˚C and 180˚C, respectively). The

compressive Young’s moduli were determined from the linear region and are presented in

Figure 5.6 together with the compressive strength at 10% strain. Both graphs reveal similar

trends: the compressive strength of both copolymers decreased significantly as a result of

annealing. The Young’s modulus of P(100/0) decreased from 69±15 MPa for non-annealed

samples to 28 ± 8 MPa following 30min annealing. For the corresponding interval in P(70/30)

scaffolds, the modulus dropped from 12±3 to 5.9±1.0 MPa. No significant change in

compressive modulus was observed between 30 and 120min annealing in P(100/0) scaffolds,

whereas P(70/30) showed a decrease in modulus from 5.9±1.0 to 2.3±0.3 MPa.

The observed decrease in compressive strength with increasing annealing time/channel

diameters is somewhat confusing. In contrast to most other methods used in the fabrication of

porous scaffolds, the “wall thickness” increases with increasing pore size as a result of the

coarsening of the two phases during annealing. Normally, such a change in morphology would

cause an increase in mechanical strength22. However, there are several other factors that may

come into play and explain our results. For example, following their preparation in the internal

mixer the polymer blends were allowed to slowly cool down to room temperature, while all

annealed samples were quenched in liquid N2 at the end of the treatment. It is therefore quite

possible that the annealed samples were less crystalline than their non-annealed counterparts,

and therefore exhibited lower compressive strength. This would also explain the small

differences in compressive modulus observed for the varying times of annealing (t=30, 60,

and 120min). In addition, increasing annealing time caused a certain increase in the porosity

of the scaffolds (see Table 1), which is consistent with decreasing mechanical strength21.

Furthermore, any degradation of the co-polymer during annealing would similarly have

resulted in a decrease in mechanical resistance.

It was also noted that P(70/30) samples showed a great capacity to partially regain their

initial shape following 60% compression. Furthermore, this capacity increased with increasing

time of annealing. Non-annealed samples were found to recover to approximately 80% of

their initial length, whereas samples that had been annealed for 120min showed up to 95%

recovery. P(100/0) samples showed a similar trend with an approximate recovery of 50% and

70% of the initial length for non-annealed and annealed samples, respectively.

Page 131: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

114

Strain

0,0 0,2 0,4 0,6 0,8 1,0

Str

ess

(MP

a)

0

5

10

15

20

25

30

P(100/0) 0minP(100/0) 120minP(70/30) 0minP(70/30) 120min

Fig 5.5: Representative stress-strain curves for porous P(100/0) and P(70/30). One curve is shown for each co-polymer for annealing times of 0 and 120 minutes.

Annealing time (min)

0 30 60 90 120

Str

ess

at

10

% s

tra

in (

MP

a)

0

1

2

3

4

5

6

P(100/0)P(70/30)

Annealing time (min)

0 30 60 90 120

Com

pre

ssiv

e m

od

ulu

s (M

Pa)

0

20

40

60

80

100

P(100/0)P(70/30)

A) B)

Fig 5.6: Compressive Young’s moduli (A) and stress at 10% strain (B) of porous P(100/0) and P(70/30) for different times of annealing.

5.4.3 Permeability of scaffolds

The intrinsic permeability of a porous material is a measurement of the ease with which a

fluid flows through it. In most tissue engineering applications, scaffolds should be designed to

achieve the most efficient mass transport properties[36]. This property can be determined

using Darcy’s law[1,37,38], which states that the flow rate through a sample is proportional to

Page 132: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

115

the applied pressure. In an earlier study, Hui et al.[39] found a threshold level of fluid

conductance below which bone xenografts in rabbit models failed to connect to the host tissue.

(Fluid conductance is the ratio of induced flow to pressure drop across the specimen). In fact,

the permeability is an indicative measure of several parameters, including porosity, pore size,

interconnectivity, and orientation of pores1, and may thus offer a simple means to obtain

relevant information related to specific scaffolds. For this study, the permeability of porous

P(100/0) and P(70/30) was measured for different pore sizes. The values of the permeability

constant, κ, as a function of annealing time (i.e. pore size) are presented in Figure 5.7.

Annealing time (min)

0 30 60 90 120

log (

κκ κκ)

-18

-16

-14

-12

-10

-8

P(100/0)P(70/30)

Fig 5.7: A log-linear plot of the permeability constant κ (m2) vs. annealing time for porous P(100/0) and P(70/30).

From Figure 5.7, it may be seen that the intrinsic permeability of the specimens tested

spanned seven orders of magnitude, ranging from 8.5 × 10-16 to 6.7 × 10-11 m2. P(100/0)

scaffolds were consistently more permeable than P(70/30) for the corresponding time of

annealing, which may be expected in view of the larger pore sizes of the former. Increasing

the annealing time from 30 to 120min, induced an 18-fold increase in P(70/30) scaffolds,

whereas the same interval caused a 70-fold increase in P(100/0) samples. For comparison, the

permeability of P(100/0) samples, annealed 120min, falls within the range of values (2.7 × 10-

11 to 2.0 × 10-8 m2) reported in the literature for human vertebral and femoral trabecular

bone[40].

