rights / license: research collection in copyright - …7241/eth... · year of my thesis so much...

151
Research Collection Doctoral Thesis Engineering chondrogenic micro-environments for tissue engineering applications Author(s): Mhanna, Rami Publication Date: 2013 Permanent Link: https://doi.org/10.3929/ethz-a-009921642 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: truongdieu

Post on 16-Aug-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Research Collection

Doctoral Thesis

Engineering chondrogenic micro-environments for tissueengineering applications

Author(s): Mhanna, Rami

Publication Date: 2013

Permanent Link: https://doi.org/10.3929/ethz-a-009921642

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

DISS. ETH NO. 21108

Engineering Chondrogenic Micro-

Environments for Tissue Engineering Applications

DISSERTATION

Submitted to

ETH ZURICH

for the degree of

DOCTOR OF SCIENCES

by

Rami Mhanna

M. Eng., Biomedical Eng., The University of Melbourne

born May 5th, 1982

Lebanese

Accepted on the recommendation of

Prof. Dr. Marcy Zenobi-Wong, examiner Dr. Daniel Eberli, co-examiner

Prof. Dr. Ivan Martin, co-examiner

2013

To My Parents

Often you shall think your road impassable, somber and companionless.

Have will and plod along; and round each curve you shall find a new companion.

Mikhail Naimy

Acknowledgments

I would like to express my deep gratitude to the people that helped realizing

this thesis and turning the PhD dream into reality.

A special thank you to Marcy Zenobi-Wong, my supervisor, for being always

there, supporting and responding to any question I had so quickly. I never had to worry

of sending you an abstract a few days before the deadline, I knew you would respond

with all corrections only a few hours after. Thank you very much for the support in the

last few months of my thesis in which I felt really in need for. I would also like to thank

you for your patience on my questions, initial limited cell lab experience and for

accepting my stubbornness in our meetings. Thank you for the dinners and BBQs at

your place, the lovely Spanish dinner, group retreat I still remember the pancakes. I

would like to express my deep gratitude to Janos Vörös for accepting me as a PhD

student. Thank you Janos for creating such a stimulating research and social

environment, your lab is so attractive that people do not want to leave it. Thank you

for your great input to most of the raised questions and to your subjective productive

advices. Thank you for joining the coffee, beer, soccer and volleyball games. I would

like to express my sincere appreciation to Prof. Ivan Martin and Dr. Daniel Eberli for

accepting to be my co-referees. It is a great honor for me that you review my thesis.

I would like to thank all my collaborators and colleagues in the Find and Bind

consortium, in particular Dr. Matthias Schnabelrauch and Dr. Jana Becher who have

provided me with a variety of materials that were essential for my thesis. I would like

to thank Dr. Markus Rottmar for the work with staurosporine, Dr. Felix Theiss for

believing in my stretching chamber, Dr. Li Zhang and Famin Qiu for the collaboration in

drug delivery using their magnetic helices. I would also like to thank Nadiia Kondratiuk

for her help in the immunostaining, Dr. Eva Beurer for her help with the contact angle

measurements and Celine Gandar for her help with the Qgel experiments.

I would like to thank all the people from CERL and LBB who have helped

realizing this thesis. Thank you Ece, you have been a great support in a large part of

this thesis. Thank you Gemma, your input and biology experience was very valuable in

several aspects of the thesis. Thank you Chris for your amazing images and for the

great time in New Orleans, and the Houston conference was great with you and

Gemma. Thank you Mischa for helping me with the mechanical testing. Thanks Orane

for your advices, discussions, beers, shishas and snowboarding lessons. Thanks Marta

for hosting me in Goteborg, dinners, beers and the visit to Lebanon. Thanks Alex

Larmagnac for the discussions and going out watching the Champions League, going

out with you was always fun. Thanks Bernd for all the nice times we spent going out

and more beers. Thanks Tatiana, it was always so much fun hanging out with you.

Thanks Esther, Dario, Elsa, Deborah, Pablo, Alex Tanno, Mike, Raphael Zahn and

Grüter, Blandine, Geraldine, Pascal, Sophie, Benjamin, Nicolas, Takumi, Andreas

Binkert and Dahlin, Kaori, Ana, Peter, Anette, Gosia, Alfredo, Claudio, Max, Anne,

Martin, Bebeka, Prayanka, Victoria, Tomaso, Harald, Laszlo, Juliane, you have all been

so much fun to work with and I can definitely remember happy moments with each

and every one of you. Special thanks to my students, Philippe, Aditya and Queralt

whose input was totally essential in this thesis. I enjoyed a lot working with all of you

each in a special way. Philippe, I still remember from you “life is always good”. Addy

you were so enthusiastic hard working and fun. Queralt, you were so enthusiastic, hard

working and fun for a bit longer time than Addy. Addy and Queralt, you made the last

year of my thesis so much productive but also so much fun, I am very glad to have met

you and worked with you. I would like to thank Leena Jaatinen for being my soul mate

during a big part of my PhD. Thank you for the great times we spent together, for the

amazing Saint Feliu conference, I will never forget that week. Thank you for all the

advices, support and for being there whenever I needed. Thank you for pushing me

into cycling, I think I will do it for the rest of my life, thanks for the snowboarding

lessons and all the amazing times we had.

Abstract

Mature articular cartilage has very limited ability to self-repair after injury or

disease. In autologous chondrocyte implantation procedures (ACI), chondrocytes are

cultured on tissue culture plastic where they undergo dedifferentiation assuming a

fibroblast-like phenotype and gene expression profile. Dedifferentiated chondrocytes

re-express cartilage specific genes when cultured in materials like alginate hydrogels

which induce a round morphology. The cellular microenvironment plays a crucial role

in directing proliferation, adhesion, metabolic and catabolic activities of cells. Cells

interact with their microenvironment via membrane receptors that recognize changes

in the extracellular matrix (ECM), oxygen levels, mechanical stimuli and small

molecules. The aim of this thesis was to engineer 2D and 3D microenvironments which

will improve conditions for cartilage tissue engineering applications such as ACI. To

achieve this goal we applied the layer-by-layer technique to optimize chondrocyte

culture conditions in 2D substrates and prevent dedifferentiation. We then worked on

designing biomimetic microenvironments by incorporating cartilage-specific molecules

into 3D scaffolds. The scaffolds developed in this thesis could restore the cartilage

phenotype of dedifferentiated chondrocytes or induce chondrogenic differentiation of

stem cells while enhancing cell proliferation. Finally, we investigated the

microenvironment conditions (mechanical load, oxygen tension and stiffness) which

promote expression of superficial zone protein (SZP), a key protein in cartilage

lubrication.

With the use of the layer-by-layer technique we were able to prepare natural

ECM-based films made of type I collagen (Col1)/chondroitin sulfate (CS) or

Col1/Heparin (HN) on stretchable polydimethylsiloxane (PDMS) substrates. The

Col1/CS films were stable in media while Col1/HN films were not. Preliminary studies

showed that chondrocytes did not restore the expression of cartilage markers when

grown on these films which indicates that a 3D environment is probably more

important to maintain the cartilage phenotype than the interaction with ECM

molecules.

Given the results of the 2D system, the focus of the thesis shifted to engineered

3D microenvironments. We observed that sulfation of alginate induced cell

proliferation while maintaining the chondrogenic phenotype of encapsulated

chondrocytes. Remarkably, alginate sulfate hydrogels showed a cartilage-like, opaque

appearance after 5 weeks in culture whereas the unmodified alginate samples

remained translucent. The problems associated with limited availability of healthy

early passage chondrocytes drives the work towards using human mesenchymal stem

cells (hMSCs) for repair. We therefore studied the possibility of inducing

chondrogenesis of hMSCs using a bio-inspired hydrogel. Interactions of MSCs with the

triple helical collagen mimetic, GPC(GPP)5-GFOGER-(GPP)5GPC-NH2, and the

fibronectin adhesion peptide, RGD, were studied in degradable or non-degradable PEG

gels. GFOGER-modified degradable gels induced the highest cell proliferation and were

the most chondrogenic of the investigated conditions. Finally we studied the in-vitro

conditions which promote expression of the superficial zone marker SZP. We observed

that bovine chondrocytes in monolayer showed a drastic decrease in SZP expression,

similar in trend to the commonly reported downregulation of Col2. Chondrocytes

embedded in alginate beads for 4 days re-expressed SZP but not Col2. Cyclic

mechanical strain and normoxic conditions upregulated SZP expression whereas Col2

expression was upregulated only in alginate beads under hypoxic conditions. This work

suggests that the distribution of environmental signals in the different zones of

cartilage in-vivo is responsible for the layer-specific distribution of matrix proteins

throughout the thickness of articular cartilage.

To conclude, the cell microenvironment can be engineered to induce cell

proliferation, maintain the cartilage phenotype of chondrocytes, regulate expression of

cartilage zonal markers and promote chondrogenic differentiation of stem cells. The

results of this thesis provide insight into several crucial aspects of the

microenvironment and should lead the way to the discovery and application of novel

promising materials, such as alginate sulfate derivatives, which could transform

current treatment strategies for repair and regeneration of cartilage lesions.

Riassunto

La cartilagine articolare matura ha una capacità molto limitata di

autorinnovamento in seguito a lesioni o malattie. Nelle procedure d’impianto di

condrociti autologhi (ACI), i condrociti sono solitamente coltivati in fiasche di plastica

per colture cellulari in cui vanno incontro a de-differenziamento, assumendo un

fenotipo e un profilo di espressione genica, tipici dei fibroblasti. I condrociti de-

differenziati ri-esprimono geni specifici della cartilagine quando coltivati in materiali

come idrogel di alginato che inducono una morfologia rotondeggiante. Il

microambiente cellulare gioca un ruolo cruciale nel dirigere proliferazione, adesione, e

attività metaboliche e cataboliche delle cellule. Queste ultime interagiscono infatti con

il microambiente circostante attraverso recettori di membrana che riconoscono

cambiamenti nella matrice extracellulare (ECM), nei livelli di ossigeno, stimoli

meccanici e piccole molecole. Lo scopo di questa tesi è stato quello di ingegnerizzare

microambienti in 2D e 3D che potessero migliorare le condizioni per applicazioni

d’ingegneria tissutale della cartilagine come l’ACI. Per raggiungere tale obiettivo,

abbiamo innanzitutto utilizzato la tecnica "layer-by-layer" che ha permesso di

ottimizzare le condizioni di coltura dei condrociti in substrati 2D e al tempo stesso

prevenirne il de-differenziamento. In seguito abbiamo lavorato alla progettazione di

materiali bio-mimetici in cui molecole specifiche della cartilagine fossero incorporate

in scaffold 3D. Gli scaffold sviluppati in questa tesi potrebbero essere utili per

ristabilire il fenotipo cartilagineo dei condrociti de-differenziati o indurre il

differenziamento condrogenico di cellule staminali, aumentando al tempo stesso la

proliferazione cellulare. Infine abbiamo investigato le condizioni del microambiente

(carico meccanico, tensione di ossigeno e rigidezza) che promuovessero l’espressione

della "proteina della zona superficiale" (SZP), una proteina chiave nella lubrificazione

della cartilagine.

Con l’uso della tecnica "layer-by-layer" abbiamo preparato dei film basati sulla

matrice extracellulare naturale e contenenti collagene di tipo I (Col1)/condroitin

solfato (CS) o Col1/eparina (HN) su substrati estensibili di polidimetilsilossano (PDMS).

Tra i due tipi di film, solo quelli contenenti Col1/CS si sono rivelati stabili nel terreno di

coltura. Studi preliminari hanno dimostrato che i condrociti non erano capaci di

ripristinare l’espressione di marcatori della cartilagine se cresciuti su questi film. Tal

evidenza sperimentale indica che un ambiente tridimensionale è probabilmente più

importante dell’interazione con le molecole della matrice extracellulare, al fine di

mantenere il fenotipo cartilagineo.

Dati i risultati del sistema bidimensionale, abbiamo rivolto l’attenzione verso

l’ingegnerizzazione di microambienti in 3D. Abbiamo osservato che la solfatazione di

alginato promuoveva proliferazione cellulare, mantenendo al tempo stesso il fenotipo

condrogenico di condrociti incapsulati. Sorprendentemente gli idrogel di alginato

solfato assumevano un aspetto opaco simile alla cartilagine dopo 5 settimane in

coltura mentre i campioni di alginato non modificato rimanevano traslucidi. I problemi

associati alla disponibilità limitata di condrociti sani a passaggio precoce indirizza il

lavoro verso l’utilizzo di cellule staminali mesenchimali umane (hMSCs) per la

rigenerazione. Di conseguenza abbiamo studiato la possibilità di indurre la

condrogenesi di cellule hMSCs usando un idrogel bio-ispirato. In particolare, abbiamo

investigato le interazioni di hMSCs con un peptide che mima il collagene a tripla elica,

GPC(GPP)5-GFOGER-(GPP)5GPC-NH2, e con un peptide di adesione tipico della

fibronectina, RGD, all’interno di gel di polietilenglicole (PEG), degradabili e non. Tra le

condizioni prese in esame, i gel contenenti il peptide GFOGER e degradabili si sono

rilevati i migliori sia per l’induzione di proliferazione cellulare che per la condrogenesi.

Infine, abbiamo condotto uno studio per identificare le condizioni in vitro che

promuovessero l’espressione di SZP, il marcatore della zona superficiale della

cartilagine. A tale proposito, abbiamo osservato che i condrociti bovini cresciuti come

mono-strato mostravano una notevole diminuzione nell’espressione di SZP, con una

tendenza simile alla riduzione di Col2 comunemente riportata. Condrociti incorporati

in perline di alginato per 4 giorni ri-esprimevano SZP ma non Col2. Una sollecitazione

meccanica ciclica e condizioni di normossia inducevano l’espressione di SZP mentre

l’espressione di Col2 risultava aumentata solo in perline di alginato mantenute in

ipossia. Questo lavoro suggerisce che la distribuzione strato-specifica di proteine della

matrice attraverso lo spessore della cartilagine articolare.

In conclusione, il microambiente cellulare può essere ingegnerizzato per

indurre proliferazione cellulare, mantenere il fenotipo cartilagineo, regolare

l’espressione di marcatori specifici delle diverse zone della cartilagine e promuovere il

differenziamento condrogenico delle cellule staminali. I risultati di questa tesi hanno

permesso la comprensione di diversi aspetti cruciali del microambiente e aprono la

strada alla scoperta e l’applicazionedi nuovi materiali ideali per l’ingegneria tissutale

della cartilagine come l’alginato solfato.

Contents

1 Tissue Engineering & Articular Cartilage ................................................... 1

1.1 The structure of articular cartilage ........................................................ 2

1.2 Cartilage disease, injury and management ............................................ 4

1.2.1 Non-surgical approaches (focus on chondroitin sulfate) .................. 5

1.2.2 Surgical approaches .......................................................................... 6

1.3 Autologous chondrocyte implantation (ACI) ......................................... 7

1.3.1 History and development of ACI ....................................................... 7

1.3.2 The limitations of ACI ....................................................................... 8

1.4 3D culture systems ................................................................................ 9

1.4.1 Preventing dedifferentiation in monolayer cultures....................... 10

1.4.2 Biomimetic systems ........................................................................ 11

1.4.3 Mechanical stimulation .................................................................. 12

2 Scope of the thesis ................................................................................ 15

3 Materials and methods .......................................................................... 19

3.1 Materials ............................................................................................. 19

3.2 Instruments ......................................................................................... 21

3.3 Protocols .............................................................................................. 24

4 Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates ................................................................................ 35

4.1 Current monolayer culturing techniques ............................................. 35

4.2 Build-up of Col1/CS and Col1/HN films on PDMS ................................ 37

4.3 Effect of substrate on film build-up ..................................................... 41

4.4 Film topography .................................................................................. 42

4.5 Film thickness and stability .................................................................. 44

4.6 Assessment of cell adhesion and integrin mediated spreading ........... 46

4.7 Chapter summary ................................................................................ 51

5 Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers ... 53

5.1 Improving chondrogenic performance of chondrocytes in 3D may be

achieved using biomimetic materials. ................................................... 53

5.2 Preparation and characterization of sulfated alginate ......................... 55

5.3 Morphology of chondrocytes encapsulated in sulfated alginate (DSs =

0.8) ........................................................................................................ 57

5.4 Assessment of cell proliferation within the hydrogels ......................... 59

5.5 Decoupling stiffness from cell spreading ............................................. 60

5.6 RhoA and integrin signalling of chondrocytes in alginate sulfate ........ 62

5.7 Cyclin D1 expression is upregulated in alginate sulfate samples ......... 64

5.8 Expression of cartilage markers for dedifferentiated chondrocytes within

alginate sulfate hydrogels ...................................................................... 65

5.9 Immunohistological staining and gross appearance of alginate sulfate

hydrogels ............................................................................................... 66

5.10 Chapter summary ............................................................................ 69

6 GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells ......... 71

6.1 Chondrogenic differentiation and the microenvironment ................... 71

6.2 Hydrogel modification did not affect mechanical properties .............. 73

6.3 Cell viability ......................................................................................... 74

6.4 Cell morphology .................................................................................. 75

6.5 Cell proliferation was highest in RGD and GFOGER degradable hydrogels

............................................................................................................. 77

Contents

6.6 Gene expression .................................................................................. 78

6.7 GAG production in peptide-modified gels ........................................... 79

6.8 Histology and immunostaining ............................................................ 80

6.9 Chapter summary ................................................................................ 84

7 Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression ........................................................................ 85

7.1 Chondrocyte dedifferentiation and superficial zone protein ............... 85

7.2 Both SZP and Col2 undergo dedifferentiation during serial passaging 87

7.3 Redifferentiation of serially passaged primary chondrocytes .............. 88

7.4 Design and evaluation of a 3D tension and compression chamber

compatible with the strex strain machine ............................................. 90

7.5 Application of cyclic mechanical strain to 2D and 3D constructs ......... 92

7.6 Effect of mechanical strain on SZP expression ..................................... 94

7.7 Effect of oxygen tension on expression of SZP .................................... 95

7.8 Cell morphology and the expression of SZP ......................................... 98

7.9 Chapter summary .............................................................................. 105

8 Conclusions and outlook ...................................................................... 107

9 References .......................................................................................... 109

Curriculum Vitae .................................................................................................. 125

List of Figures and Tables

Figure 1. 1 | Articular cartilage structure and composition ......................................................................... 4

Figure 1. 2. | ACI procedure as described in Brittberg et al. 1994 ............................................................... 8

Figure 1. 3. | Schematic representation of the build-up of ECM films and application of load on seeded

cells. ............................................................................................................................................................ 11

Table 3. 1. | Genes used in qRT-PCR analysis ............................................................................................ 23

Figure 4. 1. | Buid-up of Col1/CS multilayer ............................................................................................... 38

Figure 4. 2. | Buid-up of Col1/HN multilayer .............................................................................................. 39

Figure 4. 3. | Buid-up exponential (Col1/HN) vs. linear (Col1/CS) ............................................................. 40

Figure 4. 4. | Build-up of Col1/CS on gold .................................................................................................. 42

Figure 4. 5. | Morphology of Col1/CS and Col1/HN films ........................................................................... 43

Figure 4. 6. | Films’ thickness and stability. ............................................................................................... 45

Figure 4. 7. | Morphology of bovine chondrocytes seeded ECM films ...................................................... 48

Figure 4. 8. | Integrin mediated cell spreading .......................................................................................... 50

Figure 5. 1. | Reaction scheme for the synthesis of alginate sulfate ......................................................... 56

Figure 5. 2. | Morphology of cells encapsulated in alginate and alginate sulfate ...................................... 58

Figure 5. 3 | Cell proliferation in gels of various alginate sulfate (AlgSulf) contents ................................. 60

Figure 5. 4 | Mechanical properties of alginate sulfate (AlgSulf) hydrogels and effect on cell proliferation

Figure 5. 5. | Integrin mediated cell spreading and proliferation within alginate sulfate (AlgSulf)

hydrogels and RhoA activity. ...................................................................................................................... 63

Figure 5. 6. | Cyclin D1 gene expression. ................................................................................................... 65

Figure 5. 7. | Expression of relevant cartilage markers. ............................................................................. 66

Figure 5. 8. | Gross appearance and immunostaining of matrix molecules. .............................................. 67

Figure 6. 1. | Compressive moduli of the hydrogels.. ................................................................................ 74

Figure 6. 2. | Cell viability in the hydrogels.. .............................................................................................. 75

Figure 6. 3. | hMSCs cultured for 21 days in chondrogenic media in modified PEG gels prior to fixation

and staining for nuclei (DAPI, blue) and actin filaments (phalloidin, red).. ................................................ 76

Figure 6. 4. | Assessment of cell proliferaiton............................................................................................ 77

Figure 6. 5. | Gene expression of relevant cartilage markers .................................................................... 79

Figure 6. 6. | GAG/DNA quantification in the various conditions .............................................................. 80

Figure 6. 7. | Alcian blue staining of hMSC cultured in pellets and in PEG gels of different compositions.

hMSC cultured in pellets with control medium .......................................................................................... 81

Figure 6. 8. | Type II collagen immunostaining of hMSC cultured in pellets and in PEG gels of different

compositions .............................................................................................................................................. 82

Figure 7. 1. | Effect of chondrocyte dedifferentiation on gene expression of SZP and type II collagen. ... 87

Figure 7. 2. | Redifferentiation of serially passaged chondrocytes ............................................................ 89

Figure 7. 3. | Design for the 3D molds A and the final 3D chamber B. ....................................................... 91

List of Figures

Table 7. 1. | Actual strains measured in the alginate gel and the corresponding strain setting in the

STREX device. .............................................................................................................................................. 92

Figure 7. 4. | Design for application of 3D cyclic strain. ............................................................................. 93

Figure 7. 5. | The effect of mechanical strain on SZP mRNA levels was measured by qRT-PCR ................ 95

Figure 7. 6. | The effect of oxygen tension on SZP and Col2 mRNA levels was quantified by qRT-PCR. .... 97

Figure 7. 7. |Morphology of chondrocytes cultured in 2D and 3D substrates. .......................................... 99

Figure 7. 8. | Effect of cell morphology and dimensionality on SZP gene expression .............................. 100

Figure 7. 9. | Effect of RGD sequences on morphology and SZP gene expression. .................................. 102

Figure 7. 10. | Morphology and gene expression of SZP for chondrocytes cultured in 3D MMP sensitive

PEG hydrogels with and without RGD peptides ....................................................................................... 104

Chapter 1

1

1 Tissue Engineering & Articular Cartilage

Tissue engineering is an interdisciplinary field that utilizes cells, biomaterials,

biochemical (e.g. growth factors) and physical (e.g. mechanical loading) signals, as well

as their combinations to generate tissue-like structures [1]. The goal of tissue

engineering is to provide biological substitutes that can maintain, restore or improve

the function of damaged tissues [2]. Skin grafts were the first tissue-based therapies

and introduction of techniques to preserve cells and tissues enabled allograft skin

banking [3-5]. However, limited donor availability and rejection of the grafts by the

immune system drove the concept of in-vitro grown tissues. Although the first tissue

engineered skin products were introduced in the late 1970s and early 1980s giving rise

to modern tissue engineering, the term “tissue engineering” was only coined in 1987

[6-9]. The success of engineering skin grafts boosted interest in applying similar

concepts to other tissues and organs [10]. However, the relatively simple structure, the

limited vascular demands of skin and the ease of growing keratinocytes in-vitro are not

common to most tissues. The dream of regenerating tissues in-vitro faced major

hurdles associated with the engineering of complex 3-dimentional vascularized

multicellular tissues.