In an attempt to identify key scaffold properties useful in the prediction of scaffold

performance in vivo, permeability alone and the permeability/porosity ratio has been

Page 133: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

116

suggested to be a more suitable property than pore size distribution or pore size alone[1]. We

made observations in support of this proposal, while comparing permeability values obtained

for porous structures fabricated with different techniques. For example, Shimko and

Nauman[37] reported the permeability of poly(methyl methacrylate) scaffolds fabricated using

the solvent casting/particle leaching method. The constructs were prepared with salt porogens

in the size range of 200-700µm and a final porosity of 70%. And although both pore size and

porosity were significantly higher for all but the P(100/0) 120min samples, the permeability

range obtained (6.6 × 10-16 to 1.4 × 10-10 m2) was comparable to that found for the scaffolds

tested here. Conversely, calcium phosphate scaffolds for bone reconstruction that exhibited

pore morphologies very similar to ours (pore sizes 200-700µm, porosity >70%), showed

permeability on the order of 10-10 m2.[1] Altogether, these results may serve to illustrate the

importance of the high degree of interconnectivity achieved by binary blends, resulting in a

microstructure that is more conducive to fluid flow.

5.4.4 Smooth muscle cell adhesion on protein-modified scaffolds

Covalent immobilization of fibronectin (Fn) was shown in a previous study to drastically

improve smooth muscle cell (SMC) adhesion and proliferation on P(100/0) films[28].

Compared to unmodified films and films exhibiting amine groups only, covalently

immobilized Fn improved cell adhesion 3.5 and 1.9 times, respectively. Fn-modified surfaces

were furthermore able to support 5-fold increase in cell numbers over six days in culture,

whereas no proliferation was observed on the two control surfaces. Adsorbed Fn was

similarly found to enhance cell adhesion on this material, but did not to the same extent

support proliferation (2-fold increase over six days)[28].

In this study, the same surface modification scheme was applied to promote cell

attachment in 3D matrices. Utilizing a custom-made reaction chamber (see Fig. 5.1), designed

to facilitate homogeneous surface modification and subsequent cell seeding, cylindrical

P(100/0) scaffolds were aminated through a reaction with 1,6-hexanediamine. Ninhydrin was

used as a colorimetric indicator of the presence of amine groups on the scaffold exterior and

interior surfaces. The successful and thorough aminolysis of the samples was confirmed

visually by the homogeneous, blue staining of several cross-sections of the two scaffold types

(Fig. 5.8). Following aminolysis, the scaffold surfaces were activated with dimethyl

pimelimidate (an amino-reactive cross-linker) to couple Fn covalently to the modified

surfaces. In addition, scaffolds of two different pore sizes (30 vs. 60min, annealing

Page 134: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

117

temperature = 200˚C) were used to observe the effect of pore size on the efficiency of cell

seeding.

Fig 5.8: (A) Aminated P(100/0) scaffold cross-sections stained blue by ninhydrin reactant. (B) Un-modified scaffolds were not stained by the ninhydrin treatment. Each scaffold was cut with a razor blade and numbered as indicated in (C).

SEM images of adhering cells bridging a pore gap and spreading out on a pore wall close

to the scaffold surface can be seen in Figure 5.9. Figure 5.10 shows the number of adhering

cells in the two types of scaffold 7h after seeding. Specifically, increasing the pore size from

85 to 137µm improved cell numbers from approximately 400 to 1500 cells/mg of scaffold.

Given the similar geometry of the samples, these results indicate a more efficient penetration

into the bulk of the scaffolds exhibiting larger pores. Seeing as no difference in porosity (void

volume) was observed for these two preparations (Table 5.1), the increase in cell number was

a direct effect of enhanced diffusion caused by larger pore sizes. This effect becomes ever

clearer when considering the smaller surface area available for cell adhesion in samples of

larger pore sizes. The obtained cell seeding efficiency was 43±4% and 17±10% for the

“60min” and “30min” samples, respectively, whereas it was a mere 4±2% for aminated

0123

C)

0123

C)

Page 135: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

118

“30min” samples. Keeping in mind that rather large quantities of cells were visible in the

space outside the scaffolds in the reaction chambers after 7h incubation, these results can be

considered as good. It is expected that by automating the chamber setup so as to achieve flow

of the cell-seeding suspension across the scaffolds throughout the adhesion period, seeding

efficiency should be significantly improved.

Fig 5.9: Scanning electron micrographs of SMC adhering in a P(100/0) scaffold, 8h after cell seeding.

Cells spanning a pore void (A) and cells spread out on the Fn-modified copolymer surface (B) are

indicated by white arrows.

Cel

l num

ber/

mg

scaf

fold

0

500

1000

1500

2000

P(100/0) 60minP(100/0) 30minP(100/0) 30min, no Fn

Fig 5.10: Number of SMC adhering on surface-modified porous P(100/0) scaffolds 7h after cell

seeding. The scaffolds were previously annealed for 30min or 1h (n=3).

Our results show that biochemical surface modification within three-dimensional is

possible using the methods described. Furthermore, as reported elsewhere[18], pore size was

found an important parameter that influences the efficiency of dynamic in vitro cell seeding

Page 136: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

119

techniques. The data presented here are certainly in support of these findings, although further

investigations will be needed to more accurately establish optimal pore diameters for cell

seeding and culture.

5.5 Conclusions

Scaffolds with entirely interconnected porosity were prepared from binary co-

polymer/polystyrene blends and investigated for their applicability in tissue engineering. The

original mean pore sizes of P(100/0) and P(70/30) were 1.3 and 3.2µm, respectively. Pore

sizes could be efficiently controlled by means of static annealing for 30-120min, resulting in

pore sizes roughly in the range of 15-350µm. The efficiency of in vitro seeding of smooth

muscle cells was investigated in fibronectin-modified P(100/0) scaffolds of two different pore

sizes. Considerably higher numbers of adhering cells were found in the scaffolds with larger

pore sizes, indicating the importance of this parameter to promote effective cell intrusion into

bulk materials. The number of adhering cells was also indicative of successful protein

immobilization.

Furthermore, the influence of scaffold pore size on compressive strength, permeability

was investigated. Compressive moduli ranged from 2.3±0.3 to 67±15 MPa and decreased with

increasing pore size. The reverse trend was found for scaffold permeability, which generated

values of κ from 8.5 × 10-16 to 6.7 × 10-11 m2. Importantly, in comparison with other scaffolds

that were prepared by particulate leaching and showing larger pore size and higher porosity,

the scaffolds studies here showed similar permeability, clearly indicating the pivotal role of

pore morphology.

Especially, with regard to the permeability and successful cell adhesion tests, we

propose that these scaffolds may be well suited for future studies of three-dimensional tissue

culture in vitro.

5.6 Acknowledgements

This work was supported by the NSERC through a Discovery grant (PV) and a

Strategic grant (PV) and a post-doctorate fellowship (PS). Appreciation is extended to Dr.

Martin Bureau and Manon Plourde, at Industial Materials Institute (NRC), Boucherville, QC,

for the use of and assistance with testing of mechanical properties.

Page 137: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

120

5.7 References

(1) Li SH, de Wijn JR, Li JP, Layrolle P, de Groot K. Macroporous biphasic calcium phosphate scaffold with high permeability/porosity ratio. Tissue Eng. 2003; 9:535-548.

(2) Maquet V, Jerome R. Design of macroporous biodegradable polymer scaffolds for cell transplantation. Mater. Sci. Forum 1997; 250:15-42.

(3) Griffith L. Emerging design principles in biomaterials and scaffolds for tissue engineering. Ann. N. Y. Acad. Sci. 2002; 961:83-95.

(4) Mikos AG, Thorsen AJ, Czerwonka LA, Bao Y, Langer R, Winslow DN, Vacanti JP. Preparation and characterization of poly(L-lactic acid) foams. Polymer 1994; 35:1068-1077.

(5) McGlohorn JB, Holder WD, Grimes LW, Thomas CB, Burg KJL. Evaluation of smooth muscle cell response using two types of porous polylactide scaffolds with differing pore topography. Tissue Eng. 2004; 10:505-514.

(6) Draghi L, Resta S, Pirozzolo MG, Tanzi MC. Microspheres leaching for scaffold porosity control. J. Mater. Sci.: Mater. Med. 2005; 16:1093-1097.

(7) Harris LD, Kim BS, Mooney DJ. Open pore biodegradable matrices formed with gas foaming. J. Biomed. Mater. Res. 1998; 42:396-402.

(8) Whang K, Thomas CH, Healy KE, Nuber G. A Novel Method to Fabricate Bioabsorbable Scaffolds. Polymer 1995; 36:837-842.

(9) Mooney DJ, Baldwin DF, Suh NP, Vacanti LP, Langer R. Novel approach to fabricate porous sponges of poly(D,L-lactic-co-glycolic acid) without the use of organic solvents. Biomaterials 1996; 17:1417-1422.

(10) Park A, Wu B, Griffith LG. Integration of surface modification and 3D fabrication techniques to prepare patterned poly(L-lactide) substrates allowing regionally selective cell adhesion. J. Biomater. Sci.: Polym. Ed. 1998; 9:89-110.

(11) Liu XH, Ma PX. Polymeric scaffolds for bone tissue engineering. Ann. Biomed. Eng. 2004; 32:477-486.

(12) Mikos AG, Temenoff JS. Formation of highly porous biodegradable scaffolds for tissue engineering. Electr. J. Biotechnol. 2000; 3:114-119.

(13) Sachlos E, Czernuszka J. Making tissue engineering scaffolds work. Review on the application of solid freeform fabrication technology to the production of tissue engineering scaffolds. Eur. Cells Mater. 2003; 5:29-40.

(14) Yang SF, Leong KF, Du ZH, Chua CK. The design of scaffolds for use in tissue engineering. Part 1. Traditional factors. Tissue Eng. 2001; 7:679-689.

Page 138: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

121

(15) Kim BS, Nikolovski J, Bonadio J, Mooney DJ. Cyclic mechanical strain regulates the development of engineered smooth muscle tissue. Nat. Biotechnol. 1999; 17:979-983.

(16) Oberpenning F, Meng J, Yoo JJ, Atala A. De novo reconstitution of a functional mammalian urinary bladder by tissue engineering. Nat. Biotechnol. 1999; 17:149-155.

(17) Lee SH, Kim BS, Kim SH, Choi SW, Jeong SI, Kwon IK, Kang SW, Nikolovski J, Mooney DJ, Han YK, Kim YH. Elastic biodegradable poly(glycolide-co-caprolactone) scaffold for tissue engineering. J. Biomed. Mater. Res., Part A 2003; 66A:29-37.