Cartilage, an avascular unicellular tissue, was the next target for tissue

engineering which generated early success. The high prevalence of osteoarthritic

disease as well as cartilage injuries incurred by athletes create a large market for

cartilage tissue engineering solutions [1]. Although engineering cartilage at first might

seem to be an easy task, the tissue’s stratified structure, unique tensile, compressive

and lubricative properties and compositional and ultrastructural complexity makes it

difficult if not impossible with current technologies to fully replicate the tissue [11-13].

Current available tissue engineering strategies for cartilage repair do not recapitulate

the cartilage microenvironment which provides a multitude of cues necessary for

cartilage homeostasis [14].

Introduction

2

1.1 The structure of articular cartilage

Articular cartilage is the tissue that covers the surface of articulating joints

providing an almost frictionless interaction and absorbing shocks from daily physical

activities. Articular cartilage is composed of extracellular matrix containing a single

type of sparsely distributed highly specialized cells (chondrocytes) and lacking

vasculature, lymphatic vessels and nerves (Figure 1.1) [13]. Chondrocytes vary in

morphology and metabolic activity depending on their location within the thickness of

the tissue. The distribution and organization of the extracellular matrix also varies

allowing the identification of four different zones from the surface of articular cartilage

to the subchondral bone: the superficial zone, transitional zone, deep zone and

calcified cartilage zone. The superficial zone is the thinnest zone comprising an

acellular thin sheet of collagen fibrils. The ellipsoid-shaped chondrocytes below this

sheet are arranged with their major axis parallel to the cartilage surface. The thin sheet

of collagen fibrils provides this layer with higher tensile strength compared to deeper

layers allowing it to resist high shear forces on the surface of the tissue. The dense

sheet of collagen also acts as a barrier preventing the leakage of cartilage molecules

out of and/or the entrance of antibodies and cytokines into the cartilage. Disruption of

the collagen layer is one of the first observed physical alterations in experimentally

induced degeneration of articular cartilage [15]. Chondrocytes of this layer synthesize

different amounts and types of matrix proteins compared to deeper layers,

particularly, they synthesize superficial zone protein (SZP) which is highly homologous

to lubricin [16], a major contributor of cartilage lubrication [17]. The transitional zone

has several times the volume of the superficial zone with a matrix composition and cell

morphology intermediate between the superficial and deep zones. Cells of the

transitional zone have a spheroidal morphology and synthesize thicker collagen fibrils

and proteoglycans at a higher concentration compared to cells of the superficial zone.

The deep zone has the highest concentration of proteoglycans and thickest collagen

fibrils. Chondrocytes in the middle layer assume a spheroidal morphology and are

often aligned in columns perpendicular to the surface of cartilage. The calcified

Chapter 1

3

cartilage zone separates the deep zone and the subchondral bone and provides a

physical barrier between cartilage and bone.

The cellular microenvironment plays a pivotal role in the function of cells [18].

In chondrocytes several factors are known to influence chondrocyte migration,

proliferation, matrix synthesis and degradation. Moreover chondrocytes of the

different layers in cartilage have a different microenvironment as shown in Figure 1.1.

Chondrocytes are surrounded by an extracellular matrix which is predominantly

composed of type II collagen and proteoglycans [13]. Chondrocytes possess cell

membrane receptors called integrins which mediate interactions with matrix proteins.

Integrin α10β1 recognizes the GFOGER sequences in collagen [19] while the receptor

CD44 mediates cell interaction with hyaluronic acid and chondroitin sulfate [20]. The

oxygen levels vary through the cartilage thickness from 7-10% on the surface down to

0.1% near the subchondral bone [21]. Mechanical stimulation also varies in type and

magnitude throughout cartilage [22]. In stem cells, the microenvironment changes

considerably throughout the stages of chondrogenesis. Initially, MSCs produce

fibronectin followed closely by an increase in type I collagen expression which

promotes their motility leading to condensation, a pre-requisite step to MSC

chondrogenesis. Differentiation into chondrocytes follows with cells decreasing

fibronectin expression and replacing type I collagen with up-regulation of type II

collagen and aggrecan [23].

Introduction

4

Figure 1. 1 | Articular cartilage structure and composition. Electronic micrographs of cells from the

various cartilage zones (left panel), full thickness articular cartilage (middle panel) and collagen and

matrix composition and structure (right panel). Reproduced with permission from [24, 25].

1.2 Cartilage disease, injury and management

Osteoarthritis is a common joint disease affecting over 200 million people

worldwide. In the United States alone, more than 40 million people are affected by

arthritis (approximately 15% of the total population) and the number is estimated to

grow to 60 million by 2020 [26]. Osteoarthritis is a low mortality disease which causes

pain and can lead to disability in advanced stages. Osteoarthritis is primarily a disease

of the aging population [27], but is occurring more commonly in the younger

population especially those engaged in high-impact sport activities. Sports-related

increase in the risk of osteoarthritis in soccer and ice hockey athletes has been

reported to be related to knee injuries and not to physical activity during sports

engagement [28].

Chapter 1

5

Articular cartilage has very limited ability to self-repair due to its avascular

nature, limited proliferative ability of mature chondrocytes and limited access to

progenitive repair cells. Furthermore, the ability of healthy chondrocytes to migrate to

injured sites is restricted by the surrounding extracellular matrix (ECM) [29]. Currently

no treatment is available to fully regenerate damaged cartilage. Treatments for

cartilage can be categorized into non-surgical and surgical treatments, with the choice

being dependent on the severity of the injury. Tissue engineering holds strong promise

in providing functional replacement for cartilage lesions. However, improvement of

the three elements of tissue engineering namely cells, biomaterials and culture

conditions are necessary to achieve ideal reconstruction of cartilage defects.

1.2.1 Non-surgical approaches (focus on chondroitin sulfate)

The pharmacological agents used for osteoarthritis treatment can be

categorized into three main groups: 1) non-steroidal anti-inflammatory agents and

analgesics (NSAIDs); 2) symptomatic slow-acting drugs for OA (SYSADOA); 3)

chondroprotective or truly disease-modifying agents known as disease-modifying OA

drugs (DMOADs) [30]. Paracetamol is a safe recommended oral analgesic but does not

always provide sufficient relief from severe pain [31, 32]. NSAIDs are commonly used

when significant inflammation occurs. However, several serious side-effects are

associated with long-term use of NSAIDs including gastric erosion, kidney injury, ulcer

disease, acute renal complications among others [33-36]. NSAIDs are useful for pain

relief but they do not have direct influence on the progression of osteoarthritis,

therefore safer disease modifying agents are required. Some ECM molecules such as

chondroitin sulfate (CS), hyaluronan and glucosamine are believed to have

chondroprotective properties [27, 37-42]. In cartilage, chondroitin sulfate is attached

to the core protein of proteoglycans which contributes to the tissue’s compressive

properties by entrapping water. Chondroitin sulfate has been shown to stimulate ECM

production, supress inflammation and inhibit cartilage degeneration [40, 43-46].

Chondroitin sulfate and glucosamine can be taken orally unlike hyaluronan which

Introduction

6

requires injection. The safety and tolerability of chondroitin sulfate were confirmed in

previous reports and it is currently available in several products such as Condrosulf®

and Matrix® [47, 48].

1.2.2 Surgical approaches

For patients who are diagnosed with full-thickness cartilage injuries and where

non-surgical pharmacological treatments fail in restoring function and pain relief, a

surgical procedure may be required. The current surgical approaches include lavage

and debridement, microfracture (marrow stimulation), osteochondral grafting and cell

based therapies (autologous chondrocyte implantation). Lavage involves washing away

small molecules, debris and inflammatory mediators and debridement is the removal

of larger cartilage, meniscal fragments and osteophytes. Lavage and debridement can

provide temporary relief but does not directly address the cartilage injury.

Microfracture or marrow stimulation is the creation of holes in the subchondral bone

which causes bleeding and the formation of a fibrin clot. Bone marrow stem cells can

then migrate to the clot and form a fibrocartilage repair tissue. The inferior mechanical

and biochemical properties of the resulting repair tissue are major disadvantages of

the method. Osteochondral grafting is usually done by taking autologous

osteochondral plugs from low-weight-bearing regions of cartilage and repairing the

chondral injuries with a mosaic of harvested plugs. The main drawbacks of this method

are donor site morbidity and low integration of the grafts with the surrounding tissue.

The gap between the graft and the surrounding cartilage is normally filled with

fibrocartilage which leads to strong alterations in load distribution. Finally autologous

chondrocytes implantation (ACI) is the process of using autologous chondrocytes

which are then expanded in the laboratory to fill the cartilage defect [49, 50]. ACI is a

promising technique for the treatment of cartilage and is described in more details in

the following sections.

Chapter 1

7

1.3 Autologous chondrocyte implantation (ACI)

Autologous chondrocyte implantation is the first approach that applies the

principle of tissue engineering to treat cartilage defects. In this thesis, we focus on

providing the proper microenvironment to enhance current culturing techniques with

the aim of addressing the areas where ACI currently falls short.

1.3.1 History and development of ACI

Autologous chondrocyte transplantation was developed in the late 1980s by

Peterson and co-workers and described in humans by Brittberg et al. in 1994 [51-53].

In this procedure a biopsy of healthy cartilage tissue (approx. 300 milligrams) is taken

from a low-weight-bearing region of the knee. The tissue is enzymatically digested in

the laboratory to isolate the cells which are then are grown in-vitro to reach sufficient

cell numbers. In the original embodiment, a periosteal flap large enough to cover the

defect is harvested from the proximal medial tibia and sutured to the surface of the

cartilage leaving a small gap for cell injection. The spaces between the sutures are

filled with fibrin glue creating a water-tight seal that prevents cell leakage and cells are

injected to fill the defect site (Figure 1.2). Although ACI presented a major

advancement in cartilage therapy, periosteal hypertrophy encouraged the research for

improved repair strategies [54, 55]. In order to address the above problems, a

membrane based on type I/III porcine collagen has been developed and is used to

replace the periosteal flap in ACI procedures [56-58]. The membrane commercialized

as (Chondrogide) has a porous surface on the side that faces the defect which allows

cell attachment and a smooth compact surface that prevents cell leakage.

Research is currently focused on the use of cells from more promising sources,

chondroinductive/conductive biomaterials, gene therapy, simplifying the surgical

procedures, mechanical stimulation and preconditioning of cells [14]. The work in this

thesis contributes to the research on the use of stem cells, biomaterials with

Introduction

8

biochemical cues and cell affinity, growth factors and in-vitro preconditioning with

mechanical stimulation.

Figure 1. 2. | ACI procedure as described in Brittberg et al. 1994. Reproduced with permission from

[52], Copyright Massachusetts Medical Society

1.3.2 The limitations of ACI

ACI relies on the growth of cells in the lab in monolayers which causes the loss

of the cartilage phenotype or what is commonly known as dedifferentiation [59].

Chondrocyte dedifferentiation is associated with morphological and gene expression

changes where cells behave more as fibroblasts [60, 61]. Dedifferentiated

chondrocytes produce type I collagen rather type II collagen and thus cells implanted

in ACI procedures often produce fibrous tissue and not hyaline cartilage [62]. The

current research focuses on developing a variety of methods to either prevent

dedifferentiation during serial expansion or to restore the cartilage phenotype of cells

prior or after the delivery to the defect site [63, 64]. Another disadvantage of ACI is

Chapter 1

9

that it relies on the use of articular chondrocytes which are limited in supply. Much of

the research now focuses on using different cell sources which can provide the same

repair properties as chondrocytes [65, 66]. Stem cells are a promising cell source and

various studies have reported the induction of chondrogenesis using TGFβ and

dexamethasone [67, 68]. Materials which are able to induce proliferation while

maintaining the cartilage phenotype of chondrocytes or inducing chondrogenesis of

stem cells would be highly desirable.

1.4 3D culture systems

Chondrocytes cultured on tissue culture plastic undergo dedifferentiation

where they assume a fibroblast like phenotype and alter their gene expression

producing proteins typical of fibroblasts [62]. Chondrocytes gene expression has been

shown to be tightly linked to its morphology where a round cell is typical of a

differentiated chondrocyte while a flattened morphology triggers chondrocyte

dedifferentiation [60, 69-71]. Culturing chondrocytes in 3D gels such as agarose [60] or

alginate [72] preserves the cartilage phenotype. Nevertheless, chondrocytes

proliferation in alginate and agarose is very limited thus it might not be the most

suitable method to culture chondrocytes when high cell yields are required. Synthetic

gels such as poly(ethylene glycol) (PEG) are attractive for 3D culturing of cells, as they

can be functionalized with specific adhesion sequences, growth factor binding sites

and protease-sensitive motifs [73-76]. Providing adhesion sequences within these

hydrogels would increase cell proliferation and may potentially maintain the cartilage

phenotype. Moreover, application of mechanical stimulation and low oxygen tension

can also help restoring the cartilage phenotype. Attempts to maintain the cartilage

phenotype in 2D cultures by cell spreading restriction or drugs have also been pursued

and are further discussed below.

Introduction

10

1.4.1 Preventing dedifferentiation in monolayer cultures

Experiments that were previously performed to achieve a cartilage phenotype

in 2D focus mainly on preventing cell spreading such as inhibiting RhoA/ROCK

signalling using Y-27632. Consequently, preventing actin polymerization by adding

cytochalasin D or stabilizing existing actin filaments with jasplakinolide helped recover

the cartilage phenotype in dedifferentiated cells [70, 71]. The addition of concanavalin

A to chondrocyte cultures also caused cell rounding and increased proteoglycan

synthesis [77]. In these experiments it was clearly demonstrated that a round

chondrocyte phenotype correlates with maintaining the cartilage phenotype. In a

collaboration with Rottmar et al. (submitted to Exp. Cell. Biol.), it was concluded that

the cartilage phenotype can be maintained in spread cells with a disrupted

cytoskeleton. A more high-tech approach involves the use of micro-patterned

substrates where the cells can be constrained to a round morphology [63, 78]. While

these approaches were shown to be beneficial, proliferation is constrained by the

patterns and thus the system does not serve the requirement of high cell numbers for

ACI. On the other hand the use of drugs might be associated with toxicity and would

require costly clinical trials before it is accepted for use in patients. An approach that

may help address these issues is to engineer substrates that allow cell proliferation

while maintaining the cartilage phenotype using safe molecules and possible external

stimuli.

The layer-by-layer (LbL) technique has been widely used to produce nanofilms

for biomedical applications [79, 80]. The LbL approach allows the build-up of thin films

simply by the alternating deposition of polyanions and polycations [81]. Since the

introduction of the LbL technique, a large body of research has been produced

describing the build-up of films of different compositions [82-89]. The majority of

these films rely on synthetic polymers with relatively high charge densities. Films such

as PSS/PAH or PLL/HA promote cell adhesion and are not cytotoxic [83, 84].

Nevertheless, polymers that diffuse out of the layers and could be taken up through

the cell membrane raise strong concerns of cytotoxicity [87]. The growth of cells has

Chapter 1

11

been shown to be dependent on the stiffness, chemistry and charge of the films [82-

86].

The preparation of films from natural polymers is limited to a few combinations

which are rarely stable [82, 88, 89]. The build-up can be achieved using ECM molecules

such as collagen which is positively charged at a pH lower than 5.5 and CS or any other

negatively charged molecule [90, 91]. Chondrocytes cultured in contact with ECM

molecules and concomitantly exposed to mechanical load are expected to have a high

chondrogenic activity optimal for cartilage tissue engineering applications (Figure 1.3).

Such systems resemble the natural environment in which chondrocytes exist and thus

are expected to be more physiologic.

Figure 1. 3. | Schematic representation of the build-up of ECM films and application of load on seeded

cells.

1.4.2 Biomimetic systems

Biomimetic systems are systems designed based on biological structures [92].

Biomimetic approaches were shown to be beneficial in cartilage tissue engineering

applications. Mimicking the natural chondrocyte round morphology in-vitro by

encapsulation in alginate and agarose leads to restoration of the cartilage phenotype

[60, 72]. Molecules that are naturally present in cartilage such as chondroitin sulfate

and hyaluronic acid were shown to have chondroprotective properties and are used

for cartilage treatment [27, 49]. Scaffolds containing these molecules or certain

Introduction

12

moieties of these molecules such as sulfate groups might have beneficial effects on

chondrocytes. Mechanical stimulation at physiological levels is known to stimulate

expression and synthesis of cartilage markers [93, 94]. Chondrocytes cultured in

hypoxic conditions were shown to synthesize more type II collagen and aggrecan [21,

95]. The fibronectin cell binding domain RGD has been shown to be involved in cell

adhesion, proliferation and differentiation of stem cells [96, 97]. However, the spread

morphology induced by this sequence may not be optimal for chondrocytes, which

form primary interactions with collagen 2 and hyaluronic acid in native cartilage [96].

Recently the GFOGER peptide was identified as the collagen sequence responsible for

chondrocyte adhesion [19, 98]. Incorporation of the GFOGER sequence in hydrogels

would provide a better cartilage mimic than the commonly used RGD sequences. The

GFOGER peptide might improve the chondrogenic performance of 3D encapsulated

chondrocytes and chondrogenesis of stem cells.

1.4.3 Mechanical stimulation

Mechanical stimulation varies throughout cartilage where cells of the

superficial layer are subject to shear, stress and compressive forces, cells of the

transitional layer and deep layer are subject to compressive forces and hydrostatic

pressure and cells of the calcified layer are subject mainly to hydrostatic pressure [22].

Mechanical loading experiments have been widely performed to obtain information

about mechanisms of cell mechanotransduction and to enhance proliferation, gene

expression, and protein synthesis in chondrocytes [94]. Mechanical signal sensing by

cells may occur through changes in the cell membrane sensed through integrins and

ion channels and their downstream signalling pathways [94, 99]. It has been shown

that in chondrocytes the application of compression at physiological levels increases

matrix protein and GAG synthesis [100]. Excessive loading, on the other hand, leads to

the production of metalloproteinases (MMPs) and aggrecanases (ADAMTS-5) that

degrade the ECM proteins [101]. Systems have been developed for the application of

tensile load, compressive load, hydrostatic pressure, shear and perfusion [102-108].

Chapter 1

13

Notably, stem cell fate can be steered through the application of phenotypic loading.

To illustrate: application of tensile loads help steer stem cells towards ligament [103],

tendon [104], or bone tissue depending on the tensile load parameters [105, 109]

while shear loads can help stem cells differentiate towards cardiac muscle [110] or

endothelial cells [107]. Finally hydrostatic pressure or compression can lead to

chondrogenic differentiation [94, 111-113]. Mechanical stimulation of chondrocytes

modulates the expression of chondrocyte marker genes such as type II collagen,

aggrecan and COMP and the synthesis of glycosaminoglycans (GAGs) [93, 106, 114,

115]. Expression of the superficial zone marker, superficial zone protein is influenced

by shear, stress and surface motion [116-118]. The mechanisms through which

chondrocytes perceive load and react to their surrounding environment are not fully

elucidated but may comprise stretch activated ion-channels and integrins [119, 120].

Understanding these mechanisms will provide basis for developing new tools in the

area of cartilage tissue engineering and possibly discovering new useful molecules in

the treatment of osteoarthritis and cartilage injuries.

Introduction

14

Chapter 2

15

2 Scope of the thesis

Cartilage disease and injury affects over 200 million people worldwide.

Osteoarthritis of cartilage is more prevalent in the aging population causing an

immense socioeconomical burden on societies. Cartilage does not heal after injury or

disease and thus requires medical intervention. Currently no disease modifying drugs

are clinically available and the surgical procedures are not yet capable of fully

recovering damaged articular cartilage. Cell based techniques such as ACI are

promising however they have limitations caused mainly by the dedifferentiation of the

transplanted cells. The use of biomimetic materials to engineer 2D and 3D substrates

for cartilage tissue engineering would potentially lead to a better outcome for ACI-like

procedures. This thesis focuses on providing the right microenvironment in

monolayers and 3D scaffolds for cartilage tissue engineering applications.

In chapter 4 we used the LbL approach to improve cell culturing conditions and

prevent dedifferentiation of in-vitro cultured chondrocytes. The layer-by-layer (LbL)

technique was used before to produce nanofilms made of ECM macromolecules.

However, most of these studies resulted in either non-stable or cell repellent films. The

build-up was also never characterized on polydimethylsiloxane (PDMS) which is a

common elastic biomaterial used in a wide range of biomedical applications. In

particular to our interest, PDMS can be used in combination with the produced natural

films to apply mechanical strain on adherent cells. We assessed the topography, build-

up, thickness and stability of type 1 collagen (Col1)/CS or Col1/heparin (HN) on PDMS

substrates using quartz crystal microbalance with dissipation (QCM-D) and atomic

force microscopy (AFM). Integrin-mediated cell adhesion was assessed by studying the

cytoskeletal organization of mammalian primary chondrocytes seeded on different end

layers and number of layers. This work provided important information on the build-up

of films made from cartilage relevant molecules on a stretchable substrate. However,

Scope

16

the chondrocytes exhibited a fibroblastic morphology on these films which lead us to

focus on 3D scaffolds in the subsequent chapters of this thesis.

In chapter 5 we focused on developing a biomimetic 3D scaffold for cartilage

tissue engineering. Inspired by the presence of sulfate groups in cartilage, we

hypothesized that sulfation of alginate which is known to restore the cartilage

phenotype would result in a more chondrogenic material. Hydroxyl groups of alginate

were converted to sulfates by incubation with sulfur trioxide-pyridine complex

(SO3/pyridine), yielding a sulfated material crosslinkable with calcium chloride.

Dedifferentiated passage 3 bovine chondrocytes were encapsulated in alginate and

alginate sulfate hydrogels for up to 35 days. Cell morphology, proliferation, expression

and synthesis of type II collagen, type I collagen and aggrecan were assessed by

quantitative real time PCR (qRT-PCR) and immunohistochemistry. The alginate sulfate

hydrogel provided a 3D microenvironment which promoted both chondrocyte

proliferation and maintenance of the chondrogenic phenotype. The results of this

study were remarkably striking especially concerning the unexpected increase in cell

proliferation emphasizing how small microenvironmental changes can completely alter

the response of cells encapsulated in a 3D scaffold.