(18) Zeltinger J, Sherwood JK, Graham DA, Mueller R, Griffith LG. Effect of pore size and void fraction on cellular adhesion, proliferation, and matrix deposition. Tissue Eng. 2001; 7:557-572.

(19) Kim BS, Nikolovski J, Bonadio J, Smiley E, Mooney DJ. Engineered smooth muscle tissues: Regulating cell phenotype with the scaffold. Exp. Cell Res. 1999; 251:318-328.

(20) Washburn NR, Simon CG, Tona A, Elgendy HM, Karim A, Amis EJ. Co-extrusion of biocompatible polymers for scaffolds with co-continuous morphology. J. Biomed. Mater. Res. 2002; 60:20-29.

(21) Reignier J, Huneault MA. Preparation of interconnected poly(ε-caprolactone) porous scaffolds by a combination of polymer and salt particulate leaching. Polymer 2006; 47:4703-4717.

(22) Sarazin P, Roy X, Favis BD. Controlled preparation and properties of porous poly(L-lactide) obtained from a co-continuous blend of two biodegradable polymers. Biomaterials 2004; 25:5965-5978.

(23) Sarazin P, Favis BD. Morphology control in co-continuous poly(L-lactide)/polystyrene blends: A route towards highly structured and interconnected porosity in poly(L-lactide) materials. Biomacromolecules 2003; 4:1669-1679.

(24) Yuan ZH, Favis BD. Coarsening of immiscible co-continuous blends during quiescent annealing. AIChE Journal 2005; 51:271-280.

(25) Na YH, He Y, Shuai X, Kikkawa Y, Doi Y, Inoue Y. Compatibilization effect of poly(ε-caprolactone)-b-poly(ethylene glycol) block copolymers and phase morphology analysis in immiscible poly(lactide)/poly(ε-caprolactone) blends. Biomacromolecules 2002; 3:1179-1186.

(26) Washburn NR, Weir M, Anderson P, Potter K. Bone formation in polymeric scaffolds evaluated by proton magnetic resonance microscopy and X-ray microtomography. J. Biomed. Mater. Res., Part A 2004; 69A:738-747.

(27) Bramfeldt H, Sarazin P, Vermette P. Characterization, degradation, and mechanical strength of poly(D,L-lactide-co-ε-caprolactone)-poly(ethylene glycol)-poly(D,L-lactide-co-ε-caprolactone). J. Biomed. Mater. Res., Part A 2007; 83A: 503-511.

Page 139: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

122

(28) Bramfeldt H, Vermette P. Enhanced smooth muscle cell adhesion and proliferation on protein-modified caprolactone-based copolymers. J. Biomed. Mater. Res., Part A, accepted.

(29) Lebaron RG, Athanasiou KA. Extracellular matrix cell adhesion peptides: Functional applications in orthopedic materials. Tissue Eng. 2000; 6:85-103.

(30) Hu YH, Winn SR, Krajbich I, Hollinger JO. Porous polymer scaffolds surface-modified with arginine-glycine-aspartic acid enhance bone cell attachment and differentiation in vitro. J. Biomed. Mater. Res., Part A 2003; 64A:583-590.

(31) Pompe T, Kobe F, Salchert K, Jorgensen B, Oswald J, Werner C. Fibronectin anchorage to polymer substrates controls the initial phase of endothelial cell adhesion. J. Biomed. Mater. Res., Part A 2003; 67A:647-657.

(32) Serrano MC, Pagani R, Manzano M, Comas JV, Portoles MT. Mitochondrial membrane potential and reactive oxygen species content of endothelial and smooth muscle cells cultured on poly(ε-caprolactone) films. Biomaterials 2006; 27:4706-4714.

(33) Lieb E, Hacker M, Tessmar J, Kunz-Schughart LA, Fiedler J, Dahmen C, Hersel U, Kessler H, Schulz MB, Gopferich A. Mediating specific cell adhesion to low-adhesive diblock copolymers by instant modification with cyclic RGD peptides. Biomaterials 2005; 26:2333-2341.

(34) Sarazin P, Favis BD. Stability of the co-continuous morphology during melt mixing for poly(epsilon-caprolactone)/polystyrene blends. J. Polym. Sci.: Polym. Phys. 2007; 45:864-872.

(35) Yuan ZH, Favis BD. Macroporous poly(L-lactide) of controlled pore size derived from the annealing of co-continuous polystyrene/poly(L-lactide) blends. Biomaterials 2004; 25:2161-2170.

(36) Martin Y, Vermette P. Bioreactors for tissue mass culture: Design, characterization, and recent advances. Biomaterials 2005; 26:7481-7503.

(37) Shimko DA, Nauman EA. Development and characterization of a porous poly(methyl methacrylate) scaffold with controllable modulus and permeability. J. Biomed. Mater. Res.: Appl. Biomater. 2007; 80B:360-369.

(38) Grimm MJ, Williams JL. Measurements of permeability in human calcaneal trabecular bone. J. Biomechanics 1997; 30:743-745.

(39) Hui PW, Leung PC, Sher A. Fluid conductance of cancellous bone graft as a predictor for graft-host interface healing. J. Biomechanics 1996; 29:123-132.

(40) Nauman EA, Fong KE, Keaveny TM. Dependence of intertrabecular permeability on flow direction and anatomic site. Ann. Biomed. Eng. 1999; 27:517-524.

Page 140: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

123

GENERAL CONCLUSION AND SUGGESTED FUTURE WORK

GENERAL CONCLUSION

This thesis project was designed to constitute the first part of a work whose final

objective is the creation of functional vascular beds within three-dimensional scaffold

constructs for tissue engineering. It is the author’s expectation, that the methods and results of

this work may constitute a helpful knowledge base and direct future work towards the ultimate

objective.

Briefly, this thesis has resulted in the synthesis of a series of biocompatible co-polymers,

whose material properties vary from highly crystalline, mechanically strong, and slow-

degrading to highly amorphous, malleable and fast-degrading, as specified in Chapter 2.

These co-polymers may be further manipulated with relative ease to obtain materials

possessing characteristics appropriate for a specific application. This can be achieved either

by changing the monomer composition and/or molecular weight of the copolymers, or, as

shown in Chapter 3, by simple blending strategies. In addition, by preparing binary blends

composed of one fast-degrading and one slow-degrading phase, porous materials may

develop as hydrolysis proceeds.

Two of the characterized co-polymers (P(100/0) and P(70/30)) were considered

interesting candidates for scaffold fabrication. In order to improve cell-material interactions

on these materials, an aminolysis technique that was previously applied to pure PCL films,

was here applied to the described co-polymers, thereby introducing reactive amine groups on

the surface. Following aminolysis, films were further modified by covalent protein

immobilization using dimethylpimelimidate. This method strongly enhanced material

interactions with smooth muscle cells (SMC); specifically adhesion and proliferation, while

preserving a smooth muscle cell phenotype, although very different results were obtained for

the two co-polymers. It was found necessary for the LA-conainting P(70/30) to be pre-wet

prior to aminolysis and protein immobilization in order to increase cell numbers on these

surfaces. However, cell numbers were still not comparable to the ones obtained for P(100/0).

As for the different proteins preparations of this study, results implied that covalently

immobilized fibronectin and fibrin layers were the most suited for medium- to long-term

culture.

Page 141: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

124

The same two co-polymers were investigated for the applicability of immiscible binary

blends for the fabrication of porous scaffolds with highly interconnected porosity. Static

annealing was found to constitute an efficient means to increase pore sizes. Cylindrical

scaffolds were then tested for compressive strength and permeability, and were both affected

by pore size. Encouraging preliminary results of smooth muscle cell adhesion in cylindrical

scaffolds, suggested the successful translation of the technique used for two-dimensional

surface modification to this three-dimensional scaffold. The results for permeability and SMC

adhesion within these three-dimensional co-polymer scaffolds indicated the pivotal role of

sufficient pore size. Taken together, the porous scaffolds developed in this thesis were judged

as promising for further tissue-culture studies in a bioreactor environment.

SUGGESTED FUTURE WORK

Based on the results of the studies described in this thesis and the many question marks

and ideas that have appeared along the way, the following work is proposed for the further

advancement of the project:

• Carry out a detailed surface study to identify and solve the problems encountered

during attempts to immobilize proteins on P(70/30) to ameliorate cell adhesion and

proliferation on these films.

o The ninhydrin assay needs to be repeated on aminated pre-wet films.

o The ninhydrin assay may also be useful to quantify free NH2 goups on film

surfaces following covalent protein immobilization, both on P(100/0) and

P(70/30). This method may then be used to assess the success rate of protein

immobilization, which may be interesting to compare between materials, and

between dry and pre-wet P(70/30).

o An alternative route may be to detach immobilized Fn by enzymatic cleavage

(e.g. by trypsination) and quantify the protein content by a Lowry or total

protein assay. The digestion would of course reduce the polymers into peptides

and amino acids depending on the time of digestion. In the Lowry assay, the

Folin-Ciocalteu’s phenol reagent reacts with mainly tyrosine residues in the

protein, and results in an absorbance maximum at 750nm. Yet another

Page 142: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

125

technique is the quartz-crystal microbalance with dissipation monitoring.

Although coating a crystal with co-polymer may be rather difficult, amine-

terminated monolayers can be generated on gold-coated crystals, on which the

subsequent surface-modification steps (DMP-activation and protein-

immobilization) can be carried out. The conformational state may then easily

be evaluated using specific antibodies.

o One hypothesis is that the P(70/30) films have PEG chains oriented towards the

surface to a higher extent than P(100/0). If that is the case the P(70/30) films

are less reactive than P(100/0). High resolution XPS may perhaps be useful on

thinner and flatter films, prepared e.g. by spin-coating. Solid state NMR may

also be used to determine molecular structure and intermolecular packing

patterns.

o The conformation of the immobilized proteins (adsorbed and covalently

coupled) may be investigated using AFM. Try a trial and error-approach using

different monoclonal antibodies towards Fn may turn the previously

unsuccessful attempt to visualize the protein using immunofluorescence

around.

• Further characterization of the phenotype of smooth muscle cells grown on protein-

modified surfaces. This may also be further developed to promote dedifferentiation/

differentiation by growth factor delivery. Of interest would be the stimulation of 1) a

migrating phenotype in order to populate the scaffold bulk with SMC, and 2) a

contractile phenotype, as found in native blood vessels, and other mature SMC-based

tissues. This study should first be carried out in copolymer films, and later on in 3D

scaffolds. Suggested growth factors are: TGF-alpha, which acts to increase SM alpha-

actin. (Riha et al., Surgery 2007; 141, p.394-402): A combination of PDGF-BB and

IL-1 beta, which together act to lower levels of the contractile markers alpha-actin,

myosin heavy chain, and calponin, thereby possibly stimulating a migrating and ECM-

producing phenotype. (Chen et al., Proceedings Natl Acad Sciences 2006; 103, p.