Much of today’s research focuses on exploring new cell sources such as stem

cells for cartilage tissue engineering applications. We therefore studied in chapter 6

the possibility of inducing chondrogenesis of human mesenchymal stem cells (hMSCs)

by controlling microenvironmental cues involved in adhesion and matrix degradation

in stem cells. Interactions of human mesenchymal stem cells (hMSCs) with the triple

helical collagen mimetic, GPC(GPP)5-GFOGER-(GPP)5GPC-NH2, and the fibronectin

adhesion peptide, RGD, were studied in MMP-sensitive (degradable) or MMP-non-

sensitive (non-degradable) PEG gels formed by Michael addition chemistry. Viability,

proliferation, cytoskeletal morphology, and chondrogenic differentiation of

encapsulated hMSCs were evaluated. The type of adhesion sequence and whether it

was presented together with MMP-sensitive motifs or not had a profound influence on

the response of encapsulated hMSCs. The GFOGER peptide enhanced proliferation in

Chapter 2

17

degradable PEG gels and provided a better chondrogenic microenvironment compared

to the RGD peptide or gels with no adhesion peptide.

Finally, the microenvironmental conditions which promote expression of

cartilage zonal markers such as SZP in the superficial zone and type II collagen in the

lower cartilage zones were studied in chapter 7. Chondrocytes from 6 months old

calves were expanded in monolayer culture and the expression of SZP in alginate bead

and monolayer culture was quantified with qRT-PCR and immunostaining. The effect of

cyclic tensile strain and oxygen tension on expression of SZP and type II collagen in 2D

and 3D cultures was quantified. Bovine chondrocytes in monolayer showed a drastic

decrease in SZP expression which could be fully restored in alginate beads after 4 days

of culture. This finding indicates that although cells of the superficial zone have a more

discoidal shape than cells of the deeper layers they still require a 3D environment to

maintain their phenotype. Cyclic mechanical strain and normoxic conditions improved

SZP expression whereas Col2 was upregulated only in alginate beads under hypoxic

conditions. The results of this chapter indicate that engineering stratified articular

cartilage will require gradients both of oxygen and mechanical signals.

The thesis concludes with a summary of findings and an outlook to challenges

and future possibilities.

Scope

18

Chapter 3

19

3 Materials and methods

In this chapter, descriptions of the materials, instruments and protocols used in

this thesis are given.

3.1 Materials

Cell culture and chondrocyte isolation solutions: Phosphate buffer saline (PBS),

fetal bovine serum (FBS), cell culture media (DMEM-Glutamax), antibiotic-antimycotic

(Anti-Anti) and trypsin/EDTA were from Invitrogen AG, Basel, Switzerland. ITS+ Premix,

non-tissue culture treated 24-well plates and 40 μm cell strainers were from Becton

Dickson AG Allschwill, Switzerland. Formaldehyde, L-proline, dexamethasone, tris(2-

carboxyethyl)phosphine hydrochloride solution (TCEP), acetic acid, Spectra-Por Float-

A-Lyzer G2 black 3.5-5 kDa, hexane, cycloheximide, sodium dodecyl chloride (SDS),

sodium chloride (NaCl), pronase E from Streptomyces Griseus, bovine serum albumin

(BSA), cycloheximide, collagenase type II from Clostridium Histolyticu,

ethylenediaminetetraacetic acid (EDTA), sodium phosphate, 1,9-dimethyl-methylene

blue (DMMB), sodium formate, formic acid, ethanol, Papain from papaya latex, L-

cysteine and chondroitin 4-sulfate sodium salt from bovine trachea were purchased

from Sigma Aldrich Chemie GmbH, Buchs, Switzerland. L-Ascorbic acid phosphate

magnesium salt was obtained from Wako (IG instrumenten-Gesellschaft AG, Zurich).

Human fibroblast growth factor-2 (FGF-2) and human transforming growth factor beta

3 (TGF-β3) were from Peprotech, Rocky Hill, NJ, USA.

Materials and methods

20

Materials used in preparation of extracellular matrix multilayers: Bovine type

1 collagen (Col1) was from Becton Dickson AG Allschwill, Switzerland. Chondroitin 4-

sulfate from bovine trachea (CS) was from Fluka, Switzerland. Heparin sodium salt

(HN), hydrochloric acid (HCl), 4-(2-hydroxyethyl)piperazine-1-ethanesulfonic acid

(HEPES), N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC) and N-

hydroxysulfosuccinimide sodium salt (NHS) were purchased from Sigma Aldrich

Chemie GmbH, Buchs, Switzerland. 150mM NaCl solutions used for preparation of

polyelectrolytes had a pH of ~6 and 150mM with 1mM HCl had a pH of ~3. All

polyelectrolyte and rinsing solutions were prepared with deionized ultra-pure water

(DIW) and filtered with a 0.2 µm sterile filter prior to use.

Materials used in preparation of hydrogels: Alginate Pronova UP LVG (low

viscosity (20-200 mPas) sodium alginate in which at a minimum 60% of the monomer

units are α-L-guluronate) was from Novamatrix, Norway. DOWEX ion exchanger,

tetrabutyl ammonium bromide, (D-(+)-glucono-δ-lactone), SO3/pyridine and anhydrous

DMF were purchased from Sigma Aldrich Chemie GmbH, Buchs, Switzerland.

Atelocollagen bovine type 1 collagen (Koken) was from Holzel Diagnostika, Koln,

Germany. CaCO3 particles of 5 µm size were obtained from PlasmaChem GmbH, Berlin.

MMP-degradable and non-degradable PEG powder without RGD-peptides (Reference

1004 and 1010) and disc casters were obtained from QGel, Lausanne, Switzerland.

Antibodies, peptides and staining materials: Immunoglobin G (IgG),

chondroitinase ABC (C2905), Triton X100, phalloidin TRITC labeled mixed isomers

(phalloidin-rhodamine) and Immunoglobin G (IgG), Fluoroshield™ with DAPI were

purchased from Sigma Aldrich Chemie GmbH, Buchs, Switzerland. Alpha 1, alpha 3 and

alpha5 integrin antibodies were obtained from Chemicon International. Type II

collagen antibody II-II6B3, type I collagen antibody M38, proteoglycan, hyaluronic acid

binding region (12/21/1-C-6) and AIIB2 beta1 integrin antibody were from

Developmental Studies Hybridoma Bank, University of Iowa. Monoclonal antibody

against native bovine lubricin (3A4) was from MD Bioproducts, Zurich, Switzerland.

Alexa 488 goat anti-mouse conjugated IgG IgM (H+L) and 4´,6-diamidino-2-

Chapter 3

21

phenylindole dilactate (DAPI) were from Invitrogen AG, Basel, Switzerland. The RGD

peptide with QGel linker was obtained from QGel, Lausanne, Switzerland. The GFOGER

peptide with the sequence GPC(GPP)5-GFOGER-(GPP)5GPC-NH2 was obtained from

professor Richard Farndale (University of Cambridge, Cambridge, UK) and preparation

of the peptide for hydrogel incorporation was followed according to an established

protocol [98]. 2mg GFOGER peptide was reduced in 10 mM acetic acid 2 mM TCEP

then heated for 2 min to 70 °C and allowed to form triple helices at 4 °C for 24 hours.

The triple helical trimer (11.1 KDa) was dialyzed against 10 mM acetic acid with a 3.5-5

KDa cutoff dialysis column to remove unfolded peptides and the TCEP.

3.2 Instruments

Quartz crystal microbalance with dissipation (QCM-D) and crystal

preparation: Assessment of the build-up of layer-by-layer films was followed using

QCM-D (Q-Sense E4, Gothenburg, Sweden). QCM crystals coated with thin PDMS

layers were cleaned for 30 minutes in 2% SDS then rinsed with DIW. The coated

crystals were then plasma treated in air for 2 minutes, mounted in the QCM-D flow cell

and 1ml of 150mM NaCl was injected. After the signal was stable adsorption of the

various layers and observation of the build-up was followed. Overtones 1, 3, 5 and 7

were monitored and the 3rd overtone was used for the assessment of films’ build-up.

QCM gold coated crystals were coated with a thin layer of PDMS using a method

previously described by Voros et al. [121]. Briefly, base silicone oil and crosslinker were

well mixed at a ratio of 10:3 (w/w) and diluted 1:100 (w/w) in hexane. The solution of

base silicon oil and crosslinker in hexane was spin coated on the gold coated QCM

crystals at 2000 rpm for 40 seconds. The crystals were then cured overnight at 70 °C to

allow polymerization of PDMS and to evaporate the hexane. Measurements of the

resonant frequency of the QCM crystals were taken before and after preparing the

thin PDMS layers and were used to calculate the thickness of the PDMS coating. The

PDMS film thickness calculated using the measured resonant frequency shift was 39.2

Materials and methods

22

±2.1 nm. The presence of PDMS coating on the QCM crystals and the films on the

PDMS were confirmed by contact angle measurements.

Atomic force microscopy (AFM): Atomic force microscopy (AFM) images were

taken using a JPK AFM instrument with CSC38 cantilevers (MikroMash).

Measurements were performed in contact mode in liquid at set points ranging from

0.1 V to 0.5 V for areas of 10 x 10 µm at a frequency fX range of 0.1 Hz to 1 Hz.

Quantitative real time polymerase chain reaction (qRT-PCR): Genes were

designed using Primer Select software and gene expression of relevant cartilage

markers was analysed using a StepOnePlus qRT-PCR machine (Applied Biosystems).

Homogenization of hydrogels was performed using a rotor stator homogenizer (TH 220

tissue homogenizer, Omni International, LabForce, Nunningen, Switzerland). Total

RNA was quantified using a spectrophotometer (Nanodrop ND-1000) and the

measured 260/280 ratio was consistently 2.0 ± 0.1 in all samples. Ribosomal protein

L13 (RPL13) was used as a housekeeping gene as it has been recently demonstrated to

be more stable than GAPDH [122]. The Livak method was used for analysis of qRT-PCR

data [123]. Investigated primers are listed in Table 3.1.

Microscopy: Images of the cells’ cytoskeletal organization and

immunohistological staining were taken with a confocal microscope (Carl Zeiss AG/LSM

510, equipped with a 40x 0.6 NA objective). Average cell area and image processing

were carried out using ImageJ 1.43 image processing program. For the GFOGER

experiments (Chapter 5) a Leica SP8 two photon microscope (Leica Microsystems,

Bensheim, Germany) was used to enable imaging deep within the opaque samples.

Macroscopic images of whole hydrogels were taken using a stereomicroscope (Leica

Chapter 3

23

WILD M650, equipped with a 6x objective). Other transmission images were acquired

using a Leica DFC420 C digital microscope camera.

Table 3. 1. Genes used in qRT-PCR analysis

Mechanical properties of the hydrogels: Compressive moduli of hydrogels

were measured using a texture analyser (TA.XTPlus, Stable Microsystems, UK). Samples

were compressed without preload at a speed of 0.01 mm/s and the force was

Materials and methods

24

measured at 0.1 mm compression, where the gels showed linear behavior. Young’s

moduli 𝐸 were calculated according to the formula:

𝜎 = 𝐸 ∙ 𝜀 (Equation 1)

where σ is the measured force per hydrogel area and ε is the strain.

Plate reader: The Synergy H1 Hybrid Reader (Biotek) was used to quantify

proliferation, DNA, RhoA activity and GAG. The plate reader allows top and bottom

UV-visible absorbance, fluorescence intensity, and luminescence detection in

microwell plates.

Mechanical strain machine and strain chambers: The STREX device (ST-140-10,

B-Bridge International) was used to apply mechanical strain to cells seeded in 2D and

3D. Silicon oil was well mixed 1:10 (w/w) with its crosslinker, then degassed PDMS was

poured in chamber molds (32 x 32 mm) and allowed to cure at 80 °C for 6 hours. For

the application of mechanical strain in 3D, STREX PDMS chambers were modified to

contain a series of support ridges to hold the gel as it is stretched. The homogeneity of

the applied strain was assessed by measuring the displacement of 10-30 μm glass

beads embedded in an alginate gel before and after strain. Assessment of

homogeneity was performed with 10% strain on 18 different areas within the strained

gel with 4 images taken in each area (n=3). Measurements of the strain values were

also determined for the other magnitudes applicable by the device.

3.3 Protocols

Chondrocyte isolation: Chondrocytes were harvested from the knees of 6

month old calves obtained from the local slaughterhouse as previously described

[124]. Cartilage shavings were minced using a sterile blade and treated with 0.2%

pronase in DMEM containing 1% antibiotic-antimycotic for 2 hours at 37°C, 7% CO2

with gentle stirring. Following the pronase digestion, the tissue was washed 3 x 2 min

with DMEM containing 1% antibiotic antimycotic then incubated 6 hours in 0.03%

Chapter 3

25

collagenase in DMEM supplemented with 1% antibiotic antimycotic at 37°C, 7% CO2

with gentle stirring. Cells were separated from the digested matrix by filtering through

a 100 μm cell strainer followed by a 40 μm cell strainer. Cells were counted and

viability was determined using an automated cell counter (Countess™ Automated Cell

Counter, Invitrogen AG, Basel, Switzerland). Cell viability in all isolations was above

90%. The isolated cells were seeded at 10,000/cm2 in DMEM supplemented with 1%

antibiotic antimycotic, 10% FBS and 50 µg/mL L-ascorbic acid. At 80-90% confluency

cells were released using trypsin/EDTA and seeded at 5,000/cm2 for the later passages.

Cell culture (Chapter 4): Passage 2 bovine chondrocytes were suspended in

DMEM supplemented with 1% Anti-Anti and 50 µM cycloheximide to prevent de novo

protein synthesis. The cells were seeded on different films of Col1/CS and Col1/HN

prepared on PDMS pieces for 4h at 37°C, 7% CO2. The films were blocked with 1% BSA

in PBS for 30 minutes prior to cell seeding. In order to determine if the cell spreading

was integrin mediated, cells were incubated with antibodies for beta1, alpha1, alpha3

and alpha5 integrins for 15 min prior to seeding on 2 bilayers of Col1/CS denoted as

(Col1/CS)2. IgG was used as control.

Cell culture (Chapter 5): Passage 3 or 4 bovine chondrocytes were mixed with

solutions of alginate, alginate sulfate or a mixture of alginate and alginate sulfate

sterile filtered through a 0.2 µm sterile Millipore filter. The final polymer concentration

was always kept at 2% (w/v) in 150 mM NaCl. Gels were cast in 30 μL volume disks

using the disc caster (Qgel, AG) and polymerized by placing the caster in a 50 mL falcon

tube containing 102 mM CaCl2 solution for 8 min. Disks were removed from the caster

with a spatula and incubated for 8 min free floating in a 60 mm petri dish containing 51

mM CaCl2 to allow equilibration of CaCl2 throughout the gel. Disks were then washed 3

x 2 min each in PBS containing 3 mM CaCl2 and 1 x 2 min in cell culture media

containing 3 mM CaCl2. Each disk was incubated in one well of non-treated 24-well

plates containing 1 mL cell culture media to which an additional 3 mM CaCl2 was

added. Medium was changed every 3 days. Cultures were carried out for 7, 21 or 35

days.

Materials and methods

26

Cell culture of human mesenchymal stem cells (hMSCs, chapter 6): Passage 2

human mesenchymal stem cells (hMSC, Lonza, Switzerland) were seeded and

expanded until passage 6 in DMEM supplemented with 10% FBS, 1% antibiotic-

antimycotic and 1 ng/mL FGF-2 which has been shown to maintain the multilineage

and chondrogenic potential of the stem [125]. Cells were released with trypsin/EDTA

and encapsulated in degradable or non-degradable PEG gels modified with 100 µM

RGD or 100 µM GFOGER peptide or non-modified according to the manufacturer’s

protocol (QGel, Lausanne, Switzerland). The cell/PEG solution was cast in discs of 30 µL

volume each. After 45 min incubation at 37 °C, discs were removed from the caster

and cultured in a non-tissue culture 24 well-plate pre-filled with 1 mL chondrogenic

media/well (DMEM 31966, 1% antibiotic antimycotic, 50 µg/mL L-ascorbic acid, 1%

ITS+ premix, 40 µg/mL L-proline, 100 ng/mL dexamethasone, 10 ng/mL TGF-β3). Non-

modified non-degradable gels cultured in medium without dexamethasone and TGF-β3

were designated as controls. Pellets were also prepared by centrifuging a 250,000 cell

suspension for 5 min at 300 g in 15 mL falcon tubes then culturing the pellets in 500 µL

chondrogenic or control media as differentiation controls. The media was changed

every 3 days.

Chondrocyte dedifferentiation assessment (Chapter 7): Cartilage tissue was

minced then placed in 350 μL RLT Plus Buffer (Qiagen AG, Zurich, Switzerland) and

homogenised using a rotor stator homogenizer (TH 220 tissue homogenizer, Omni

International, LabForce, Nunningen, Switzerland). Chondrocytes isolated from at least

3 different calves were seeded at 10,000 cells/cm2, these cells were designated as

passage 0 (P0). After 4 days of culture, part of the P0 cells seeded in a 25 cm2 tissue

culture polystyrene (TCP) flask were washed in PBS and lysed for mRNA analysis. At

80% confluence after 7±1 days, cells were trypsinized with 0.25% Tripsin/EDTA for 5

min, counted and seeded at 5,000 cells/cm2. Cells were passaged every 4±1 days or

lysed by adding 350 µL RLT plus buffer after washing once with PBS for mRNA

quantification. This was repeated till passage 4 (P4). The 260/280 ratio measured by

the spectrophotometer in all samples was consistently 2.0 ± 0.1. Total RNA was

Chapter 3

27

reverse transcribed starting with 500 ng RNA and gene expression of SZP and Col2 was

determined using qRT-PCR (StepOnePlus, Applied Biosystems). Ribosomal protein L13

(RPL13) was used as a housekeeping gene. Analysis of qRT-PCR data was performed

following the method of Livak.

Cell culture (Chapter 7): Chondrocytes at passage 3 or 4 were seeded on TCP at

5,000 cells/cm2 or embedded in alginate beads at 6x10⁶ cells/mL. To prepare alginate

beads, a cell pellet was suspended in 1.2% (w/v) alginate (Pronova UPLVG, Novamatrix)

and collected in a syringe connected to a 21-gauge needle. All alginate solutions were

sterilized by filtering through a 0.2 µm sterile Millipore filter. The alginate/cell mixture

was dispensed in a 102 mM CaCl2 mixture under continuous gentle stirring for 10min.

Beads were washed 3 x 2 min each with PBS and cultured in a 6-well plate containing

10-11 beads/well, 3 mL media/well. Cells in beads or TCP were incubated for 3-4 days

in a humidified incubator at 37°C under normoxic conditions (21% pO2) or hypoxic

conditions (1% pO2).

Application of mechanical strain (Chapter 7): For the application of strain in

2D, PDMS chambers were plasma treated for 2 min in air. The chambers were then

incubated with 0.1 mg/mL Col1 (Koken CosmoBio) in 150 mM NaCl, 1 mM HCl for 30

min to coat the chambers with a thin Col1 layer. Chambers were washed once with

150 mM NaCl and once with cell culture media followed by seeding passage 3

chondrocytes at 10,000 cells/cm2. 5% tensile strain at 1 Hz was applied for 2 hours, 24

hours post seeding for 4 days or once after 4 days of culture. Chondrocytes were

seeded on non-strained chambers and in 25 cm2 TCP flasks as controls. For application

of 3D strain, 0.75 mL of 2% alginate (w/v) + chondrocytes was injected in modified

PDMS chambers. Alginate was prepared using CaCO3 as a source for calcium ions

which results in uniform, slow-gelation of the material which can be injected [126,

127]. The gel was prepared by mixing 886 μL 2.25% alginate in 150 mM NaCl, 41 μL of

100 mg/mL CaCO3 in DIW, 73 μL of 200 mg/mL freshly prepared D-(+)-glucono-δ-

lactone (GDL) in DIW and 6 million cells. The solution was allowed to gel for 15-30 min

at 37°C. The hydrogels were covered with 5 mL of cell culture media after gelation and

Materials and methods

28

cultured for 24 hours. Mechanical strain was applied at 5.8% magnitude and 1 Hz for 2

hours, 24 hours post seeding for 4 days or once after 4 days of culture. Samples were

lysed 2 hours after the final strain was applied.

Layer-by-layer build-up assessment using QCM-D: Assessment of the build-up

of the films of Col1/CS and Col1/HN was followed using QCM-D (Q-Sense E4,

Gothenburg, Sweden). QCM crystals coated with thin PDMS layers were cleaned for 30

minutes in 2% SDS then rinsed with DIW. The coated crystals were then plasma treated

in air for 2 minutes, mounted in the QCM-D flow cell and 1ml of 150 mM NaCl was

injected. After the signal was stable, 0.5ml of Col1 dissolved at 0.1 mg/ml in 150 mM

NaCl, 1 mM HCl (pH ~3) was injected. After 30 min, crystals were rinsed with 150 mM

NaCl, 1 mM HCl, followed by injection of 1 mg/ml of CS or HN dissolved in 150 mM

NaCl (pH ~6). After 15 min incubation, a rinsing step with 150 mM NaCl was

performed. The steps were repeated for the desired number of layers. To test the

stability of the films at physiologic pH and in cell culture media, 1 ml of 10 mM HEPES,

pH 7.4, 150 mM sodium chloride (HEPES-2 buffer) was injected for 1 hour. Afterwards,

serum free media containing 1% antibiotic antimycotic was injected and the signal was

observed for 24 hours. Overtones 1, 3, 5 and 7 were monitored and the 3rd overtone

was used for the assessment of films’ build-up.

Chemical cross-linking of layer by layer films: Cross-linking was performed on

films of (Col1/CS)10 and (Col1/HN)10 prepared on PDMS pieces. 400 mM EDC and 100

mM NHS were freshly prepared in 150 mM NaCl. The substrates with the films were

crosslinked in the EDC/NHS solution for 12 h at 4 °C. [84] Films were then rinsed with

150 mM NaCl followed by deionized water and dried with a low stream of N2. Contact

angle measurements and cell adhesion assessment were then performed on the

prepared films.

Assessment of films’ topography and thickness: Silicon oil was well mixed 1:10

(w/w) with its cross linker. 3 mL of the degassed PDMS were poured in 35 mm cell

culture Petri dishes and allowed to cure at 80 °C for 6 hours. The Petri dishes were

then plasma treated in air for 2 min. The PDMS was rinsed with 150 mM NaCl and

Chapter 3

29

build-up of the films of Col1/CS or Col1/HN was initiated. Films were prepared using

the same protocol described in the QCM-D section with different number of layers and

different end layers. AFM images were taken using a JPK AFM instrument with CSC38

cantilevers (MikroMash). Fibril thickness was determined by measuring the height of

10 individual fibrils in a given image using the JPK image processing program and

calculating the average thickness and standard deviation. The thickness of the films

was measured by AFM using a previously-described method [128]. Briefly, a 2x2 µm

area of the sample was scanned several times at a high set point of approximately 5 V

and high frequency (10-20 Hz) to remove the film. A 10x10 µm area of the sample

including the scratched area was then scanned and the film thickness was determined

by measuring the average difference between 30 randomly chosen points of the

unscratched area and the scratched area using the JPK image processing program.