2665-2670). VEGF and bFGF stimulate migration in SMC (Couper et al. Circulation

Res 1997; 81, p.932-939; Wang and Keiser, Circulation Res 1998; 83, p.832-840).

Page 143: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

126

• Characterization and promotion of ECM deposition by SMC cultured on prepared

surfaces and scaffolds. Suitable techniques are e.g. Western blots (protein) and

Northern blots (gene expression, RNA). Suitable ECM proteins are Carry out a

comparative study of immobilized adhesion peptides vs. whole proteins to identify

effects on cell adhesion, proliferation, and ECM production. Interesting growth factors

to stimulate protein production may be: TGF-beta (Collagen type I, fibronectin)

(Schmidt et al. Eur J Clin Invest 2006; 36, p. 473-482);

• Polystyrene was used as a model porogen in this work but must be replaced by a

biocompatible polymer to avoid inflammatory response to any residual PS trapped in

the 3D constructs after extraction.

• Maximize the void volume obtainable by the binary co-continuous blend technique.

This may be done by varying the volume fraction of the porogen phase. Also worth

looking into would be the parameters during melt mixing. Lee and Han (Polymer

1999; 40, p. 6277-6296) were also able to increase the range of co-continuity to 30-

70% of the porogen phase by working close to / below the critical flow temperature of

the two components.

• A study of P(100/0) and P(70/30) material properties (crystalline content, degradation,

mechanical strength) following annealing as described in Chapter 5. A degradation

study should also be carried out of 3D scaffolds, and parallel measurements of the

mechanical properties of these structures as a function of degradation time.

• Apply melt blending as a technique for preparation of the co-polymer blends described

in Chapter 3, in order to obtain co-continuous porosity as a result of differentiated

degradation. Investigate whether viable long-term cell culture may be accomplished on

such materials, and if so, if cells will invade the forming void space.

Page 144: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

127

• Combine the co-polymer blend preparation of Chapter 3 with the melt-

mixing/annealing method described in Chapter 5, in order to fabricate scaffolds of

bimodal porosity.

• In vitro: develop culture protocols and optimize conditions for tissue culture in a

perfusion bioreactor. Specifically, the effects of shear stress and soluble growth factors

on SMC phenotype should be evaluated to establish in what way smooth muscle cells

(or other cell types) may be controlled in this environment.

• Investigate the applicability of techniques like high resolution magnetic resonance

imaginng and µ-CT (microcomputed tomography) to image the interior of the

scaffolds. The µ-CT technique has been used to measure bone thickness in

regenerating bone (see e.g. Sell et al. Calcified Tissues Int 2005; 76, p.355-364) and

has also been proposed as a non-invasive monitoring system for vascularization (ref:

www.medicalnewstoday.com/articles/20243.php). It may therefore be interesting to

follow the vascularization and tissue growth within the 3D scaffolds using this

technique.

• Additionally, it would be interesting to study the possibility of growth factor

immobilization, possibly in combination with whole adhesive proteins, on scaffold

surfaces to promote particular cellular behavior. Such strategies would, if effective,

help to decrease the quantitative requirements of growth factors, and thus keep costs

down.

• In vitro: use protein-modified scaffolds, such as described in Chapter 5, to pre-culture

smooth muscle cells for a time sufficient to allow thorough cellular invasion of the

scaffold interior as well as the deposition of ECM. Subsequently, the scaffolds may be

seeded with endothelial or a mix of EC and tissue-specific cells, and cultured in a

bioreactor. Objective: to achieve initial differentiation of EC into tubular networks,

possibly induced by growth factor delivery, followed by EC-controlled recruitment of

SMC for vessel stabilization.

Page 145: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

128

• In vivo: subcutaneous implantation of surface-modified, P(100/0) and P(70/30)

sponges seeded with SMC alone or EC+SMC, to verify the biocompatibility of the

scaffolds and to find out whether vascular ingrowth may be promoted.

Page 146: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

129

APPENDIX

A1. Protocol for synthesis of P(ε-CL-co-D,L-LA)-PEG-P(ε-CL-co-D,L-LA)

Co-polymer synthesis using simultaneous, ring-opening co-polymerization of ε-CL and D,L-

LA in the presence of PEG diols (MW=7500), catalyzed by stannous octoate (SnOct).

MATERIAL

ChemGlass two- or three-neck round bottom flask (RBF), max. 250mL, 24/40 neck.

Rubber septum

Stirring shaft and blade of appropriate size

Air tight glass stop cock (with Teflon fittings and o-ring) to fix the stirring shaft

Motor with adjustable rotation speed

Robust stand and clips

Oil bath with temperature control equipped with magnetic stirrer

Equipment for degassing and purging with either N2 or Ar

STEP 1: MONOMER PREPARATION

• PEG diols: dry under vacuum at 50°C for 48h.

• ε-CL monomers: use as received

• D,L-LA monomers: recrystallize from ethyl acetate

1. Dissolve monomers in ethyl acetate in a crystallizing dish. Heat under stirring

to approximately 70°C. Carefully add solvent drop by drop until dissolution. It

is important that the minimal amount of solvent necessary to dissolve the

monomer is used.