Cell viability: Cell viability was assessed using the Live/Dead® viability assay

from life technologies. Calcein AM is a green fluorophore that can only be activated

upon esterase cleavage by viable cells, while ethidium homodimer (EthD-1) is a red dye

that can only penetrate the membrane of a dead cell where it binds to DNA in the

nucleus. Cells in 2D or 3D were washed three times in warm PBS and subsequently

incubated in a solution containing 2 µM calcein AM and 4 µM EthD-1 for 30 minutes at

37 °C. Samples were then washed three times in warm PBS and then imaged directly

with a Zeiss LSM 510 Confocal Microscope (Carl Zeiss AG, Oberkochen, Germany) at a

minimum of three separate imaging planes per gel.

Cell morphology: Cell nuclei and the f-actin were visualized using double

labelling by DAPI and rhodamine-labelled phalloidin respectively. Cells were fixed with

5% formalin for 30 minutes at 4 °C, washed 3x with PBS, permeabilized with 0.4%

Triton X-100 for 20 minutes, and washed again. F-actin fibers were stained for 1 h

total with 0.13 µg/mL phalloidin. After 30 minutes, 2 µg/mL of DAPI was added at 1:1

(v/v) to the phalloidin solution for the remaining 30 minutes. All staining was done at

4 °C and in complete darkness. After staining, samples were washed in PBS and z-

Stacks were acquired using a two photon microscope and (SP8 Leica Microsystems,

Materials and methods

30

Bensheim, Germany). Samples grown in monolayer cultures or in translucent hydrogels

were imaged using a confocal microscope (Carl Zeiss AG/LSM 510, equipped with a 40x

0.6 NA objective).

Proliferation within hydrogels: Proliferation was assessed using a BrdU assay

Calbiochem® BrdU Cell Proliferation Assay (Calbiochem, Millipore, Switzerland). The

assay was performed according to the manufacturer’s recommendations with slight

modifications. Briefly, the gels were placed in media containing 1:200 BrdU label in a

96-well plate for 24 hours, post which the medium was carefully replaced with 270 μL

of dissolving buffer (0.055 M sodium citrate, in 0.03 M EDTA, 0.15 M NaCl, pH 6.8) and

the plate was rocked for 20 min at 1000 rpm at room temperature. For PEG gels the

dissolving solution used was trypsin. A volume of 30 µL was transferred to a new 96-

well plate and cells were centrifuged for 10 min at 1000 rpm, dried with an N2 stream

and fixed using the fixative solution supplied in the kit. Staining was then carried out

by an Anti-BrdU antibody followed by HRP conjugation and substrate addition

according to the manufacturer’s protocol. Absorbance was measured at dual

wavelengths (450-540) and the difference was blanked with the signal obtained from

cell free gels using a plate reader (Synergy H1 Hybrid Reader, Biotek).

DNA quantification in gels: Gels were collected at day 7 and day 21 for

assessment of DNA content and dissolved in 100 µL of papain lysis buffer (10 mM

EDTA, 100 mM sodium phosphate, 10 mM L-cysteine, 125 µg/mL Papain type III at pH

6.3) overnight at 60 ˚C with shaking at 1000 rpm. DNA quantification was carried out

with Quant-iTTM PicoGreen dsDNA Kit (Invitrogen). All samples were diluted 1:10 in TE

buffer supplied by the kit. DNA standard was diluted in TE Buffer to 2 μg/mL, 200

ng/mL, 20 ng/mL, 2 ng/mL and 0.2 ng/mL concentrations. Lysis buffer was diluted 1:10

and Picogreen reagent was diluted 1:200 in TE buffer. A volume of 100 µL of each

sample and DNA standard was transferred to a 96-well plate in triplicates and an equal

volume of PicoGreen solution was added in each well. Measurement was performed

using an excitation wavelength of 480 nm and emission wavelength of 520 nm with a

plate reader.

Chapter 3

31

1,9-Dimethylmethylene blue (DMMB) assay: DMMB dye was prepared

according to Estes et al. [129]. Briefly, 21 g of DMMB was dissolved in 5 mL ethanol

with 2 g sodium formate and mixed into 800 mL distilled water. The pH was adjusted

to 3.0 with formic acid followed by bringing up the solution to 1 L. Gels were collected

at day 7 and day 21 for assessment of GAG content and dissolved in 100 µL of papain

lysis buffer (10 mM EDTA, 100 mM sodium phosphate, 10 mM L-cysteine, 125 µg/mL

Papain type III at pH 6.3) overnight at 60 ˚C with shaking at 1000 rpm. Chondroitin 4-

sulfate (Sigma) was diluted in papain buffer dilutions from 0 to 35 µg/mL were used to

create the standard curve. A volume of 40 µL of each sample or standard was pipetted

into a 96-well plate. A volume of 125 µL of the DMMB dye was added to each well and

optical density was measured at 595 nm with a plate reader. All GAG quantification

was normalized to DNA content for each gel.

Gene expression: Samples in 2D were washed once in PBS followed by addition

of 350 μL of RLT plus buffer immediately to chambers or TCP. Alginate and alginate

sulfate samples were washed once with PBS and dissolved completely in 1 mL sodium

citrate dissolving buffer. Samples were centrifuged at 10,000 rpm for 2 min, re-

suspended in PBS and centrifuged at 14,500 rpm for 2 more minutes. The pellet was

then lysed in 350 μL of RLT Plus Buffer. Expression of superficial zone protein was

determined by qRT-PCR using RPL13 as a housekeeping gene. In alginate sulfate

samples, 0.5% (w/v) BSA was added to the qRT-PCR mixture to prevent interference of

the sulfate groups with the amplification reaction [130]. PEG gels and pellets were

homogenised using a rotor stator homogenizer (TH 220 tissue homogenizer, Omni

International, LabForce, Nunningen, Switzerland). A 540 μL of RNA free water and 10

µL Protease K solution (Qiagen) were added to each sample. Samples were incubated

for 10 min at 55 °C with continuous shaking at 1000 rpm. RNA was isolated using the

RNeasy Mini Kit (Qiagen) following the fibrous tissue isolation protocol given by the

manufacturer. Total RNA was quantified using a spectrophotometer (Nanodrop ND-

1000) and the measured 260/280 ratio was consistently 2.0 ± 0.1 in all samples. Total

RNA was reverse transcribed and gene expression of type II collagen (Col2), type I

collagen (Col1) and aggrecan (Agg) were determined using quantitative real time PCR

Materials and methods

32

(StepOnePlus, Applied Biosystems). Ribosomal protein L13 (RPL13) was used as a

housekeeping gene and the Livak method was used for analysis of qRT-PCR data [123].

Immunostaining: Samples were washed once with PBS, fixed with 4%

paraformaldehyde, 0.2% Triton-X100 for 30 min then washed once with PBS. Alginate

samples were stored in PBS containing 10 mM CaCl2 to prevent gel dissolution.

Samples were embedded in Optimum Cutting Temperature (O.C.T) compound and

frozen on a dry ice block for 5 min and 6 μm-thick slices were cut using a microtome

(CryoStar NX70, ThermoScientific). The slices were fixed in ethanol then washed with

10 mM CaCl2 in PBS. Antigen retrieval was performed using pronase at 1 mg/mL for 15

min at 37 °C for Col2 and Col1 and using chondroitinase at 0.02 U/mL for 40 min at

37 °C for proteoglycans. Samples were then blocked with 5% BSA in PBS for 1 h at

room temperature. Samples were incubated with primary antibodies specific against

Col2 (II-II6B3, Developmental Studies Hybridoma Bank), Col1 (M38, Developmental

Studies Hybridoma Bank) and proteoglycan, hyaluronic acid binding region (12/21/1-C-

6, Developmental Studies Hybridoma Bank) at 1:10 dilutions in 1% BSA in PBS

overnight at 4 °C. Samples were washed 3x in PBS and incubated with 1:400 Alexa-488

goat anti-mouse secondary antibody for 1 h at room temperature. Controls were

prepared following the same procedure but omitting the primary antibody. Samples

were imaged using a confocal microscope (Carl Zeiss AG/LSM 510, equipped with a 40x

0.6 NA objective).

Alcian blue staining: Frozen sections of the hydrogels were defrosted at room

temperature, washed in PBS for 5 minutes and fixed with 4% formaldehyde in PBS for

7 minutes. Samples were washed 3x 5 minutes each in PBS, hydrated with DIW and

then stained in alcian blue solution (pH 2.5) for 20 minutes. Slides were washed in

running tap water until unspecific staining was washed away, samples were rinsed in

DIW and dehydrated using ascending alcohol solutions (95% to absolute ethanol) for 3

minutes each. Sections were cleaned in xylene, air-dried and mounted with resinous

mounting medium (Mounting Medium ref. 4112, Richard-Allan Scientific). A cover slip

Chapter 3

33

was added on the top of the section and sealed with nail polish. Images were acquired

using a Leica DFC420 C digital microscope camera equipped with a 40x objective.

Statistical analysis: All quantitative data were obtained from at least 3

independent donors and expressed as the mean ± standard error. Statistical evaluation

was carried out by analysis of variance (ANOVA) and post-hoc Tukey’s tests where P

values of less than 0.05 were considered significant. Statistical analysis was performed

using OriginPro version 8.1.

Materials and methods

34

Chapter 4

35

4 Layer-by-Layer Films Made from

Extracellular Matrix Macromolecules on

Silicone Substrates

The work in this chapter has been published in R.F. Mhanna et al., Biomacromolecules 2011;

12: 609-616

In the introduction we identified the current problems with monolayer cultures

mainly represented by chondrocyte dedifferentiation. We hypothesized that the use of

molecules from ECM origin might be a tool to improve the chondrogenic performance

of serially passaged chondrocytes. Furthermore, such molecules can also be used to

improve biocompatibility of implanted prosthetic devices. In this chapter we

demonstrate the build-up of Col1/CS and Col1/HN films prepared using the layer-by

layer technique.

4.1 Current monolayer culturing techniques

The layer-by-layer (LbL) technique has been widely used to produce nanofilms

for biomedical applications [79, 80]. The LbL approach allows the build-up of thin films

simply by the alternating deposition of polyanions and polycations [81]. A wide variety

of polymeric compositions both synthetic [131] and natural [82, 88, 132] have been

studied since the first introduction of the technique by Decher and coworkers [81]. LbL

films have been used to improve cytocompatibility of poly ethylene terephthalate and

poly(l-lactic acid) [133, 134]. Films can be tuned to favour cell adhesion [83, 85] or

hinder cell adhesion through the addition of antifouling end layers such as

poly(ethylene glycol) [121, 135].

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

36

Despite their success in cell cultures, the use of LbL engineered multilayers in

biomedical applications raised the concern of cytotoxicity of these films. Recent

studies show that polymers used to prepare films such as the traditional

polystyrenesulfonate/polyallylamine hydrochloride (PSS/PAH) multilayers can be

cytotoxic if released to the culture media [87]. Thus having systems of natural

polymers such as ECM macromolecules or other naturally occurring proteins is of high

importance. Some ECM molecules such as chondroitin sulfate (CS) and heparin (HN)

contain sulfate groups that render these polymers negatively charged at physiological

pH [91]. These can then be used in combination with type I collagen (Col1) [90] which

is protonated at a pH below its isoelectric point (5.5) [136] to form natural thin films,

for a variety of applications. CS is the most abundant GAG in plasma [137] and has

been shown to increase proteoglycan synthesis and to have roles in

mechanotransduction of articular cartilage [37, 38]. CS is available as a food

supplement and as a disease protective drug for arthritis patients [27, 39]. On the

other hand HN is a well-known anticoagulant and has been used in various biomedical

applications [91]. All these polymers have been used for generation of scaffolds for

tissue engineering especially Col1 which has been well tested in-vivo for the delivery of

chondrocytes in autologous chondrocyte implantation [29, 138-140].

Polydimethylsiloxane (PDMS) is a common elastic biomaterial used in a wide

range of biomedical applications including microfluidic devices [141], breast implants

[142], contact lenses [143], cardiac pacemakers, glaucoma drainage devices [144] and

the study of cells under static and cyclic strain [108, 145]. PDMS is naturally

hydrophobic and cell repellent. One of the most common methods used to modify

PDMS is plasma treatment which creates free hydroxyl groups on its surface rendering

PDMS hydrophilic [146]. This process is reversible as PDMS can recover hydrophobicity

within days or even hours at normal temperature and in dry environment due to the

migration of non-crosslinked monomers to the surface. Hydrophobicity recovery can

be delayed by storing PDMS in water and low temperature [147, 148]. A more efficient

Chapter 4

37

and more stable modification of the PDMS surface is by adsorption of hydrophilic

polymers on its surface [144].

In this study we assessed the build-up, stability, topography and thickness of

films prepared from (Col1/CS) and (Col1/HN). Moreover, we assessed wettability and

cell adhesion on these LbL films. The current work provides insight for the use of these

layers for biomedical applications which make use of PDMS substrates.

4.2 Build-up of Col1/CS and Col1/HN films on PDMS

The QCM-D data showed a clear build-up of the Col1/CS and Col1/HN systems

on the PDMS coated crystals (Figures 4.1 and 4.2). After Col1 adsorption the frequency

decreased about 70 Hz for all crystals with a high increase in dissipation indicative of

loose and highly hydrated layers. After the first Col1 layer, adsorption of HN or CS

caused an increase in the frequency of about 50 Hz with a decrease in the dissipation

in both systems. In the following layers, build-up of the Col1/HN system had larger

frequency drop steps than the Col1/CS (Figure 4.3).

In the Col1/CS system the adsorption of CS caused a decrease in the dissipation

and increase in frequency indicating that the film became stiffer as a negative layer

was adsorbed, a phenomenon often observed in the stratification of LbL films (Figure

4.1). The injection of CS solution (pH~6) might also initiate crosslinking of collagen

monomers induced by the local change in the film pH resulting in the formation of

fibrils. The adsorption of Col1 resulted in a decrease in the frequency and increase in

the dissipation regardless of the layer number.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

38

Figure 4. 1. | Buid-up of Col1/CS multilayer. Normalized resonance frequency shift and dissipation shift

of the 15 MHz detection frequency for the build-up of 6 bilayers of Col1/CS (a). A zoom in of the

frequency shifts showing the rinsing steps following adsorption of the first Col1 layer and first CS layer

are depicted in (b) and (c) respectively.

In the Col1/HN system the adsorption of HN resulted in a decrease in the

dissippation and increase in the frequency until the 3rd bilayer (Figure 4.2). This

behavior was slowly inverted with the subsequent layers which exhibited smaller

dissipation and larger frequency steps. A decrease in the frequency and increase in the

dissipation was observed in all Col1 adsorption steps. The value of the frequency and

dissipation shifts increased with the number of layers.

Chapter 4

39

Figure 4. 2. | Buid-up of Col1/HN multilayer. Normalized resonance frequency shift and dissipation shift

of the 15 MHz detection frequency for the build-up of 6 bilayers of Col1/HN (a). A zoom in of the

frequency shifts showing the rinsing steps following adsorption of the first Col1 layer and the first HN

layer are depicted in (b) and (c) respectively.

A linear build-up was observed with the Col1/CS films (Figure 4.3), which is in

aggreement with the results reported by Zhang et al. [82] on the build-up of

Col1/hyaluronic acid (HA). On the other hand, the build-up in the Col1/HN system was

exponential, indicating a less compact structure that permits the diffusion of one or

both polymers (Figure 4.3). The exponential growth of LbL films constituting

polysacharides such as chitosan (CH) and HA have been previously observed and well

characterized [149, 150]. Diffusion of CH molecules is believed to drive the exponential

growth of these films. In the CH/HA films, isolated islets are formed during the first few

layers which coalesce to vermiculate structures as the number of deposited layers

increases. This growth behavior differs from the systems presented in this study and

other systems prepared with Col1 as the cationic polymer [82, 88], which exhibit a

fibrillar structure covering the surface from the initial layers.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

40

Figure 4. 3. | Buid-up exponential (Col1/HN) vs. linear (Col1/CS). Normalized resonance frequency

shifts of the 15MHz detection frequency for the build-up of 6 bilayers of Col1/CS (○) and Col1/HN (□) as

a function of the number of deposited bilayers.

Injecting HEPES-2 buffer in a final step in the Col1/CS system resulted in a

decrease in the frequency with an increase in the dissipation. This phenomenon can be

attributed to the loss of the collagen charges due to the pH change that causes an

increased hydration of the film and consequently a dissipation increase. Injection of

media in a following step caused a slight increase in the frequency which reached a

plateau after 2 hours. No significant change in the dissipation was observed which

indicates that the film might lose some of the weakly bound CS molecules that can be

replaced by sugars or other charged molecules in media. On the other hand the

Col1/HN system showed a high increase in frequency with a high increase in

dissipation upon addition of HEPES-2 buffer. This is a highly unusual finding in QCM

experiments, and has been reported to be due to either the formation of long

viscoelastic structures that remain anchored to the substrate [151] or to viscoelastic

effects only observed in case of thick highly dissipative films [152]. The comparison of

our results to other QCM-D studies of multilayers such as Richert et al. [149] or Kujawa

et al. [150] is difficult as the studied species here are not only of varying molecular

Chapter 4

41

weight but also of different structure and charge density. In addition, in our case the

formation of fibres induced by pH change further complicates the system.

Build-up with Col1 was only possible at pH less than 5. As the isoelectric point

of Col1 is about 5.5 [136], if Col1 is rinsed with a solution of pH higher than 5, most of

the amino charged groups of collagen will be eliminated preventing electrostatic

adsorption of negatively charged CS or HN. At low pH Col1 is highly charged causing

strong repulsion between the Col1 monomers. At higher pH such as physiological pH,

the charges carried by Col1 are fewer which facilitates their self-assembly into larger

fibers [153].

4.3 Effect of substrate on film build-up

The build-up of these ECM-based films was dependent on the underlying

substrate. The initial Col1 adsorption step on the gold substrate was much larger than

on PDMS indicated by the almost 8-fold larger decrease in frequency and increase in

dissipation (Figure 4.4). After the 2nd bilayer, the frequency shifts of collagen become

similar to the shifts observed on PDMS. Furthermore, adsorption of CS on Col1 appears

to cause a further decrease in the frequency and decrease in dissipation. The decrease

in the frequency in this case can be attributed to the large amount of collagen

available for CS binding. It was clear from our experience that results obtained on gold

could not be used to estimate the build-up on PDMS. Therefore it is essential to assess

build-up of films on the actual substrates used in a given application.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

42

Figure 4. 4. | Build-up of Col1/CS on gold. Normalized resonance frequency shift and dissipation shift of

the 15 MHz detection frequency for the build-up of 6 bilayers of Col1/CS on gold.

4.4 Film topography

AFM images of Col1/CS and Col1/HN showed a fibrillar structure of the films

(Figure 4.5). The fibrils in the Col1/CS system increased in diameter from 10.1±1.4 nm

for (Col1/CS)2 (Figure 4.5A) up to ~19.2±3.25 nm for (Col1/CS)10 (Figure 4.5D).

Individual fibrils of about 1-2 µm in length were observed after 2 bilayers (Figures 4.5A

and B). The length of the fibrils increased with number of layers deposited and after 10

bilayers fibrils of 4-5 µm in length could be observed (Figure 4.5D). In the Col1/HN

system, films had sparse fibrils of 9.2±1.2 nm diameter for (Col1/HN)2 (Figure 4.5E and

F). A meshwork of fibrils was observed for (Col1/HN)5 and (Col1/HN)10 with no

significant change in fibrils` diameter (Figure 4.5G and H). The addition of a collagen

layer clearly increased the thickness of the fibrils in both systems which can be

observed in (Figure 4.5B and 5F).

Chapter 4

43

Figure 4. 5. | Morphology of Col1/CS and Col1/HN films. 10x10 µm AFM Images for various assemblies

of different layer numbers of Col1/CS and Col1/HN taken at room temperature in liquid (150 mM NaCl,

pH~6) in contact mode. (Col1/CS)2 (A), (Col1/CS)2Col (B), (Col1/CS)5 (C), (Col1/CS)10 (D), (Col1/HN)2 (E),

(Col1/HN)2Col (F), (Col1/HN)5 (G) and (Col1HN)10 (H).

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

44

4.5 Film thickness and stability

The thickness of the films was measured with AFM by scratching with the tip of

the cantilever (Figure 4.6). This method ensures an accurate measurement of the film

thickness assuring no damage to the underlying PDMS substrate. Difficulties in

obtaining perfect squared scratches were due to the thick fibrils in the Col1/CS system

which would be bundled together when scanned with the cantilever for many times.

The thickness of the (Col1/CS)10 film was 15.6 ± 6.8nm (Figure 4.6A) while (Col1/HN)10

was 26.4 ± 9.4 (Figure 4.6C). In order to confirm the stability results obtained from the

QCM-D curves, scratches on (Col1/CS)10 and (Col1/HN)10 were performed before and

after incubating the films with HEPES-2 buffer and tissue culture media. The thickness

and also the topography of (Col1/CS)10 were stable for 24 h in HEPES-2 buffer and in

tissue culture media (Figure 4.6B) which is in accordance with the QCM-D data. On the

other hand films of Col1/HN maintained their fibrillar topography when moved to

HEPES-2 buffer or DMEM but the thickness of the films was reduced to 8.9 ± 3.5 nm in

DMEM (Figure 4.6D).

Chapter 4

45

Figure 4. 6. | Films’ thickness and stability. AFM image with a scratched region for (Col1/CS)10 in 150

mM NaCl, pH~ 6 as prepared (A), and after incubation in media for 24 h (B), AFM image with a

scratched region for (Col1/HN)10 in 150 mM NaCl, pH~6 as prepared (C) and after incubation in media

for 24 h (D). The scanned area was 10x10 µm and the scratch was 2x2 µm.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

46

4.6 Assessment of cell adhesion and integrin mediated spreading

In order to investigate cell adhesion to our synthesized substrates and not

extraneous proteins, cells were incubated with 1) cycloheximide, to avoid influences of

adhesive proteins synthesized by the cells themselves and 2) in serum free media to

prevent influences of serum proteins on cell adhesion (Figure 4.7). Under these

conditions cell spreading was superior on ECM films compared to PDMS and tissue

culture plastic (TCP) (Figure 4.8A), suggesting that normal spreading on TCP is

dependent on proteins synthesized by the cells or serum proteins. The average cell

area on the (Col1/CS)2 film was 2025 ± 1085 μm2 which was significantly greater than

the average cell area on TCP 661 ± 611 μm2 using the Student’s t test (p < 0.05). Cells

under these conditions do not adhere well to PDMS, however, the ECM film modified

PDMS allowed strong cell adhesion, where even application of 10% tensile strain to a

PDMS membrane for 24 h was possible without noticeable cell detachment. No

significant difference in cell adhesion and spreading was found between the different

films except the (Col1/HN)10 (Figure 4.7C) which exhibited a high percentage of round

cells with a significantly (p < 0.05) lower average cell area of 977 ± 781 μm2. Adhesion

to (Col1/HN)10 was possibly reduced due to the high film hydration indicated by the

large increase in dissipation as observed with QCM-D. The stiffness of the substrate

has been shown to play an important role on cell adhesion. In a study by Richert et al.

[84], chondrosarcoma cells adhered and spread only on cross-linked Poly(L-

lysine)/hyaluronan (PLL/HA) films but not on native films of the same composition.

In our work, cell spreading was limited by crosslinking, where the cells had a

more restricted spreading on the substrate, expressed more filopodia and exhibited a

ruffled morphology (Figure 4.7E and F). The uncrosslinked, native films used in this

study were not as hydrated as the non-cell adhesive (PLL/HA) films and allowed better

access to collagen binding sites, whereas crosslinking increased the stiffness of the

substrate and limited the cell's access to collagen binding sites.

Chapter 4

47

In the Col1/CS system, cell morphology was not affected by the number of

layers, a result which can be explained by the similarity in topography of the films

(Figure 4.7B, D and F). The end layer also had no significant effect on cell adhesion.

Since chondrocytes do not possess specific receptors for HN, and their adhesion is

mainly dictated by Col1, this observation indicates that Col1 must be accessible for the

cells even for CS and HN ending films. This contrasts with the study by Zhang et al. [82]

who found chondrosarcoma cultured on Col1 ending films formed pseudopod-like

attachment on the collagen fibrils, but such structures did not form on hyaluronic acid

(HA) ending films. These results suggest that although HA might form a completely

uniform film over the Col1, CS and HN layers are not continuous and allow cells to

access collagen from deeper layers.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

48

Figure 4. 7. | Morphology of bovine chondrocytes seeded ECM films. Phalloidin staining of bovine

chondrocytes on different films and integrin mediated adhesion assessment. Cells were incubated for 4h

on (Col1/HN)2 (A), (Col1/CS)2 (B), (Col1/HN)10 (C), (Col1/CS)10 (D), cross linked (Col1/HN)10 (E), cross

linked (Col1/CS)10 (F). Images were taken using a 40x objective, scale bar 20µm.

Chapter 4

49

In order to prove that cell adhesion was integrin-mediated, cells were

incubated with antibodies to alpha1, alpha3, alpha5, beta1 integrins and IgG as

control and then cultured on films of (Col1/CS)2 (Figure 4.8). Beta1 integrin forms

dimers with several alpha subunits and controls interaction with several ECM

molecules including collagen, fibronectin and vitronectin. Alpha1 binds the monomeric

collagen fibril [154] and laminin, alpha3 binds laminin and thrombospondin, and

alpha5 binds fibronectin and fibrinogen [155]. Cell adhesion and spreading were

hindered only with the beta1 integrin antibodies (Figure 4.8C), while the other

integrins had no effect on adhesion or cell spreading. Less cell adhesion was observed

when beta1 integrin antibodies were used compared to the control and other

antibodies. Furthermore, cell area was significantly lower for cells incubated with

beta1 integrin antibodies (364 ± 142μm2 compared to IgG 1927 ± 733μm2 ) whereas

incubation with other antibodies caused no significant difference in average cell area.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

50

Figure 4. 8. | Integrin mediated cell spreading. Phalloidin staining of bovine chondrocytes seeded on

(Col1/CS)2 and incubated for 4h with antibodies to specific integrins. Tissue culture plastic (A), (Col1/CS)2

with IgG antibody (B), (Col1/CS)2 with beta1 integrin antibody (C), (Col1/CS)2 with alpha1 integrin

antibody (D), (Col1/CS)2 with alpha3 integrin antibody (E), (Col1/CS)2 with alpha5 integrin antibody.

Images were taken using a 40x objective, scale bar 20µm.

Chapter 4

51

4.7 Chapter summary

The goal of this study was to coat deformable PDMS substrates with specific

ECM macromolecules, so that cell/ECM interactions could take place under controlled

conditions. The thickness and morphology of films made from Col1/CS on PDMS were

stable in cell culture media. Films made from Col1/HN maintained their topography

but had a loss in the total film thickness. The substrate has a great influence on the

build-up of the films where adsorption of Col1 on gold was 8 times larger than on

PDMS. Cell adhesion was mediated mainly by beta1 integrins and cells had a

constrained spreading and ruffled morphology on crosslinked films compared to non

crosslinked. This study concerns the use of ECM molecules to coat surfaces for

studying specific ECM/cell interactions. This work is of major importance for studies of

mechanical loading of cells on PDMS. Further gene expression studies should help

reveal the role of chondroitin sulfate, heparin and other glycosaminoglycans in

cartilage mechanobiology.

Layer-by-Layer Films Made from Extracellular Matrix Macromolecules on Silicone Substrates

52

Chapter 5

53

5 Chondrocyte Culture in 3D Alginate Sulfate

Hydrogels Promotes Proliferation While

Maintaining Expression of Chondrogenic

Markers

The work in this chapter has been submitted to Biomaterials, Mhanna et al. (2013A).

In the previous chapter, we were able to prepare films made from extracellular

molecules relevant for cartilage engineering. We observed that stable films containing

the chondroprotective chondroitin sulfate molecule could be made at varying

thicknesses. However, cells cultured on these films maintained a spread fibroblast like

phenotype which motivated us to develop a 3D system that may provide similar

biological cues. To achieve this, alginate a commonly used biomaterial known to

maintain the cartilage phenotype was modified with sulfate groups and the material

was investigated for its potential use for cartilage tissue engineering.

5.1 Improving chondrogenic performance of chondrocytes in 3D

may be achieved using biomimetic materials.

Articular cartilage does not regenerate after injury or disease. Developing

strategies to restore the function of the damaged tissue is a major objective for tissue

engineers [12, 156]. Cell based approaches used for treatment of cartilage lesions

include microfracture, osteochondral grafting and autologous chondrocyte

implantation (ACI) [13, 14, 50]. Recently ACI increased in popularity and showed

promising results for treating medium and large chondral defects [49, 157, 158]. ACI

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

54

first described by Brittberg et al. [52] in 1994, is a two-step surgical procedure. In the

first stage a cartilage tissue biopsy harvested from the patient is enzymatically digested

to separate cells from the tissue and the cells are grown in a laboratory to reach

approximately 12 million cells. In the second stage, the cells are injected in the defect

and covered with a periosteal flap or membrane [57]. One problem associated with

use of these cells is the dedifferentiation which occurs during monolayer expansion

[59, 159]. Dedifferentiated chondrocytes express fibroblastic genes such as type I

collagen and produce fibrous tissue rather than hyaline cartilage in the defect site [50,

160]. Additionally, donor site morbidity from harvesting the periosteal flap and the

need for a second surgery are some of ACI’s other limiting factors. Taken together,

these limitations have opened the quest for 3D scaffolds for cartilage tissue

engineering. Three dimensional culture systems such as agarose [60] and alginate [72,

161, 162] restore the function and phenotype of chondrocytes. However, cell

proliferation is limited in alginate and agarose restricting their use in priming cells for

ACI-like procedures.

The cartilage extracellular matrix (ECM) is composed of 10-20% proteoglycans

[50]. The core proteins of proteoglycans are heavily modified by glycosaminoglycans

(GAGs) including chondroitin sulfate, keratan sulfate and dermatan sulfate. These

linear polysaccharides have sulfate and carboxylic groups which contribute to the high

fixed negative charge of the proteoglycans at physiologic pH. The GAG side chains also

give the tissue its compressive and osmotic swelling properties by entrapping water

[13] .

Sulfation of molecules has purported effects on the biological activity of

biopolymers. Chondroitin sulfate has been shown to have therapeutic benefits for hip

and knee osteoarthritis [27, 39]. Heparan sulfate has a high affinity to a plethora of

growth factors crucial for cartilage homeostasis [163, 164]. Chitosan sulfation

enhanced fibroblast adhesion and contraction of a collagen lattice compared to the

unsulfated material [165]. Freeman et al. [166] studied the binding affinity of 10

different heparin-binding factors (including FGF, IGF and VEGF) to alginate sulfate.

Chapter 5

55

They found that except FGF most of the factors bind equally well or better to sulfated

alginate compared to heparin. Reem et al. showed that scaffolds containing a mixture

of alginate and sulfated alginate caused attenuated TGFβ1 release and consequently

improved chondrogenesis of entrapped mesenchymal stem cells compared to scaffolds

lacking alginate sulfate [167]. The scaffolds used in these studies were made by

freezing alginate sulfate into a macroporous scaffold [166]. In this study we report for

the first time a sulfated alginate hydrogel which is crosslinkable with calcium.

We hypothesized that modification of alginate with sulfate groups provides a

more chondrogenic environment compared to pure alginate. To test this hypothesis

alginate was modified with sulfate groups and chondrocytes were encapsulated in

alginate and alginate sulfate at varying ratios. Cell morphology, proliferation, RhoA

GTPase activity, gene expression, and synthesis of cartilage markers were analyzed.

5.2 Preparation and characterization of sulfated alginate

Alginate 2 g (5.0 mmol) was dissolved in 400 mL water and 40 g DOWEX

Marathon C ion exchanger that was previously charged with an equal mass of

tetrabutyl ammonium bromide. The mixture was stirred overnight, filtered and

isolated by lyophilisation resulting in a total final mass of 2.7 g. Alginate tetrabutyl

ammonium salt 2 g (2.4 mmol) was suspended in 200 ml dry DMF and a 5-fold excess

SO3/pyridine per disaccharide repeating unit was added and the mixture was stirred at

room temperature for 1 h. The opaque solution was precipitated in acetone, brought

to pH=12 (ethanolic NaOH) for 10 minutes and then neutralized. The precipitate was

filtered, dissolved in water and purified by dialysis. Lyophilisation gave the pure

product of 1.8 g in mass. The process of alginate sulfate synthesis is demonstrated in

Figure 5.1. An important parameter to characterize the chemical composition of the

alginate sulfate is the degree of substitution by sulfation (DSS). It gives the average

number of sulfate groups per disaccharide repeating unit of alginate which is formed

from β-D-mannuronate and α-L-guluronate. Based on this definition DSS values may

range between 0 (unsubstituted alginate) and 4.0 (complete substitution of all free

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

56

hydroxyls by sulfates). The DSS was determined by estimation of the sulfur content

using an automatic elemental analyser (CHNS-932, Leco, Moenchengladbach,

Germany). The degree of sulfation of the material used in this study was DSS = 0.8.

Two other materials with DSs of 1.1 and 2.6 were synthesized but did not form

hydrogels with 102 mM calcium and were not further considered. IR spectra were

obtained on an FT-IR-Spectrometer FTS 175 (BIO RAD, Krefeld, Germany) applying ATR

technique. NMR spectra were recorded in D2O at 343 K on a Bruker Advance 300 MHz

spectrometer. Spectral data found for the synthesized material were consistent with

those found in the literature for alginate sulfates prepared by different procedures

[166, 168, 169]. IR (ATR): 3445, 2946, 1613, 1417, 1232, 1169, 1021, 951, 831, 796 cm-

1. 13C-NMR (D2O): 178.1-177.6 (C 6), 102.3-101.8 (C-1), 80.9-69.9 (C-2, C-3, C-4), 66.3

ppm (C-5).

1. SO3/pyridine, DMF2. ethanolic NaOH

R = H, SO3Na

N+ Br-

DOWEX

N+2

OO

OH

OHNaOOC OH

O

HO O

NaOOC

OO

OH

OH-OOC OH

O

HO O

-OOC

OO

OR

ORNaOOC OR

O

RO O

NaOOC

Figure 5. 1. | Reaction scheme for the synthesis of alginate sulfate. In a first reaction step sodium

alginate is transformed into a tetrabutyl ammonium salt to improve the solubility in the reaction

medium followed by the sulfation of free hydroxyl groups of alginate with SO3/pyridine.

Chapter 5

57

5.3 Morphology of chondrocytes encapsulated in sulfated alginate

(DSs = 0.8)

Transmission and confocal microscopy were used to assess changes in the

cytoskeletal organization. The morphology of cells encapsulated in the different ratios

of unmodified to sulfated alginate hydrogels was similar except for cells cultured in

pure alginate sulfate (2%), which assumed a more spread, fibroblast-like morphology

after 7 days in culture (Figure 5.2E). Cells encapsulated in 1.5% alginate sulfate + 0.5%

alginate (Figures 5.2D and I) exhibited minor spreading and protruding filopodia. Cells

in 2% alginate, 0.5% alginate sulfate + 1.5% alginate and 1% alginate sulfate + 1%

alginate (Figures 5.2A-C and F-H) had a typical round morphology as commonly

observed for alginate encapsulated chondrocytes. Proliferation was visually higher in

the pure alginate sulfate hydrogel (Figures 5.2E and J) and more multi-nucleated cells

could be observed in the 1.5% alginate sulfate + 0.5% alginate (Figures 5.2D and I)

hydrogels compared to hydrogels with lower alginate sulfate content.

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

58

Figure 5. 2. | Morphology of cells encapsulated in alginate and alginate sulfate. Phalloidin/DAPI

staining and transmission imaging of P3 bovine chondrocytes encapsulated in alginate (Alg), alginate

sulfate (AlgSulf) and mixtures of Alg and AlgSulf for 7 or 21 days.

Sulfated GAGs are present in the ECM of most tissues; among these are

heparin, heparan sulfate, chondroitin sulfate, keratin sulfate and dermatan sulfate [13,

170]. These molecules have been shown to have important functions such as

anticoagulant [171, 172], growth factor and cytokine affinity [163, 164],

chondroprotective [27, 39], neuroprotective [173] and antioxidant properties [174].

Lately, Ronghua et al. [169] showed that alginate sulfate has comparable anticoagulant

properties to heparin. The use of alginate sulfate for tissue engineering applications is

limited to a few studies [167, 175]. In particular to our knowledge, a pure alginate

sulfate hydrogel was not previously reported. Different sulfating agents including

sulfuric acid-carbodiimide [166], chlorosulfonic acid-formamide [169] and various

Chapter 5

59

SO3/complexes [176, 177] have been used in the past to prepare sulfated alginate and

differences in the sulfation method may explain why calcium crosslinking was not

previously possible. Recently, it was shown that SO3/pyridine is a mild sulfating agent

for GAGs such as hyaluronan allowing proper control of the degree of sulfation (DSS),

especially if a homogeneous sulfate group distribution and low DSS values are desired

[178, 179]. We therefore decided to use SO3/pyridine for the sulfation of alginate. In

our work, addition of 3 mM CaCl2 was necessary to maintain stability of the gel for the

first 21 days of culture, however CaCl2 was removed from the cultures for the last 14

days and gels maintained stability due to high cell proliferation and matrix synthesis.

5.4 Assessment of cell proliferation within the hydrogels

Chondrocyte proliferation of cells encapsulated in 2% alginate sulfate, 2%

alginate and mixtures of the two components with 2% (w/v) total concentration was

assessed using the Calbiochem BrdU proliferation kit. Proliferation of encapsulated

cells measured by this assay corroborated the visual images of Figure 5.2. The 2%

alginate sulfate hydrogels had 5-fold higher proliferation when compared to 2%

alginate (p= 0.038) and the other mixtures (0.5% alginate sulfate + 1.5% alginate

(p=0.055), 1% alginate sulfate + 1% Alg (p=0.059) and 1.5% alginate sulfate + 0.5%

alginate (p=0.11)) (Figure 5.3).

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

60

Figure 5. 3 | Cell proliferation in gels of various alginate sulfate (AlgSulf) contents measured using a

BrdU assay after 7 days of cell encapsulation. Absorbance was normalized by relative quantitation of

the absorbance value (450nm-540nm blanked) of 2% alginate (Alg) gels.

5.5 Decoupling stiffness from cell spreading

To determine if hydrogel stiffness was responsible for cell spreading, cells were

encapsulated in alginate sulfate and alginate at a similar compressive modulus and cell

spreading was qualitatively assessed (Figure 5.4A). Alginate was prepared with

increasing concentrations from 0.4% to 2% (w/v) where initial results revealed similar

compressive moduli for 2% alginate sulfate and 0.4% alginate. The compressive

modulus of 2% alginate (E=44.4 ± 3.21) was significantly higher (p= 1.2E-5) than 2%

alginate sulfate (E=2.4 ± 0.57), however 0.4% alginate had a similar compressive

modulus (E=4.01 ± 0.66) to 2% alginate sulfate (p=0.84). With regard to the cell

Chapter 5

61

morphology, no cell spreading was observed in pure alginate hydrogels (both 0.4% and

2%) while cells showed significant spreading in alginate sulfate hydrogels (Figure 5.4B).

This finding strongly suggests that the lower stiffness of the alginate sulfate hydrogels

was not the cause of the extensive cell spreading in this material.

Figure 5. 4 | Mechanical properties of alginate sulfate (AlgSulf) hydrogels and effect on cell

proliferation. A) Average compressive modulus of alginate (Alg) gels at 2% and 0.4% (w/v) and AlgSulf

2% when subjected to unconfined compressive strain and B) Transmission images of passage 3

chondrocytes encapsulated in Alg at 2% and 0.4% (w/v) and in AlgSulf at 2% (w/v) for 7 days, scale bar

(50 µm).

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

62

5.6 RhoA and integrin signalling of chondrocytes in alginate sulfate

To explore whether cell morphology in alginate sulfate hydrogels was integrin

mediated, we tested the effect of blocking beta1 integrins on cell spreading.

Encapsulation within alginate sulfate hydrogels after incubation with the beta1

blocking antibodies showed a remarkable inhibition of cell spreading (Figure 5A). We

further investigated the effect of beta1 integrin blocking on cell proliferation. As

shown in Figure 5B, anti-beta1 integrin antibody caused a statistically significant

decrease in cell proliferation compared to untreated control (p=0.002), and to IgG

treated control (p=0.014). This evidence indicates that integrin signaling plays a role in

the physical interaction of the cells with the hydrogel and the increase in cell

proliferation. Chondrocyte proliferation has been associated with RhoA mediated

signaling pathways [180]. Therefore, we also investigated whether beta1 integrin could

act through a RhoA mediated mechanism. For this purpose, cells within 2%, 0.4%

alginate and 2% alginate sulfate maintained in culture in the presence/absence of

beta1 integrin blocking antibody or non-specific IgG were assayed for the degree of

RhoA activation. Interestingly, we found that RhoA activity was significantly higher in

2% alginate sulfate compared to pure alginate at 2% (p=0.003) and at 0.4% (p=0.026)

(Figure 5C). Moreover, beta1 integrin blocking produced a significant inhibition of

RhoA activity compared to the non-treated control (p=0.017), while IgG did not have

any significant effect (p=0.374). Those findings, taken together, indicate that beta1

integrin is responsible of cell protrusion and proliferation within alginate sulfate and

could act through the RhoA GTPase pathway.

Chapter 5

63

Figure 5. 5. | Integrin mediated cell spreading and proliferation within alginate sulfate (AlgSulf)

hydrogels and RhoA activity. A) Optical images of day 7 cultures of bovine chondrocyte encapsulated

within 2% AlgSulf gels + beta1 integrin antibody (β1) or IgG B) proliferation assay (the graph bars

represent the percentage values normalized against the untreated AlgSulf 2%) and C) OD 490 nm which

corresponds to the degree of RhoA activation in day 7 cultures within the hydrogels.

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

64

Proliferation and spreading of bovine chondrocytes on collagen based films

were shown earlier to be reduced by blocking beta1 integrins [124]. Moreover,

remodeling of three-dimensional collagen fibers and formation of aggregates by

mesenchymal stem cells was shown to be dependent on beta1 integrins [181]. We

observed here that blocking beta1 integrins prevented cell spreading and reduced

proliferation, suggesting that proliferation was integrin based and possibly through cell

synthesized collagen. Furthermore, the activity of RhoA was reduced with beta1

blocking indicating that proliferation was RhoA mediated. This is in agreement with

previous findings which report RhoA as an important regulator of chondrocyte

proliferation and differentiation.

5.7 Cyclin D1 expression is upregulated in alginate sulfate samples

To investigate if the proliferation within alginate sulfate was controlled by

genes regulating the cell cycle, expression of Cyclin D1 within alginate or alginate

sulfate hydrogels was investigated (Figure 5.6). Cyclin D1 gene expression was

upregulated 2-fold in alginate sulfate compared to alginate (p=0.74), indicating that

alginate sulfate possibly enhances proliferation by upregulation of Cyclin D1.

Chapter 5

65

Figure 5. 6. | Cyclin D1 expression. Relative gene expression of Cyclin D1 for bovine chondrocytes

encapsulated in alginate sulfate (AlgSulf) relative to alginate (Alg). RPL13 was used as a reference gene.

Beier and collaborators correlated the activity of RhoA with the increased

expression of Cyclin-D1, which is involved in regulating the cell cycle G1/S transition

[180]. Here we showed that cyclin-D1 gene expression was upregulated 2-fold in

alginate sulfate compared to alginate, thus indicating that alginate sulfate enhances

proliferation by ultimately regulating expression of cell cycle genes.

5.8 Expression of cartilage markers for dedifferentiated

chondrocytes within alginate sulfate hydrogels

The expression of relevant cartilage markers showed no significant difference in

expression between alginate sulfate and alginate indicating that the chondrocytes can

proliferate without undergoing de-differentiation which is typically associated with

expansion in monolayer culture. Interestingly, the expression of the catabolic gene

MMP13 was 11-fold lower in alginate sulfate samples after 35 days in culture (Figure

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

66

5.7). A 2-way ANOVA showed that sulfation had statistically no significant effect on

Col2/Col1 (p=0.56), Sox9/RUNX2 (p=0.48), Aggrecan (p=0.067) or MMP13 (p=0.99). On

the other hand the effect of culture time was only significant for Col2/Col1 ratio

(p=0.0087) and not significant for SOX9/RUNX2 (p=0.91), Aggrecan (p=0.16) and

MMP13 (p=0.12). Cells cultured on tissue culture plastic for 21 days exhibited a 36-fold

and 38-fold downregulation in Col2 and Aggrecan respectively compared to expression

in alginate at the same time point (data not shown).

Figure 5. 7. | Expression of relevant cartilage markers. Relative gene expression of Col2/Col1,

SOX9/RUNX2, Aggrecan and MMP13 for bovine chondrocytes encapsulated in alginate (Alg) or alginate

sulfate (AlgSulf) for 21 or 35 days. The gene expression of cells encapsulated in Alg for 21 days was used

as a reference sample and RPL13 as a reference gene.

5.9 Immunohistological staining and gross appearance of alginate

sulfate hydrogels

Encapsulation of bovine chondrocytes within the alginate sulfate hydrogels

induced an increase in cell proliferation ultimately resulting in an opaque cartilage like

Chapter 5

67

appearance of the constructs after 35 days in culture (Figure 5.8). Immunohistological

staining for Col2, Col1 and proteoglycans revealed no significant difference in protein

synthesis for cells encapsulated in alginate sulfate when compared to alginate. The

overall protein content in the alginate sulfate samples was qualitatively higher and can

be attributed to the higher cell number (Figure 5.8).

Figure 5. 8. | Gross appearance and immunostaining of matrix molecules. Gross appearance of alginate

sulfate (AlgSulf, left) and alginate (Alg, right) 30 µL gels after 35 days in culture (scale bar 2 mm).