2. Take the crystallizing dish off the hot plate and allow the solution to cool

down, and some of the solvent to evaporate (might need to be partially

covered to slow down the rate of evaporation). D,L-LA crystals will start to

form on the solvent/air interface.

3. When the recrystallization is finished, filter the crystals and wash with ethyl

acetate. Dry the crystals in room temperature and finally under vacuum.

Page 147: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

130

STEP 2: REACTION CHAMBER ASSEMBLY

1. Clean all glassware, metal, and Teflon parts, and dry under vacuum.

2. Assemble the stirrer components. The threaded glass stopper must be put in place

before attaching the stirrer blade. Insert the stirring shaft into the middle neck of the

RBF and screw the rod tight. The blade should be close to the bottom, without the

metal rod touching the bottom of the flask.

3. Close the remaining neck(s) with a rubber septum.

4. Firmly attach the top of the metal rod to the motor.

5. Place the heating/stirring plate with the oil bath on the base of the stand and mount

the motor while adjusting its height so that the RBF is in a good position in the oil

bath. Further adjustments must be made to assure that the rod is perfectly vertical, so

as not to be moving the RBF while rotating.

STEP 3: SYNTHESIS

For the synthesis of P(100/0), PEG diols and ε-CL were introduced into an RBF and melted

at 140°C. The molar ratio of ethylene glycol-to-monomer was 1:4.2. The temperature was

then lowered to 120°C and the melted mixture degassed and purged with N2. SnOct was

added at a monomer-to-catalyst ratio of 1000:1, and the polymerization was allowed to

continue for 24h at 120°C under constant stirring.

To obtain P(70/30) and P(30/70) copolymers, PEG diols and monomers were introduced into

a three-neck round-bottom flask at a molar ratio of monomer:ethylene glycol of 4.2:1. The

molar ratio of ε-caprolactone:D,L-lactide was either 70:30 or 30:70. Following melting,

degassing and N2-purging at 90°C, the temperature was briefly raised to 140°C to complete

the melting of D,L-LA and subsequently lowered to 120°C at which temperature

polymerization was carried out for 24h under constant stirring. The RBF was immersed in the

oil bath as completely as possible to avoid condensation of D,L-LA on the walls of the flask.

1. Start heating the oil bath under stirring.

2. Remove the septum and add PEG diols and monomer to the flask. Start stirring

(appropriate stirring speed is slow, approximately 60 rpm) and seal the RBF with the

septum.

Page 148: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

131

3. When all components are melted, carry out 3-4 cycles of degassing/N2 purging

through a gauge needle inserted through the septum. Stop when no more gas is being

removed from the mix, and finish by purging one last time with N2.

4. When the oil bath is stable at 120°C, inject SnOct catalyst through the septum using a

glass/steel syringe (25-100µl), making sure no air is injected. Avoid SnOct dripping

onto the flask wall. To avoid air bubbles, mount the syringe with a clip to keep it still,

and slowly aspire and eject SnOct a few times. Each time the syringe is filled, move

the piston little by little (5-10µl steps) and wait for the viscous liquid to fill the

syringe (several minutes).

5. Once the catalyst has been added to the melted monomer/diol mix, allow the synthesis

to proceed for 24h under constant stirring. As the polymerization proceeds the

reaction mixture will become increasingly viscous. It is a good idea to check on the

set up now and then. Towards the end of the reaction time it may become necessary to

stop and start the stirring, in order to let the polymer mass slide back down to the

bottom of the flask where it may be better stirred, or to stop the stirring completely to

avoid breaking the RBF.

6. At 24h, turn off the oil bath and remove it. Allow the polymer mass to cool down

(several hours or overnight).

STEP 4: RECOVERY

Polymers can be dissolved in dichloromethane in the reaction bottle, and precipitated in cold

ethanol, followed by filtration (coarse glass and standard filter paper) and drying in a vacuum

oven at 30°C until constant weight.

Page 149: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

132

Figure A.1: Setup for copolymer synthesis. Pieces for stirring are shown separately (A) and assembled (B). (C) and (D) show the setup attached to the motor unit and mounted on a stand.

A B

C D

Page 150: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

133

A2. Protocol for surface modification, cell seeding, cell count, and imaging.

Films were prepared in a hot melt press. Samples that fit into a 48-well plate can be punched

out using a 1cm diameter biopsy punch. P(100/0) is a tough and resistant material, while

P(70/30) is softer and can brake during manipulation and chemical/biochemical treatment,

especially if thin films are used (0.25mm). A gentler version of the aminolysis treatment is

performed on P(70/30) to avoid massive cracking of the surface. In the treatment of P(70/30),

the HCl protonizing step is excluded, and it is therefore recommended that any further surface

modification be carried out immediately following aminolysis.

Care has to be taken when transferring samples from one plate to another to avoid breaking

them. Make sure that films do not flip over in their wells at any time during the experiment.

This can easily occur during rinsing if glass rings are not used and also upon transfer between

wells. In the following, approximate time for each step is given in parentheses. RT = room

temperature.

STEP 1: AMINOLYSIS (5H/2H)

Instructions for P(100/0) and (P(70/30)).