Immunostaining of Col2, Col1 and proteoglycan in AlgSulf (left) and Alg (right) samples after 35 days in

culture 40x magnification (scale bar 50 μm).

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

68

Alginate sulfate has been shown to have a high affinity to most heparin binding

growth factors [166]. Moreover, a scaffold containing alginate sulfate was shown to

improve chondrogenesis of encapsulated mesenchymal stem cells by gradual delivery

of transforming growth factor beta (TGFβ-1) [167]. Bovine chondrocytes were never

reported in combination with alginate sulfate and we hypothesized that the high

affinity to growth factors would result in a more chondrogenic environment for

encapsulated chondrocytes. However the focus of this study turned into the

unexpected strong proliferation and spreading of encapsulated chondrocytes and

whether chondrocytes would maintain their cartilage phenotype after long term

culturing. Chondrocytes have been traditionally reported to redifferentiate to a

cartilage phenotype after dedifferentiation when cultured in 3D hydrogels such as

agarose [60] or alginate [72, 161, 162]. However, cell proliferation in these systems is

limited and thus the regeneration capacity is low [182]. Collagen based hydrogels were

shown to induce cell proliferation at the expense of loss of the cartilage phenotype

[182, 183]. Cells encapsulated in alginate sulfate hydrogels exhibited a high degree of

proliferation while maintaining similar levels of cartilage markers and reducing

expression of MMP13. The effect of alginate sulfate on MMP13 expression may be

similar to the anti-inflammatory effect of chondroitin sulfate on MMPs through

interleukin-1 beta (IL-1β) [27, 41, 42]. However further investigation is needed to

confirm this hypothesis.

The large differences in proliferation rates between alginate and alginate

sulfate encouraged us to further investigate if the material was suitable for cartilage

tissue engineering applications. To test this possibility, dedifferentiated bovine

chondrocytes were encapsulated for up to 5 weeks within alginate and alginate sulfate

and the expression of relevant cartilage markers was analysed using qRT-PCR. Alginate

sulfation had no significant effect on the expression of Col2/Col1, SOX9/RUNX2, Agg

and MMP13. Similarly, at the protein level cells in both systems were not different.

Remarkably, alginate sulfate constructs attained a cartilage-like opaque appearance

after 5 weeks in culture.

Chapter 5

69

5.10 Chapter summary

In conclusion, alginate sulfate is a potential hydrogel for autologous

chondrocyte implantation that allows cells to synthesize their own matrix, proliferate

and maintain their cartilage phenotype. In addition, it has a proven high affinity to

important growth factors and can thus be loaded with relevant cartilage growth

factors for improved performance. Finally, the system may have strong action on the

inflammatory responses and thus may reduce complications resulting from increased

inflammation in the implant site. In conclusion, alginate sulfate may prove to be an

ideal biomaterial for cartilage tissue engineering.

Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers

70

Chapter 6

71

6 GFOGER Modified MMP-Sensitive

Polyethylene Glycol Hydrogels Induce

Chondrogenic Differentiation of Human

Mesenchymal Stem Cells

The work in this chapter has been submitted to Acta Biomaterialia Mhanna et al. (2013B)

In the previous chapter, we saw that functionalization of alginate with GAG

mimetic groups such as sulfates was sufficient to induce strong cell proliferation and

maintain the cartilage phenotype of 3D encapsulated chondrocytes. This was a

motivation to explore the effect of using a collagen mimetic peptide on the biological

responses of cells. In this chapter, the effects of the fibronectin adhesion sequence

RGD and the relatively novel collagen mimetic peptide GFOGER on chondrogenesis of

3D encapsulated human mesenchymal stem cells were investigated. Furthermore, the

effects of presenting the peptides in a degradable or non-degradable

microenvironment were analyzed.

6.1 Chondrogenic differentiation and the microenvironment

The limited ability of cartilage tissue to self-repair after injury or disease has

driven intensive efforts to engineer replacement tissue [12, 156]. Cell-based cartilage

tissue engineering holds strong promise to provide living tissue for cartilage repair [14,

49, 52]. Chondrogenic differentiation of stem cells has been widely studied using pellet

cultures [67, 184, 185], micromass cultures [186] and in hydrogels [187]. Both natural

[187-191] and synthetic [192-194] materials have been employed to induce

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

72

chondrogenesis of stem cells in 3D. Synthetic gels such as poly(ethylene glycol) (PEG)

are attractive for 3D culturing of cells [74], as they can be functionalized with specific

adhesion sequences [193], growth factor binding sites [195, 196] and protease-

sensitive motifs [197]. A PEG hydrogel modified with a matrix metalloproteinase

(MMP)-sensitive peptide was shown to increase chondrocyte proliferation and

upregulate mRNA levels of type II collagen and aggrecan in bovine chondrocytes [73].

The extracellular microenvironment is believed to play a key role in the

biological responses of stem cells [198, 199]. Modification of hydrogels with small

functional groups such as amines, phosphates and others were shown to induce

differentiation of stem cells into the osteogenic or adipogenic pathways [200].

Furthermore, adhesion sequences such as the RGD sequence has been shown to be

involved in adhesion [197], viability [201], proliferation [202] and differentiation of

stem cells [203, 204] however, an inhibition of chondrogenesis was earlier reported in

RGD-modified alginate hydrogels [96]. Chondrocytes form primary interactions with

collagen 2 and hyaluronic acid in the native tissue [13]. Recently Knight et al. [98]

identified the GFOGER sequence as the recognition site in type I collagen and type IV

collagen. Aggregation of chondrocytes in suspension was also shown to be mediated

by the GFOGER sequence which is recognized by α10β1 integrins [19]. Moreover, type II

collagen induced GAG deposition in chondrocytes was inhibited by GFOGER integrin

blocking peptides [205].

The GFOGER sequence has been used for modification of surfaces [206-208]

and hydrogels [205, 209] to control adhesion and biological responses of chondrocytes

and stem cells. When bovine bone marrow stromal cells were encapsulated in an

agarose gel with covalently coupled GFOGER sequences, type II collagen and aggrecan

mRNA stimulation as well as GAG deposition by chondrogenic factors was inhibited

[209]. However, human mesenchymal stem cells (hMSC) encapsulated in a soft PEG

hydrogel modified with the triple helical GFOGER peptide showed enhanced

chondrogenesis compared to the unmodified and the stiffer modified hydrogels [210].

Chapter 6

73

We hypothesized that GFOGER functionalized MMP-sensitive PEG gels would

enhance the chondrogenic differentiation of encapsulated hMSC stem cells, and that

RGD and GFOGER peptides would induce differential biological responses on

encapsulated cells. To test this hypothesis, hMSC were encapsulated in degradable

(MMP-sensitive) or non-degradable PEG hydrogels containing RGD or GFOGER

sequences. Initial hydrogel compressive modulus, cell morphology, proliferation, gene

expression and synthesis of chondrogenic markers were analysed.

6.2 Hydrogel modification did not affect mechanical properties

Degradable and non-degradable hydrogels modified with 100 µM RGD, 100 µM

GFOGER or non-modified were allowed to swell in cell culture medium in the incubator

for 24 hours after production. Compressive moduli of the equilibrated hydrogels were

measured using a texture analyzer (TA.XTPlus, Stable Microsystems, UK). This step was

essential to prove that differential biological responses from the peptide conjugations

were not due to changes in the materials’ compressive moduli.

The compressive moduli of the hydrogels were not significantly different among

the different tested conditions (Figure 6.1). A two-way ANOVA test revealed that there

was no significant influence on the mechanical properties, neither due to the

degradability of the gels (p=0.35) nor to the added adhesion peptide (p=0.32). The

interaction of degradability and peptide parameters also showed no significance

(p=0.75). We can therefore assume that cell behavior was not affected by the initial

mechanical properties of the gels. Furthermore, the average compressive modulus of

the hydrogels was 3.7±0.3 kPa which is similar in magnitude to the values stated by the

manufacturer in the materials certificate of analysis.

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

74

Figure 6. 1. | Compressive moduli of the hydrogels. Compression test of hydrogels containing 100 µM

RGD, 100 µM GFOGER or without adhesion peptides (No peptide). Young's moduli are shown for the

non-degradable gels (Non-deg) as well as for the degradable gels (Deg).

6.3 Cell viability

The viability of MSCs encapsulated in the PEG hydrogels was affected by gel

degradability as well as peptide incorporation (Figure 6.2). Cell viability was higher in

degradable PEG hydrogels compared to non-degradable. Cells had a higher viability in

gels modified with the adhesion sequences, though there was not a major difference

between gels containing the RGD sequence versus those containing the GFOGER

sequence.

Chapter 6

75

Figure 6. 2. | Cell viability in the hydrogels. Merged live/dead assay of stem cells encapsulated for 21

days in non-degradable (Non-Deg) and degradable (Deg) PEG gels. The image shows live (green) and

dead (red) cells simultaneously. Images were taken with a 10x objective, scale bar 100 µm.

6.4 Cell morphology

MSCs in PEG hydrogels exhibited very different actin cytoskeletal architecture

depending on the specific integrin binding domain incorporated into the PEG matrix. If

no peptide adhesion sequence was used, cells remained round in the maximum

intensity projections of the Z-stacks (Figure 2A, D and G). In RGD and GFOGER

containing PEG gels, cells spread and developed stress fibers (Figure 2B, C, E, F, H and

I). This effect was much more pronounced in the degradable gels that permitted the

cells to remodel their matrix and physically create space in the 3D network for this

elongated morphology (Figure 2E, F, H and I). Cell protrusions were observed in non-

degradable peptide containing gels, though spreading was clearly restricted. The

elongated shape of spread cells varied strikingly in RGD-PEG versus GFOGER-PEG gels.

In RGD-PEG gels, the cells tended to be smaller and contained thin star-like projections

of cytoskeleton ending at a matrix attachment site in a sharp point (Figure 2B, E and

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

76

H). In contrast, in GFOGER-PEG gels, cells interacted with the gel using more uniform,

evenly distributed matrix attachment sites (Figure 2C, F and I).

Figure 6. 3. | hMSCs cultured for 21 days in chondrogenic media in modified PEG gels prior to fixation and staining for nuclei (DAPI, blue) and actin filaments (phalloidin, red). PEG modifications included an MMP-cleavable moiety (degradable gels, D-I) or no degradation peptides (non-degradable, A-C) as well as incorporation of the adhesion peptides RGD (B, E and H), GFOGER (C, F and I) or no adhesion peptide (A, D and G). Gel degradability improved cell proliferation while incorporation of adhesion peptides permitted substantial cell spreading. Cells in RGD-containing gels displayed star-like morphologies with thick, individually-discernible actin fibers and attachment to the matrix at the ends of stress fibers. GFOGER-modified gels induced a more homogenous spreading of cells with thinner and more dispersed actin filaments (G-I). Depicted are maximum intensity projections through 150 μm thick Z-stack images, scale = 50 μm.

Chapter 6

77

6.5 Cell proliferation was highest in GFOGER degradable hydrogels

Peptide incorporation in the gels yielded an increase in cell proliferation for

both degradable and non-degradable gels (Figure 6.4). GFOGER-modified degradable

gels yielded the best proliferation rate and were higher than both no peptide gels

(p=0.01656) and RGD-modified gels (p=0.06933). Degradability of GFOGER-modified

gels was shown to cause a significant increase (p=0.01378) in cell proliferation whereas

such effect was not seen among RGD-incorporated and no peptide gels.

Figure 6. 4. | Assessment of cell proliferaiton. Cell proliferation in gels was assessed after 7 days with a

BrdU assay. Absorbance values (450-540 nm) were normalized to non-degradable gels with no peptide

incorporation and relative quantitation was shown.

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

78

6.6 Gene expression

Gene expression of type II collagen and aggrecan were highest in GFOGER-

modified degradable gels after 7 and 21 days (Figure 6.5). Gene expression of type II

collagen and aggrecan changed significantly with time (p=0.016) and (p=0.026)

respectively. After 21 days in culture only GFOGER and RGD modified degradable

hydrogels exhibited significantly higher levels of type II collagen when compared to

pellets cultured in non-chondrogenic medium (p=0.008) and (p=0.0324) respectively.

Moreover, expression of type II collagen was higher in GFOGER-modified degradable

gels compared to non-modified degradable gels (p=0.038). On the other hand, only

GFOGER-modified degradable gels had a significantly higher level of aggrecan when

compared to pellets in non-chondrogenic medium (p=0.018).

Chapter 6

79

Figure 6. 5. | Gene expression of relevant cartilage markers. Gene expression of type II collagen and

aggrecan were determined using quantitative real time PCR (StepOnePlus, Applied Biosystems). Non-

degradable hydrogels without modification (0 µM Deg) cultured in control media (control) were used as

reference sample and RPL 13 was used as a reference gene.

6.7 GAG production in peptide-modified gels

Peptide incorporation increased GAG production in degradable gels after 7 days

and 21 days and GFOGER-modified degradable hydrogels had on average the highest

GAG content (Figure 6.6). After 7 days GFOGER-modified degradable gels had an

almost significant increase in GAG/DNA (p=0.059). Non-degradable gels led to similar

production of GAG with or without peptide at both time points.

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

80

Figure 6. 6. | GAG/DNA quantification in the various conditions. DNA was quantified using picogreen

and GAG was quantified using the DMMB assay. GAG quantity was normalized to the DNA amount.

6.8 Histology and immunostaining

GAG synthesis was confirmed qualitatively using Alcian blue staining. GFOGER

degradable hydrogels showed the most prominent staining compared to all other

hydrogel conditions (Figure 6.7). Pellets cultured in chondrogenic medium (Figure 6.7

A) showed stronger GAG staining compared to cells cultured in control medium (Figure

6E). In hydrogels with no adhesion peptide modification, staining was limited and only

visible around the cells (Figure 6 B and F). In RGD modified gels, staining was similar to

the unmodified gels (Figure 6 C and G) whereas in GFOGER modified gels staining was

very strong around the cells in the non-degradable hydrogels with little GAGs in the

matrix (Figure 6 D) and was distributed strongly all over the sample in degradable

hydrogels (Figure 6 H). Alcian blue staining in chondrogenic pellets (Figure 6E) was

comparable to degradable hydrogels modified with the GFOGER peptide although the

cell concentration in pellets is clearly higher.

Chapter 6

81

Figure 6. 7. | Alcian blue staining of hMSC cultured in pellets and in PEG gels of different

compositions. hMSC cultured in pellets with control medium (A), chondrogenic medium (E), non-

degradable PEG gels without adhesion peptides (No peptide - Non Deg, B), non-degradable PEG gels

with RGD peptides (RGD – Non Deg, C), non-degradable PEG gels with GFOGER peptides (GFOGER – Non

Deg, D), degradable PEG gels without adhesion peptides (No peptide – Deg, F), degradable with RGD

peptides (RGD-Deg, G) and degradable with GFOGER peptides (GFOGER – Deg, H). All hydrogels were

cultured for 21 days in chondrogenic medium. Images were acquired with a 40X objective, scale bar 50

μm.

Type II collagen immunostaining varied considerably among samples and was

strongest in the degradable hydrogels (Figure 7). Staining in pellets under

chondrogenic medium (Figure 7 E) was limited to a few cells which showed very weak

staining, which was not different than cells in pellets cultured under control medium

(Figure 7 A). Within hydrogels, staining was in general clearer and stronger in

degradable gels (Figure 7 F-H) compared to non-degradable gels (Figure 7 B-D).

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

82

Figure 6. 8. | Type II collagen immunostaining of hMSC cultured in pellets and in PEG gels of different

compositions. hMSC cultured in pellets with control medium (A), chondrogenic medium (E), non-

degradable PEG gels without adhesion peptides (No peptide - Non Deg, B), non-degradable PEG gels

with RGD peptides (RGD – Non Deg, C), non-degradable PEG gels with GFOGER peptides (GFOGER – Non

Deg, D), degradable PEG gels without adhesion peptides (No peptide – Deg, F), degradable with RGD

peptides (RGD-Deg, G) and degradable with GFOGER peptides (GFOGER – Deg, H). All hydrogels were

cultured for 21 days in chondrogenic medium. Images were acquired with a 40X objective, scale bar 50

μm.

A major requirement for an engineered 3D biomaterial to be used for tissue

engineering applications is to maintain cell viability for a long culturing period.

Additionally, increased proliferation may be desirable for tissue formation [211].

Although the viability of fibroblasts [197] or chondrocytes [73] is not significantly

affected by the materials degradability, stem cells have been reported to have a higher

viability in degradable hydrogels [211]. Furthermore incorporation of adhesion

sequences like RGD [212], IKVAV [211] and GFOGER [210] were shown to improve

stem cells viability in PEG gels. We show here that the presence of MMP motifs alone

Chapter 6

83

did not have a significant effect on cell proliferation or cell viability. However, the

incorporation of adhesion peptides in both degradable and non-degradable hydrogels

improved cell viability and cell proliferation which is consistent with the literature

[210-212]. In general, incorporation of peptide adhesion sequences improved

proliferation however only GFOGER degradable hydrogels had a significantly higher

proliferation (p=0.017) compared to the degradable hydrogels without any adhesion

peptides. Degradability was also only a statistically significant factor in the presence of

GFOGER sequences (p=0.014).

A round cell morphology is often linked to a cartilage phenotype [60, 213].

Chondrogenesis of bone marrow stem cells has been shown to be inhibited in RGD

modified alginate [96]. Moreover, chondrogenesis in agarose was inhibited in the

presence of RGD sequences and this inhibition was blocked by disruption of the f-actin

cytoskeleton using cytochalasin D [97]. These results are in contradiction with the

findings of the current study where chondrogenic differentiation was strongest in RGD

and GFOGER modified degradable hydrogels in which cells had a clearly more spread

morphology than the other conditions. In general the trend observed in the current

study was that the increase in proliferation was associated with increase in type II

collagen. These results indicate that cell spreading and proliferation do not necessarily

inhibit the onset of chondrogenesis and that the microenvironment and growth factors

may be stronger players. Moreover, cells encapsulated in RGD modified degradable

hydrogels had a star-like morphology with strong actin fibers whereas cells in GFOGER

had a uniaxial spreading with more diffuse and less intense actin fibers which might

better support chondrogenesis.

Several studies have employed RGD sequences for improving cell adhesion

[197], viability [201] and chondrogenesis [203, 204] of stem cells in 3D PEG gels.

However, very few studies investigated the incorporation of the GFOGER peptide in 3D

hydrogels and their influence on chondrogenesis [209, 210]. In particular the effect of

hydrogel degradability on the response of stem cells to the GFOGER peptide was not

GFOGER Modified MMP-Sensitive Polyethylene Glycol Hydrogels Induce Chondrogenic Differentiation of Human Mesenchymal Stem Cells

84

reported. Furthermore, a study comparing the effect of presenting equal amounts of

the GFOGER and RGD peptides on chondrogenesis is presented here for the first time.

We observed that the presence of MMP motifs in the PEG hydrogels greatly influenced

the cell response to GFOGER and RGD peptides in terms of cell spreading and

proliferation. Furthermore, degradability was essential to induce chondrogenesis in

peptide modified hydrogels. Finally, it was observed that when degradable PEG gels

were modified with adhesion sequences chondrogenesis was better than in pellet

cultures.

6.9 Chapter summary

GFOGER modified hydrogels represent a new biomimetic option for cell growth

in a 3D hydrogel. The GFOGER peptide induces higher proliferation of stem cells and

better maintains a chondrogenic phenotype compared to RGD in terms of expression

of the chondrogenic markers type II collagen and aggrecan and alcian blue staining.

Degradability and the presence of adhesion peptides in 3D cultures are shown here to

be necessary to maintain a high viability and initiate chondrogenesis of MSCs. Taken

together, the results of the current study suggest that GFOGER peptides presented in a

degradable hydrogel provide a suitable microenvironment for proliferation and

chondrogenesis of hMSCs.

Chapter 7

85

7 Probing the Microenvironmental Conditions

for Induction of Superficial Zone Protein

Expression

The work in this chapter is in revision in Osteoarthritis and Cartilage, Mhanna et al. (2013C).

In the previous chapters we demonstrated the importance of the

microenvironment on chondrocytes and stem cells. We showed that small sulfate

groups and peptides had profound effects on cell proliferation and expression of

chondrogenic markers. In the current chapter we investigate the effect of cell culture

dimensionality, mechanical stimulation, oxygen tension and cell morphology on the

expression of superficial zone protein and type II collagen.

7.1 Chondrocyte dedifferentiation and superficial zone protein

The development of replacement tissue for healing cartilage injuries poses a

major scientific challenge. The avascular nature, limited proliferation and restricted

access to nutrients and regenerative cells prevent the spontaneous regeneration of

cartilage defects [12, 214]. Furthermore, the dense matrix that surrounds cartilage

cells (chondrocytes) inhibits their migration to the defect site [14, 156]. Articular

cartilage is organized into the deep, middle and superficial zones [13, 215]. The

superficial zone in cartilage is believed to be crucial for long term stability and function

of the tissue by protecting the deeper layers and providing lubricants that limit friction

and prevent wear [50]. Superficial zone protein (SZP) which is highly homologous to

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

86

lubricin is mainly produced by the chondrocytes of the superficial zone and is a key

molecule involved in cartilage lubrication [17, 216, 217].

Chondrocytes have been shown to dedifferentiate when cultured on two-

dimensional (2D) tissue culture plastic where they assume a fibroblastic phenotype

[59, 61, 159]. Culturing chondrocytes in alginate [72, 161, 162] (Alg) or agarose [60]

hydrogels has been shown to maintain the cartilage phenotype. However, most studies

address the re-expression of type II collagen (Col2) and aggrecan which are present

throughout cartilage tissue but are synthesized in lower amounts by superficial zone

cells compared to cells of the deeper layers [13]. Only a few studies have investigated

the effects of dedifferentiation on the superficial zone phenotype and particularly the

expression of SZP [61].

The zones of articular cartilage are important to the function of the tissue,

making an engineered cell-based cartilage replacement that recapitulates the stratified

cartilage structure desirable. In the current chapter we focus on microenvironmental

factors which regulate the expression of superficial zone protein. Synthesis of SZP has

been shown to be affected by growth factors such as bone morphogenic protein-7

(BMP-7) [218] and transforming growth factor beta (TGFβ) [219] as well as cytokines

such as interleukin-1 beta (IL-1β) [220]. Gene expression and synthesis of superficial

zone protein is also modulated by mechanical stimulation such as shear [118], surface

motion [116] and mechanical stress [117]. Finally, differentiation under hypoxic

conditions stimulated matrix production by middle/deep chondrocytes and

proteoglycan 4 by superficial zone chondrocytes [95].

We hypothesized that an in-vitro system that most closely resembles the in-vivo

cartilage superficial zone environment would allow dedifferentiated chondrocytes to

re-express markers of the superficial zone. To test this hypothesis we designed

experiments to address the effect of 3D cultures, cyclic tensile strain, oxygen tension

and cell morphology on the expression of superficial zone protein. We designed a

mechanical loading chamber for application of homogeneous 3D strain compatible

with the commercially-available STREX machine. We then cultured chondrocytes on 2D

Chapter 7

87

substrates and 3D hydrogels exhibiting varying degrees of spreading in the presence

and absence of cyclic mechanical strain and quantified the expression of SZP. We also

studied the expression of SZP at ambient oxygen levels as well as under hypoxic

conditions and compared its expression profile to that of Col2.