Prepare: 10% (wt/wt) 1,6-hexanediamine in isopropanol (heat slightly and stir to aid

dissolution)

5mM HCl

1. Incubation in 10% (wt/wt) 1,6-hexanediamine for 2h (30 min) at RT. Rinse 3×mQ.

2. Rinse in 5mM HCl (mQ) for 1h (30 min) to protonize the aminogoups

3. Rinse 3×mQ

4. Dry in vacuum oven (pass directly to protein immobilization)

STEP 2: PROTEIN IMMOBILIZATION (2H + 0,5H+INCUBATION)

Prepare: Sterile filtration of the DMP solution (10mM, in PBS pH 8).

Sterile 48-well plates

Sterile tweezers (2)

Sterile PBS pH 7.4 d pH 8

Marked, sterile tubes for protein solutions

Page 151: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

134

Filter 0.45um and syringe for Fgn filtration

1. Incubate the films in 10mM DMP, 1h in RT. (If control surfaces are included,

remember to leave these without!) In the meantime, prepare the protein solutions (as

late as possible).

2. Aspire the liquid, and rinse twice with PBS pH 8.

3. Add the protein solutions, 300µl/well (wells with glass rings need less). Carefully push

floating films to the bottom (they will stay there).

4. Incubate 4h at 4°C and 1h in room temperature. Leave fibrinogen surfaces (for fibrin)

in RT.

5. After 2h in room temperature, aspire the fibrinogen solution (don’t rinse), add

thrombin solution (or thrombin + DMP), and incubate 3h at 37°C.

6. Rinse all surfaces in sterile PBS×2, and sterilize (step 3).

STEP 3: STERILIZATION (OVERNIGHT + 45 MIN)

1. Incubate the surfaces in PBS with a 5X concentration of antibiotics overnight at 4°C

(to conserve proteins).

2. Rinse the surfaces 2x10min PBS

3. Transfer to new sterile 48-well plates.

STEP 4: CELL SEEDING (1,5H)

Prepare: Use freshly prepared surfaces if possible.

At the suggested seeding density, have one T75 of SMC for each 48-well plate

ready for passage.

Sterile glass rings

1. Equilibrate the films in 200µl of medium for 1h at 37°C. Place sterile glass rings

where necessary.

2. Passage the cells.

3. Seed SMC at 10,000 cells/well for counting experiments, and at 5,000 cells/well for

imaging. A small Petri dish may be seeded for control fluorescent staining.

4. Change medium after 8h, and then every two-three days for longer experiments.

Page 152: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

135

STEP 5: COUNTING THE CELLS (TIME DEPENDS ON SIZE OF EXPERIMENT) OR FIXING THEM FOR

SEM IMAGING (45MIN + OVERNIGHT).

Prepare:

For counting: Centrifuge tubes (15ml) identifying the surface type and batch number

Identically marked 1,5ml tubes for Trypan Blue staining

Warm medium (or DMEM, cheaper), Trypsin-EDTA, serum

For fixing: Glutaraldehyde solutions (1,4% and 2,5%)

Cell count:

1. Add 100µl of warm serum to each of the identified centrifuge tubes.

2. Rinse the surfaces once in PBS (max 4 pools) and carefully transfer them without glass

rings to a new 48-well plate (to avoid counting cells adhering to the TCPS below).

3. Add 200µl of warm trypsin-EDTA to the samples (max. 4 pools) and incubate for 5

minutes at RT.

4. Transfer the trypsin solution of identical surfaces to the corresponding 15ml centrifuge

tube. Add 300µl of warm DMEM (or medium) to the wells to recover the remaining

cells.

5. Centrifuge SMC at 1600rpm for 4 min

6. Carefully remove the supernatant (you will probably not see the pellet) and resuspend

in 100µl of DMEM. Keep the remaining suspension in its tube.

7. Transfer the 100µl to a 1.5ml Eppendorf, add 50µl of Trypan Blue, and incubate for 7

minutes.

8. Count with hemacytometer (at least 6 fields)

9. Measure the volume of the remaining suspension in the Falcon tubes and use this to

calculate the total number of cells/surface condition.

10. Normalize the cell number to cells/area unit.

Page 153: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

136

Fixing and imaging in SEM

1. Fix the cells in 1,4% glutaraldehyde, 30 min at RT

2. Aspire the liquid and add glutaraldehyde 2,5% to fix overnight in the fridge (can be

left up to a week).

3. CHUS: Post-fixing with OsO4 and dehydration in an EtOH gradient followed by

drying in a critical point dryer.

4. Metallize the samples

5. Image in SEM, using high contrast and low current.

Fixing and fluorescent staining

Follow laboratory SOPs.

A3. Standard curve established for the ninhydrin assay used in Chapter 4.

Ninhydrin assay: standard curve

y = 5578xr2 = 0,9997

0

100

200

300

400

500

600

0,00 0,02 0,04 0,06 0,08 0,10

C (mM)

AU

(57

5nm

)

Fig. A.2: Standard curve established for ninhydrin assay.

Page 154: Université de Sherbrooke Département de génie chimique · 2010. 5. 7. · caractérisées par la microscopie électronique à balayage ainsi que par le marquage fluorescent du

137

A4. Control α-actin staining of SMC following culture on TCPS and Fn-immobilized

P(100/0) in Chapter 4.

Fig. A.3: Alpha-actin staining of cells reseeded on TCPS following six days of culture on (A) Fn-

immobilized P(100/0) and (B) TCPS.

A) B)