7.2 Both SZP and Col2 undergo dedifferentiation during serial

passaging

Serially passaged chondrocytes exhibited a massive downregulation in Col2

expression (Figure 7.1B) similar to what has been reported in previous studies. The

downregulation was significant in cartilage tissue, p0 and P1 when compared to

passage 4. The expression of SZP decreased with passaging, exhibiting a similar trend

to that reported for Col2 (Figure 7.1A). The relative quantity decreased from 78 fold in

cartilage tissue to 32 fold at P0 d4 normalized to P4 d4. The downregulation was

significant between cartilage tissue and P0 compared to passage 4, but not between

passage 2 and 4.

Figure 7. 1. | Effect of chondrocyte dedifferentiation on gene expression of SZP and type II collagen.

Chondrocytes were harvested from cartilage tissue and serially passaged by culturing on tissue culture

plastic for 4 days from passage 0 to passage 4. Analysis of SZP mRNA levels (A) and Col2 (B) showed a

decrease in expression with passaging. Gene expression was normalized to expression in P4, (n=4), *

p<0.05 and *** p<0.001.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

88

Chondrocyte dedifferentiation is a common phenomenon observed in 2D

culture and we show here that the reduction in SZP expression along with the

precipitous drop in Col2 expression can be now considered a part of the de-

differentiation phenomenon associated with monolayer culture.

7.3 Redifferentiation of serially passaged primary chondrocytes

Chondrocytes encapsulated in alginate beads for 4 days re-expressed SZP but

not Col2 (Figure 7.2). The expression of SZP was upregulated 52.3 fold (p=0.00013) in

alginate beads compared to chondrocytes cultured on TCP (Figure 7.2A). This is

comparable to the expression in cartilage tissue indicating that the cells could almost

completely restore the expression of SZP by 3D culture. The expression of Col2 was

however downregulated 2.7 fold after 4 days in alginate culture (Figure 7.2A). The

downregulation of Col2 expression in alginate beads although statistically insignificant

(p=0.08) is in contradiction to the finding of several studies which show that

chondrocytes re-express col2 in alginate beads [72, 161, 162]. In order to confirm the

expression of SZP at the protein level, immunostaining for SZP was performed on

alginate beads and TCP samples. Strong staining for SZP was found in alginate

encapsulated cells but not in cells cultured on TCP (Figure 7.2B).

Chapter 7

89

Figure 7. 2. | Redifferentiation of serially passaged chondrocytes. Passage 3 chondrocytes were

cultured on tissue culture plastic (TCP) for 4 days or encapsulated in alginate beads (Beads) for the same

period to determine if expression of SZP and Col2 could be restored. A) qRT-PCR mRNA levels for SZP

and Col2 on TCP and within alginate beads normalized to TCP gene expression (n=5) and B)

immunostaining of SZP in cells cultured on TCP and in alginate beads, scale bar 50 μm.

Chondrocytes have been traditionally reported to redifferentiate to a more

chondrogenic phenotype when cultured in 3D hydrogels [62]. The mechanism by which

3D cultures allow redifferentiation of chondrocytes is not well understood but involves

the integrity of the cytoskeleton. The prevention of cell spreading in 2D cultures by the

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

90

ROCK kinase inhibitor Y-27632 induced a cortical actin morphology and was shown to

redifferentiate chondrocytes. Likewise, interference with actin polymerization by

cytochalasin D or jasplakinolide helped restore the chondrogenic phenotype in

dedifferentiated cells [70, 71]. The addition of concanavalin A to chondrocyte cultures

also caused cell rounding and increased proteoglycan synthesis [77]. Dedifferentiation

and redifferentiation studies have often focused on the expression of Col2, Col1 and

aggrecan while little attention has been given to the behavior of the main superficial

zone marker, SZP. We observed that SZP is downregulated with passaging similar to

Col2 and its expression can be restored within 4 days of encapsulation in alginate

beads unlike type II collagen.

7.4 Design and evaluation of a 3D tension and compression

chamber compatible with the STREX strain machine

The chamber was designed using Key Creator v7.02 (Figure 7.3A) and metal

molds were produced with a milling machine (Deckel FP 3 CC, Germany). The molds

were used for casting the 3D PDMS chamber (Figure 7.3B).

Chapter 7

91

Figure 7. 3. | Design for the 3D molds A and the final 3D chamber B.

The STREX machine allows for different magnitudes of strain to be applied to a

given chamber. For each tension/compression setting available in the STREX device,

the corresponding measured strain in the strained hydrogel was measured (Table 7.1).

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

92

Table 7. 1. | Actual strains measured in the alginate gel and the corresponding strain setting in the

STREX device.

7.5 Application of cyclic mechanical strain to 2D and 3D constructs

The strain within the hydrogel was found to be fairly uniform across the width

and length of the specimen (Figure 7.4B). Application of 10% tensile strain in the STREX

machine (Figure 7.4A) resulted in a tensile strain of 7.5 ± 0.6% while the same

compressive strain resulted in 8 ± 0.5% within the hydrogel (Figure 7.4C). To determine

the homogeneity of the strain field, the displacement of glass beads (Figure 7.4D) was

measured before and after application of 10% compressive/tensile strains in areas

designated in Figure 7.4E. The difference in displacement of glass beads among these

different locations was found to be non-significant (p>0.05, Figure 7.4F). The area

between the two small round pillars (marked with arrows, Figure 7.4E) was used

thereafter for analysis of strain distribution.

Chapter 7

93

Figure 7. 4. | Design for application of 3D cyclic strain. A) STREX Device. B) New chamber designed for

3D loading. Metal tubes were inserted in the holes connecting the chamber to the machine to minimize

non-uniform deformations around the holes. C) Strain measured within a hydrogel in response to 10%

compressive/tensile strain set by the instrument. D) Glass beads embedded in an alginate gel (scale bar

200 μm). E) Schematic representation of the chamber: The gel was injected in the blue area, letters (a, b

and c) mark areas where images of glass beads were taken. Three images were taken 1 mm away from

the small round pillar marked with an arrow (a), 3 images at the same level closer to the wall (b) and 3

images were taken in the center middle (c) to cover the whole gel in one plane. To assess variation in

the z-axis, images were taken at 2 different depths resulting in a total of 18 images per gel. In each

image 4 measurements of displacement were performed. F) The homogeneity was assessed by

measuring the displacement before/after compression/tension of 10-30 μm glass beads. Displacement

was found to have no significant difference between the various locations. (n=3, p>0.05).

The STREX device has been previously employed for application of mechanical

strain in 2D; however limited studies have employed the system for 3D strain [221].

These studies did not verify uniformity of strains within the strained hydrogel and in

our experience could result in non-homogeneous strains throughout the gel.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

94

Furthermore, the previously reported chambers make it difficult if not impossible to

apply strain to alginate gels without adhesion to PDMS. The 3D pillared chamber

designed for this study was shown to allow application of homogenous strain to

alginate. The strain within the gel appeared to be lower than the strain applied to the

chamber where a 20% applied strain corresponded to approximately 14% strain in the

hydrogel. Inhomogeneity in strain was partially solved by inserting metal tubes that fit

tightly in the connection points between the chamber and the device. Additionally, the

alginate beam was cast in only the middle third of the chamber which gave a uniform

strain across the specimen width. In addition to the homogeneous 3D strain

distribution, the system allows a direct comparison between application of cyclic strain

to cells in 2D and cells in 3D. The current system is an excellent tool for application of

homogeneous 3D strain to cells embedded in a hydrogel and may be used for several

biomedical applications including but not limited to tendon, muscle and bone tissue

engineering.

7.6 Effect of mechanical strain on SZP expression

Gene expression of superficial zone protein (SZP) was upregulated in 2D and 3D

cultures by mechanical strain of 2h/day (Figure 7.5). A single application of mechanical

strain at day 4 of culture resulted in 1.5 fold upregulation in SZP while continuous

strain for 2h/day for 4 days resulted in a 2.4 fold upregulation in 3D cultures (Figure

7.5B). Cells seeded on collagen coated PDMS did not show a response to the single

strain treatment while a strain of 2h/day resulted in a 1.7 fold upregulation (Figure

7.5A). Although no statistical significant differences were obtained between group

means as calculated by two-way ANOVA (p=0.31), the data indicate a clear

upregulation trend especially with application of 2h/day strain.

Chapter 7

95

Figure 7. 5. | The effect of mechanical strain on SZP mRNA levels was measured by qRT-PCR. A)

Response of cells seeded on collagen coated PDMS to 2h strain applied 4 days post seeding or applied

every day for 2 h throughout the culture period (total of 4 times). B) Response of cells to the same strain

regime when encapsulated in alginate gels (n=4). Gene expression was normalized to unstrained

samples.

Physiologic mechanical loading has been shown to stimulate gene expression

and synthesis of chondrogenic markers [93, 99, 106, 222, 223]. The mechanism

through which chondrocytes sense mechanical load is not completely understood but

may involve stretch-activated ion channels and integrin signalling [114]. External

signals sensed by cell membrane proteins are transduced through molecular pathways

leading to the expression of certain genes [94]. To further augment the expression of

SZP, we used the commercially available STREX device to apply cyclic mechanical strain

to chondrocytes cultured in 2D and 3D [108, 221]. Tensile strains are expected to be

highest in the superficial zone and thus were expected to induce expression of

superficial zone markers [22]. We showed using this system that mechanical strain

upregulated SZP expression up to 2.4 fold in 3D and 1.7 fold in 2D. These values are

similar in magnitude to previously reported responses to mechanical stimulation [116,

117].

7.7 Effect of oxygen tension on expression of SZP

Under hypoxic conditions (1% pO2) Col 2 was upregulated while SZP was

downregulated. Cells encapsulated in alginate beads exhibited a 49.5 fold upregulation

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

96

in Col2 gene expression (p=0.014) under hypoxic conditions compared to normoxia

(Figure 7.6A). The expression of SZP on the other hand was downregulated 3 fold in

hypoxia (p=0.15). This observation is in accordance with the oxygen distribution in

cartilage tissue where the superficial zone oxygen levels are estimated to be between

7-10% (Figure 7.6B and C). In the deep and middle zones, where oxygen levels are less

than 1 %, SZP is not present and Col2 is highly abundant (Figures 7.6B and C). The

expression of SZP was also downregulated 7.7 fold in TCP cultures under hypoxia

(p=0.055), however, Col2 levels did not vary (RQ=1.1 fold, P=0.87). This indicates that

SZP is responsive to oxygen levels in 2D but Col2 is not.

Chapter 7

97

Figure 7. 6. | The effect of oxygen tension on SZP and Col2 mRNA levels was quantified by qRT-PCR. A) mRNA levels of SZP and Col2 for cells seeded on TCP or encapsulated in alginate beads in normoxic or hypoxic conditions normalized to expression in TCP normoxic conditions (n=3). B) Immunostaining of SZP and Col2 in cartilage tissue 40X magnification (scale bar 50 μm). C) Schematic representing cartilage tissue and its different layers as well as the gradients of oxygen and tensile strain and the corresponding SZP and Col2 expression.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

98

The avascular nature of cartilage leads to low oxygen levels compared to

vascularized tissues. Several studies have investigated the influence of oxygen tension

on chondrocyte metabolic and catabolic activity [21, 64, 95, 224-226]. These studies

showed an increased Col2 and aggrecan expression [64, 95, 224, 226] and improved

GAG deposition [225] in hypoxic compared to normoxic conditions. The effects of

oxygen tension on the expression of superficial zone markers are less investigated.

Schrobback et al. [95] found increased SZP mRNA levels in hypoxic compared to

normoxic conditions. However, this study used 5% oxygen as a hypoxic condition and

20% for normoxic conditions. The oxygen level in cartilage is believed to be between 7-

10% at the surface and near 0.1% adjacent to the subchondral bone [21, 95].

Therefore, using 5% oxygen as a hypoxia model might not induce different responses

in SZP expression. In our study 1% oxygen level was used as a hypoxic environment.

We hypothesized that simulation of the physiological cartilage conditions would direct

in-vitro cultured chondrocytes to express zonal cartilage phenotypes. This hypothesis

was confirmed by the results which showed that normoxic conditions were required

for SZP expression while hypoxic conditions induced Col2 re-expression and

downregulated SZP expression. The expression of Col2 did not increase during the 4

days of alginate bead culture and partial restoration of the expression could only be

achieved under hypoxic conditions. This is an interesting observation which indicates

that 3D culture is not sufficient for re-expression of the cartilage phenotype. Although

the common consensus is that encapsulation in alginate allows recovery of Col2

expression [72, 161, 162], some studies have shown this is not always the case [61].

7.8 Cell morphology and the expression of SZP

To address the interdependence of cell morphology and expression of SZP we

performed pilot studies to induce cell morphologies from round to spread and

quantified SZP mRNA levels. Chondrocytes cultured on the various substrates induced

a range of different morphologies, from highly spread on the 2D substrates to less

Chapter 7

99

spread in the 3D samples. Culturing chondrocytes on TCP and Col1 2D resulted in a

spread morphology typical of dedifferentiated chondrocytes. Chondrocytes exhibited a

slightly spread morphology in the Col1 hydrogel (Col1 3D) and maintained a

completely round morphology in the alginate gel (Figure 7.7).

Figure 7. 7. |Morphology of chondrocytes cultured in 2D and 3D substrates. Phase contrast (top) and

confocal images of phalloidin-rhodamine stained (bottom) passage 2 Bovine chondrocytes cultured on

TCP, Col1 coated PDMS (Col1 2D), Col1 3D and 0.8% Alg. All images were taken 3 days after cell seeding

using a 40x objective.

Gene expression of SZP was higher in 3D than 2D cultures (Figure 7.8). Cells

within alginate maintained the highest expression indicating that the spread

morphology in 3D does not necessarily lead to SZP upregulation.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

100

Figure 7. 8. | Effect of cell morphology and dimensionality on SZP gene expression. TCP was set as the

reference sample and RPL13 was used as a housekeeping gene.

Chondrocytes of the superficial zone have a more flat morphology compared to

those of deeper layers which suggest that these cells might show a better response to

culture conditions that induce a flat morphology such as 2D cultures. It has been

reported that passage 0 cells have a higher expression of SZP in 2D compared to 3D

which is in contradiction with our results [227]. The use of passage 0 cells in the

previous report, however, may explain their results as freshly isolated cells are

weakened by the enzymatic digestion process and may be less active when embedded

in a 3D matrix. We have also observed in preliminary studies that a spread

morphology in 3D within a collagen gel resulted in lower expression of SZP compared

to alginate which induces a completely round morphology (Figures 7.7 and 7.8).

However the different responses in alginate and collagen cultures may reflect distinct

material and biological properties rather the effect of spreading. To address this issue,

we also compared SZP expression of chondrocytes in RGD modified alginate compared

Chapter 7

101

to non-modified alginate. Cells were encapsulated in alginate and RGD modified

alginate (Novatach, Novamatrix) using the CaCO3-GDL method at 0.4% gel

concentration and casted in the 3D chambers. Mechanical stimulation was applied at

day 4 for 2 h. Cell morphology was only affected on the edge of the RGD modified gel

where cells were able to spread (Figure 7.9B). This is consistent with previous reports

that show RGD-modified alginate induces cell spreading only when alginate is made

degradable through conjugation of MMP sensitive motifs [228]. The presence of RGD

sequences did not induce any effect on SZP expression nor did it induce a stronger

response to mechanical stimulation. This indicates that the presence of such adhesion

sequences do not trigger any pathways that affect SZP (Figure 7.9A).

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

102

Figure 7. 9. | Effect of RGD sequences on morphology and SZP gene expression. A) mRNA levels of SZP

quantified by qRT-PCR for cells encapsulated in alginate and RGD-modified alginate under static and

mechanical strain conditions. B) Morphology of chondrocytes encapsulated in alginate and RGD-

modified alginate after 4 days of culture. Cell spreading was only observed in RGD-modified alginate at

the edges of the gel but not in the middle.

When PEG hydrogels were used in the same context we observed that

modification with RGD peptides lead to cell spreading in MMP sensitive PEG hydrogels

(QGel, Lausanne, Switzerland). However the higher cell spreading was associated with

lower expression of SZP (Figure 7.10). This indicates that a spread cell morphology in

Chapter 7

103

3D does not alone cause an upregulation of SZP. A more in depth analysis of the

biological pathways would possibly shed light on the regulation of SZP and its relation

to cell morphology.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

104

Figure 7. 10. | Morphology and gene expression of SZP for chondrocytes cultured in 3D MMP sensitive PEG hydrogels with and without RGD peptides. A) mRNA levels of SZP quantified by qRT-PCR for cells encapsulated in non-modified PEG, PEG modified with 70 µM RGD and PEG with 155 µM RGD normalized to TCP, RPL13 was used as a reference gene. B) Phalloidin-rhodamine stained (top) and transmission (bottom) images of passage 3 bovine chondrocytes cultured in non-modified PEG, PEG modified with 70 µM RGD and PEG with 155 µM RGD. All images were taken 4 days after cell seeding using a 63x objective, scale bar 20 µm.

Chapter 7

105

7.9 Chapter summary

Cartilage lubrication is crucial for maintaining the function of articular cartilage

and preventing its degeneration. Understanding the mechanisms that regulate the

expression of SZP will allow the design of engineered tissue with optimal lubrication

properties. We have shown that SZP expression in dedifferentiated chondrocytes can

be upregulated to cartilage tissue levels when cultured in alginate at normoxic

conditions. Moreover, we designed a chamber compatible with the STREX device that

allows application of homogeneous strain in 3D to any hydrogel. The results of the

current study provide insights into the regulation of SZP and tools to control its

expression in-vitro.

Probing the Microenvironmental Conditions for Induction of Superficial Zone Protein Expression

106

Chapter 8

107

8 Conclusions and outlook

The aim of this thesis was to engineer 2D and 3D microenvironments to

improve the culturing conditions of chondrocytes and stem cells used for cartilage

tissue engineering applications such as ACI. Using the layer-by-layer technique 2D

substrates made of natural ECM macromolecules were constructed and characterized

on stretchable PDMS. The films made of type I collagen and the chondroprotective

molecule chondroitin sulfate were stable in media. However, chondrocytes seeded

those films still adopted a fibroblastic phenotype and mechanical stimulation did not

help recovering the cartilage phenotype. Furthermore, the CS containing films did not

have any effect on expression of MMPs. This might have been a consequence of using

juvenile chondrocytes as in our studies cells were obtained from 6-month old

chondrocytes. It may be that the beneficial effects of chondroitin sulfate are more

pronounced when using cells from older donors or osteoarthritic cells. The well-

characterized effects of chondroitin sulfate are on inflammatory responses and thus

future work should focus on the use of osteoarthritic cells to determine any potential

benefits of the chondroitin sulfate containing films.

The major part of this thesis focused on 3D microenvironments. 3D scaffolds

were designed to present biomimetic cartilage-like microenvironments to

encapsulated chondrocytes and stem cells. The use of the alginate sulfate material as a

hydrogel for cartilage tissue engineering was a major breakthrough in this thesis. We

showed that by simply modifying alginate with sulfate moieties, cell proliferation was

increased 5 fold and the hydrogel attained a cartilage-like appearance after only 5

weeks in culture. Most importantly, the material could preserve the cartilage

phenotype of encapsulated chondrocytes similar to unmodified alginate which is well

known to maintain the cartilage phenotype. Future work should focus on investigating

proliferation of non-passaged chondrocytes within alginate sulfate to see if de-

differentiation can be completely eliminated and to study the performance of this

novel material in animal models.

Conclusions and outlook

108

We have also seen interesting responses for stem cells encapsulated in the

GFGOER degradable hydrogels. Our studies were carried out for a maximum of 21 days

and we observed an opaque appearance of these constructs as a result of high

collagen deposition and cell proliferation. We believe that longer times are necessary

to fully establish chondrogenic differentiation of the stem cells. Therefore future work

should investigate longer culture periods within the GFOGER modified hydrogels.

Finally, we made a step towards building a stratified articular cartilage. We first

present superficial zone protein as a new marker for chondrogenesis, whose

expression drops during passaging and is fully recovered by day 4 encapsulation in

alginate under normoxic conditions. This expression was further augmented with

mechanical strain. On the other hand, type II collagen was only partially restored in

alginate beads under hypoxic conditions. We have also seen in data not presented in

this thesis that longer time periods are required to recover type II collagen expression

and potentially different mechanical stimuli such as compression or hydrostatic

pressure could be beneficial. Future work should help identify the parameters

important in constructing stratified cartilage.

In this thesis we worked on a variety of techniques to engineer 2D and 3D

microenvironments. We were successful in preparing natural ECM based films on

stretchable substrates that may be of relevance in a variety of biomedical applications.

We presented for the first time the alginate sulfate hydrogel as a material for cartilage

tissue engineering applications. We showed that the GFOGER adhesion peptide

presented in a degradable PEG hydrogel provides a better chondrogenic

microenvironment than the more commonly used RGD peptide. Finally, we

determined parameters that regulate expression of superficial zone protein as a step

towards building stratified articular cartilage. The results of this thesis are expected to

have a high impact on the tissue engineering field particularly for cartilage

applications.

Chapter 9

109

9 References

[1] Berthiaume F, Maguire TJ, Yarmush ML. Tissue Engineering and Regenerative Medicine: History, Progress, and Challenges. Annual Review of Chemical and Biomolecular Engineering. 2011;2:403-30.

[2] Langer R, Vacanti JP. Tissue Engineering. Science. 1993;260:920-6.

[3] Webster JP. Refrigerated Skin Grafts *. Annals of Surgery. 1944;120:431-49.

[4] Polge C, Smith A, Parkes A. Revival of spermatozoa after vitrification and dehydration at low temperatures. Nature. 1949;164:666.

[5] Billingham R, Medawar P. The freezing, drying and storage of mammalian skin. Journal of Experimental Biology. 1952;29:454-68.

[6] Rheinwatd JG, Green H. Seria cultivation of strains of human epidemal keratinocytes: the formation keratinizin colonies from single cell is. Cell. 1975;6:331-43.

[7] Green H, Kehinde O, Thomas J. Growth of cultured human epidermal cells into multiple epithelia suitable for grafting. Proceedings of the National Academy of Sciences. 1979;76:5665-8.

[8] Bell E, Ehrlich H, Buttle D, Nakatsuji T. Living tissue formed in vitro and accepted as skin-equivalent tissue of full thickness. Science. 1981;211:1052-4.

[9] Viola J, Lal B, Grad O. The emergence of tissue engineering as a research field. http://www.nsf.gov/pubs/2004/nsf0450/exec_summ.pdf; 2003.

[10] Eberli D, Filho LF, Atala A, Yoo JJ. Composite scaffolds for the engineering of hollow organs and tissues. Methods. 2009;47:109-15.

[11] Chen FH, Rousche KT, Tuan RS. Technology Insight: adult stem cells in cartilage regeneration and tissue engineering. Nat Clin Pract Rheum. 2006;2:373-82.

[12] Becerra J, Andrades JA, Guerado E, Zamora-Navas P, López-Puertas JM, Reddi AH. Articular cartilage: structure and regeneration. Tissue Engineering Part B: Reviews. 2010;16:617-27.

[13] Buckwalter JA, Mankin HJ, Grodzinsky AJ. Articular cartilage and osteoarthritis. INSTRUCTIONAL COURSE LECTURES-AMERICAN ACADEMY OF ORTHOPAEDIC SURGEONS. 2005;54:465.

[14] Chiang H, Jiang CC. Repair of articular cartilage defects: review and perspectives. Journal of the Formosan Medical Association. 2009;108:87-101.

[15] Guilak F, Ratcliffe A, Lane N, Rosenwasser MP, Mow VC. Mechanical and biochemical changes in the superficial zone of articular cartilage in canine experimental osteoarthritis. Journal of Orthopaedic Research. 1994;12:474-84.

[16] Jay GD, Tantravahi U, Britt DE, Barrach HJ, Cha C-J. Homology of lubricin and superficial zone protein (SZP): Products of megakaryocyte stimulating factor (MSF) gene expression by human synovial fibroblasts and articular chondrocytes localized to chromosome 1q25. Journal of Orthopaedic Research. 2001;19:677-87.

References

110

[17] Schumacher BL, Block JA, Schmid TM, Aydelotte MB, Kuettner KE. A Novel Proteoglycan Synthesized and Secreted by Chondrocytes of the Superficial Zone of Articular Cartilage. Archives of Biochemistry and Biophysics. 1994;311:144-52.

[18] Li L, Xie T. Stem cell niche: structure and function. Annu Rev Cell Dev Biol. 2005;21:605-31.

[19] Gigout A, Jolicoeur M, Nelea M, Raynal N, Farndale R, Buschmann MD. Chondrocyte Aggregation in Suspension Culture Is GFOGER-GPP- and β1 Integrin-dependent. Journal of Biological Chemistry. 2008;283:31522-30.

[20] Rolls A, Shechter R, London A, Segev Y, Jacob-Hirsch J, Amariglio N, et al. Two faces of chondroitin sulfate proteoglycan in spinal cord repair: a role in microglia/macrophage activation. PLoS medicine. 2008;5:e171.

[21] Grimshaw MJ, Mason RM. Modulation of bovine articular chondrocyte gene expression in vitro by oxygen tension. Osteoarthritis and Cartilage. 2001;9:357-64.

[22] Wong M, Carter DR. Articular cartilage functional histomorphology and mechanobiology: a research perspective. Bone. 2003;33:1-13.

[23] DeLise AM, Fischer L, Tuan RS. Cellular interactions and signaling in cartilage development. Osteoarthritis and Cartilage. 2000;8:309-34.

[24] Woo SLY, Buckwalter JA. Injury and repair of the musculoskeletal soft tissues: workshop, Savannah, Georgia, June 1987: American Academy of Orthopaedic Surgeons; 1988.

[25] Herzog W, Longino D. The role of muscles in joint degeneration and osteoarthritis. Journal of Biomechanics. 2007;40, Supplement 1:S54-S63.

[26] Schulz R, Bader A. Cartilage tissue engineering and bioreactor systems for the cultivation and stimulation of chondrocytes. Eur Biophys J. 2007;36:539-68.

[27] Kubo M, Ando K, Mimura T, Matsusue Y, Mori K. Chondroitin sulfate for the treatment of hip and knee osteoarthritis: Current status and future trends. Life Sciences. 2009;85:477-83.

[28] Thelin N, Holmberg S, Thelin A. Knee injuries account for the sports-related increased risk of knee osteoarthritis. Scandinavian Journal of Medicine & Science in Sports. 2006;16:329-33.

[29] Ochi M, Uchio Y, Kawasaki K, Wakitani S, Iwasa J. Transplantation of cartilage-like tissue made by tissue engineering in the treatment of cartilage defects of the knee. Journal of Bone & Joint Surgery, British Volume. 2002;84-B:571-8.

[30] Lequesne M, Brandt K, Bellamy N, Moskowitz R, Menkes CJ, Pelletier JP, et al. Guidelines for testing slow acting drugs in osteoarthritis. The Journal of rheumatology Supplement. 1994;41:65-71; discussion 2-3.

[31] Zhang W, Moskowitz RW, Nuki G, Abramson S, Altman RD, Arden N, et al. OARSI recommendations for the management of hip and knee osteoarthritis, Part I: Critical appraisal of existing treatment guidelines and systematic review of current research evidence. Osteoarthritis and Cartilage. 2007;15:981-1000.

[32] Zhang W, Moskowitz RW, Nuki G, Abramson S, Altman RD, Arden N, et al. OARSI recommendations for the management of hip and knee osteoarthritis, Part II: OARSI evidence-based, expert consensus guidelines. Osteoarthritis and Cartilage. 2008;16:137-62.

[33] Stürmer T, Erb A, Keller F, Günther K-P, Brenner H. Determinants of impaired renal function with use of nonsteroidal anti-inflammatory drugs: the importance of half-life and other medications. The American Journal of Medicine. 2001;111:521-7.

Chapter 9

111

[34] Winkelmayer WC, Waikar SS, Mogun H, Solomon DH. Nonselective and Cyclooxygenase-2-Selective NSAIDs and Acute Kidney Injury. The American Journal of Medicine. 2008;121:1092-8.

[35] Laine L. Nonsteroidal anti-inflammatory drug gastropathy. Gastrointestinal endoscopy clinics of North America. 1996;6:489-504.

[36] Lanza FL, Chan FKL, Quigley EMM. Guidelines for Prevention of NSAID-Related Ulcer Complications. Am J Gastroenterol. 2009;104:728-38.

[37] Wu C-H, Ko C-S, Huang J-W, Huang H-J, Chu IM. Effects of exogenous glycosaminoglycans on human chondrocytes cultivated on type II collagen scaffolds. J Mater Sci: Mater Med. 2010;21:725-9.

[38] Sharma G, Saxena RK, Mishra P. Synergistic effect of chondroitin sulfate and cyclic pressure on biochemical and morphological properties of chondrocytes from articular cartilage. Osteoarthritis and Cartilage. 2008;16:1387-94.

[39] Michel BA, Stucki G, Frey D, De Vathaire F, Vignon E, Bruehlmann P, et al. Chondroitins 4 and 6 sulfate in osteoarthritis of the knee: A randomized, controlled trial. Arthritis & Rheumatism. 2005;52:779-86.

[40] Uebelhart D, Thonar EJMA, Zhang J, Williams JM. Protective effect of exogenous chondroitin 4,6-sulfate in the acute degradation of articular cartilage in the rabbit. Osteoarthritis and Cartilage. 1998;6, Supplement A:6-13.

[41] Jomphe C, Gabriac M, Hale TM, Héroux L, Trudeau L-É, Deblois D, et al. Chondroitin Sulfate Inhibits the Nuclear Translocation of Nuclear Factor-κB in Interleukin-1β-Stimulated Chondrocytes. Basic & Clinical Pharmacology & Toxicology. 2008;102:59-65.

[42] Legendre F, Baugé C, Roche R, Saurel AS, Pujol JP. Chondroitin sulfate modulation of matrix and inflammatory gene expression in IL-1beta-stimulated chondrocytes--study in hypoxic alginate bead cultures. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society. 2008;16:105-14.

[43] Bassleer C, Henrotin Y, Franchimont P. In-vitro evaluation of drugs proposed as chondroprotective agents. International journal of tissue reactions. 1992;14:231-41.

[44] Johnson KA, Hulse DA, Hart RC, Kochevar D, Chu Q. Effects of an orally administered mixture of chondroitin sulfate, glucosamine hydrochloride and manganese ascorbate on synovial fluid chondroitin sulfate 3B3 and 7D4 epitope in a canine cruciate ligament transection model of osteoarthritis. Osteoarthritis and Cartilage. 2001;9:14-21.

[45] Lippiello L, Woodward J, Karpman R, Hammad TA. In Vivo Chondroprotection and Metabolic Synergy of Glucosamine and Chondroitin Sulfate. Clinical Orthopaedics and Related Research. 2000;381:229-40.

[46] Ronca F, Palmieri L, Panicucci P, Ronca G. Anti-inflammatory activity of chondroitin sulfate. Osteoarthritis and Cartilage. 1998;6, Supplement A:14-21.

[47] Bourgeois P, Chales G, Dehais J, Delcambre B, Kuntz J-L, Rozenberg S. Efficacy and tolerability of chondroitin sulfate 1200 mg/day vs chondroitin sulfate 3×400 mg/day vs placebo. Osteoarthritis and Cartilage. 1998;6, Supplement A:25-30.

References

112

[48] Bucsi L, Poór G. Efficacy and tolerability of oral chondroitin sulfate as a symptomatic slow-acting drug for osteoarthritis (SYSADOA) in the treatment of knee osteoarthritis. Osteoarthritis and Cartilage. 1998;6, Supplement A:31-6.

[49] Strauss EJ, Fonseca LE, Shah MR, Yorum T. Management of focal cartilage defects in the knee-Is ACI the answer? Bulletin of the NYU hospital for joint diseases. 2011;69:63.

[50] Bhosale AM, Richardson JB. Articular cartilage: structure, injuries and review of management. British Medical Bulletin. 2008;87:77-95.

[51] Grande DA, Pitman MI, Peterson L, Menche D, Klein M. The repair of experimentally produced defects in rabbit articular cartilage by autologous chondrocyte transplantation. Journal of Orthopaedic Research. 1989;7:208-18.

[52] Brittberg M, Lindahl A, Nilsson A, Ohlsson C, Isaksson O, Peterson L. Treatment of deep cartilage defects in the knee with autologous chondrocyte transplantation. New england journal of medicine. 1994;331:889-95.

[53] Brittberg M, Peterson L, Sjouml, gren-Jansson E, Tallheden T, Lindahl A. Articular Cartilage Engineering with Autologous Chondrocyte Transplantation A Review of Recent Developments. The Journal of Bone & Joint Surgery. 2003;85:109-15.

[54] Henderson I, Gui J, Lavigne P. Autologous Chondrocyte Implantation: Natural History of Postimplantation Periosteal Hypertrophy and Effects of Repair-Site Debridement on Outcome. Arthroscopy: The Journal of Arthroscopic & Related Surgery. 2006;22:1318-24.e1.

[55] Rosenberger RE, Gomoll AH, Bryant T, Minas T. Repair of Large Chondral Defects of the Knee With Autologous Chondrocyte Implantation in Patients 45 Years or Older. The American Journal of Sports Medicine. 2008;36:2336-44.

[56] Haddo O, Mahroof S, Higgs D, David L, Pringle J, Bayliss M, et al. The use of chondrogide membrane in autologous chondrocyte implantation. The Knee. 2004;11:51-5.

[57] Behrens P, Bitter T, Kurz B, Russlies M. Matrix-associated autologous chondrocyte transplantation/implantation (MACT/MACI)—5-year follow-up. The Knee. 2006;13:194-202.

[58] Behrens P, Ehlers EM, Köchermann KU, Rohwedel J, Russlies M, Plötz W. New therapy procedure for localized cartilage defects. Encouraging results with autologous chondrocyte implantation. MMW Fortschritte der Medizin. 1999;141:49-51.

[59] Schiltz JR, Mayne R, Holtzer H. The Synthesis of Collagen and Glycosaminoglycans by Dedifferentiated Chondroblasts in Culture. Differentiation. 1973;1:97-108.

[60] Benya PD, Shaffer JD. Dedifferentiated chondrocytes reexpress the differentiated collagen phenotype when cultured in agarose gels. Cell. 1982;30:215-24.

[61] Darling EM, Athanasiou KA. Rapid phenotypic changes in passaged articular chondrocyte subpopulations. Journal of Orthopaedic Research. 2005;23:425-32.

[62] Schnabel M, Marlovits S, Eckhoff G, Fichtel I, Gotzen L, Vécsei V, et al. Dedifferentiation-associated changes in morphology and gene expression in primary human articular chondrocytes in cell culture. Osteoarthritis and Cartilage. 2002;10:62-70.

[63] Darling E, Pritchett P, Evans B, Superfine R, Zauscher S, Guilak F. Mechanical Properties and Gene Expression of Chondrocytes on Micropatterned Substrates Following Dedifferentiation in Monolayer. Cel Mol Bioeng. 2009;2:395-404.

Chapter 9

113

[64] Domm C, Schünke M, Christesen K, Kurz B. Redifferentiation of dedifferentiated bovine articular chondrocytes in alginate culture under low oxygen tension. Osteoarthritis and Cartilage. 2002;10:13-22.

[65] Candrian C, Vonwil D, Barbero A, Bonacina E, Miot S, Farhadi J, et al. Engineered cartilage generated by nasal chondrocytes is responsive to physical forces resembling joint loading. Arthritis & Rheumatism. 2008;58:197-208.

[66] Candrian C, Miot S, Wolf F, Bonacina E, Dickinson S, Wirz D, et al. Are ankle chondrocytes from damaged fragments a suitable cell source for cartilage repair? Osteoarthritis and Cartilage. 2010;18:1067-76.

[67] Mackay AM, Beck SC, Murphy JM, Barry FP, Chichester CO, Pittenger MF. Chondrogenic differentiation of cultured human mesenchymal stem cells from marrow. Tissue engineering. 1998;4:415-28.

[68] Awad HA, Halvorsen YD, Gimble JM, Guilak F. Effects of transforming growth factor beta1 and dexamethasone on the growth and chondrogenic differentiation of adipose-derived stromal cells. Tissue engineering. 2003;9:1301-12.

[69] Tew SR, Murdoch AD, Rauchenberg RP, Hardingham TE. Cellular methods in cartilage research: Primary human chondrocytes in culture and chondrogenesis in human bone marrow stem cells. Methods. 2008;45:2-9.

[70] Kumar D, Lassar AB. The transcriptional activity of Sox9 in chondrocytes is regulated by RhoA signaling and actin polymerization. Molecular and Cellular Biology. 2009;29:4262-73.

[71] Woods A, Wang G, Beier F. RhoA/ROCK signaling regulates Sox9 expression and actin organization during chondrogenesis. The Journal of biological chemistry. 2005;280:11626-34.

[72] Bonaventure J, Kadhom N, Cohen-Solal L, Ng KH, Bourguignon J, Lasselin C, et al. Reexpression of Cartilage-Specific Genes by Dedifferentiated Human Articular Chondrocytes Cultured in Alginate Beads. Experimental Cell Research. 1994;212:97-104.

[73] Park Y, Lutolf MP, Hubbell JA, Hunziker EB, Wong M. Bovine primary chondrocyte culture in synthetic matrix metalloproteinase-sensitive poly(ethylene glycol)-based hydrogels as a scaffold for cartilage repair. Tissue engineering. 2004;10:515-22.

[74] Lutolf MP, Hubbell JA. Synthetic biomaterials as instructive extracellular microenvironments for morphogenesis in tissue engineering. Nat Biotech. 2005;23:47-55.

[75] Kraehenbuehl TP, Zammaretti P, Van der Vlies AJ, Schoenmakers RG, Lutolf MP, Jaconi ME, et al. Three-dimensional extracellular matrix-directed cardioprogenitor differentiation: Systematic modulation of a synthetic cell-responsive PEG-hydrogel. Biomaterials. 2008;29:2757-66.

[76] Lee ST, Yun JI, Jo YS, Mochizuki M, van der Vlies AJ, Kontos S, et al. Engineering integrin signaling for promoting embryonic stem cell self-renewal in a precisely defined niche. Biomaterials. 2010;31:1219-26.

[77] Yan WQ, Nakashima K, Iwamoto M, Kato Y. Stimulation by concanavalin A of cartilage-matrix proteoglycan synthesis in chondrocyte cultures. Journal of Biological Chemistry. 1990;265:10125-31.

[78] Petersen EF, Spencer RGS, McFarland EW. Microengineering neocartilage scaffolds. Biotechnology and Bioengineering. 2002;78:802-5.

References

114

[79] Boudou T, Crouzier T, Ren K, Blin G, Picart C. Multiple Functionalities of Polyelectrolyte Multilayer Films: New Biomedical Applications. Advanced Materials. 2010;22:441-67.

[80] Ai H, Jones S, Lvov Y. Biomedical applications of electrostatic layer-by-layer nano-assembly of polymers, enzymes, and nanoparticles. Cell Biochem Biophys. 2003;39:23-43.

[81] Decher G. Fuzzy Nanoassemblies: Toward Layered Polymeric Multicomposites. Science. 1997;277:1232-7.

[82] Zhang J, Senger B, Vautier D, Picart C, Schaaf P, Voegel J-C, et al. Natural polyelectrolyte films based on layer-by layer deposition of collagen and hyaluronic acid. Biomaterials. 2005;26:3353-61.

[83] Semenov OV, Malek A, Bittermann AG, Vörös J, Zisch AH. Engineered polyelectrolyte multilayer substrates for adhesion, proliferation, and differentiation of human mesenchymal stem cells. Tissue Engineering Part A. 2009;15:2977-90.

[84] Richert L, Boulmedais F, Lavalle P, Mutterer J, Ferreux E, Decher G, et al. Improvement of Stability and Cell Adhesion Properties of Polyelectrolyte Multilayer Films by Chemical Cross-Linking. Biomacromolecules. 2003;5:284-94.

[85] Thompson MT, Berg MC, Tobias IS, Rubner MF, Van Vliet KJ. Tuning compliance of nanoscale polyelectrolyte multilayers to modulate cell adhesion. Biomaterials. 2005;26:6836-45.

[86] Ai H, Lvov Y, Mills D, Jennings M, Alexander J, Jones S. Coating and selective deposition of nanofilm on silicone rubber for cell adhesion and growth. Cell Biochem Biophys. 2003;38:103-14.

[87] Chanana M, Gliozzi A, Diaspro A, Chodnevskaja I, Huewel S, Moskalenko V, et al. Interaction of Polyelectrolytes and Their Composites with Living Cells. Nano Letters. 2005;5:2605-12.

[88] Johansson JÅ, Halthur T, Herranen M, Söderberg L, Elofsson U, Hilborn J. Build-up of Collagen and Hyaluronic Acid Polyelectrolyte Multilayers. Biomacromolecules. 2005;6:1353-9.

[89] Hahn SK, Hoffman AS. Preparation and characterization of biocompatible polyelectrolyte complex multilayer of hyaluronic acid and poly-l-lysine. International Journal of Biological Macromolecules. 2005;37:227-31.

[90] Jiang F, Hörber H, Howard J, Müller DJ. Assembly of collagen into microribbons: effects of pH and electrolytes. Journal of Structural Biology. 2004;148:268-78.

[91] Volpi N, Cusmano M, Venturelli T. Qualitative and quantitative studies of heparin and chondroitin sulfates in normal human plasma. Biochimica et Biophysica Acta (BBA) - General Subjects. 1995;1243:49-58.

[92] Vincent JFV, Bogatyreva OA, Bogatyrev NR, Bowyer A, Pahl A-K. Biomimetics: its practice and theory. Journal of The Royal Society Interface. 2006;3:471-82.

[93] Giannoni P, Siegrist M, Hunziker EB, Wong M. The mechanosensitivity of cartilage oligomeric matrix protein (COMP). Biorheology. 2003;40:101-9.

[94] Wang JHC, Thampatty BP. Chapter 7 Mechanobiology of Adult and Stem Cells. In: Kwang WJ, editor. International Review of Cell and Molecular Biology: Academic Press; 2008. p. 301-46.

Chapter 9

115

[95] Schrobback K, Malda J, Crawford RW, Upton Z, Leavesley DI, Klein TJ. Effects of oxygen on zonal marker expression in human articular chondrocytes. Tissue Engineering Part A. 2012;18:920-33.

[96] Connelly JT, García AJ, Levenston ME. Inhibition of in vitro chondrogenesis in RGD-modified three-dimensional alginate gels. Biomaterials. 2007;28:1071-83.

[97] Connelly JT, García AJ, Levenston ME. Interactions between integrin ligand density and cytoskeletal integrity regulate BMSC chondrogenesis. Journal of Cellular Physiology. 2008;217:145-54.

[98] Knight CG, Morton LF, Peachey AR, Tuckwell DS, Farndale RW, Barnes MJ. The Collagen-binding A-domains of Integrins α1β1 and α2β1Recognize the Same Specific Amino Acid Sequence, GFOGER, in Native (Triple-helical) Collagens. Journal of Biological Chemistry. 2000;275:35-40.

[99] Millward-Sadler SJ, Wright MO, Davies LW, Nuki G, Salter DM. Mechanotransduction via integrins and interleukin-4 results in altered aggrecan and matrix metalloproteinase 3 gene expression in normal, but not osteoarthritic, human articular chondrocytes. Arthritis & Rheumatism. 2000;43:2091-9.

[100] Sah RLY, Kim Y-J, Doong J-YH, Grodzinsky AJ, Plass AHK, Sandy JD. Biosynthetic response of cartilage explants to dynamic compression. Journal of Orthopaedic Research. 1989;7:619-36.

[101] Lee JH, Fitzgerald JB, DiMicco MA, Grodzinsky AJ. Mechanical injury of cartilage explants causes specific time-dependent changes in chondrocyte gene expression. Arthritis & Rheumatism. 2005;52:2386-95.

[102] Scherberich A, Galli R, Jaquiery C, Farhadi J, Martin I. Three-Dimensional Perfusion Culture of Human Adipose Tissue-Derived Endothelial and Osteoblastic Progenitors Generates Osteogenic Constructs with Intrinsic Vascularization Capacity. STEM CELLS. 2007;25:1823-9.

[103] Altman GH, Horan RL, Martin I, Farhadi J, Stark PR, Volloch V, et al. Cell differentiation by mechanical stress. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2002;16:270-2.

[104] Juncosa-Melvin N, Matlin KS, Holdcraft RW, Nirmalanandhan VS, Butler DL. Mechanical stimulation increases collagen type I and collagen type III gene expression of stem cell-collagen sponge constructs for patellar tendon repair. Tissue engineering. 2007;13:1219-26.

[105] Chen Y-J, Huang C-H, Lee I-C, Lee Y-T, Chen M-H, Young T-H. Effects of Cyclic Mechanical Stretching on the mRNA Expression of Tendon/Ligament-Related and Osteoblast-Specific Genes in Human Mesenchymal Stem Cells. Connective Tissue Research. 2008;49:7-14.

[106] Davisson T, Kunig S, Chen A, Sah R, Ratcliffe A. Static and dynamic compression modulate matrix metabolism in tissue engineered cartilage. Journal of Orthopaedic Research. 2002;20:842-8.

[107] Wang H, Riha GM, Yan S, Li M, Chai H, Yang H, et al. Shear Stress Induces Endothelial Differentiation From a Murine Embryonic Mesenchymal Progenitor Cell Line. Arteriosclerosis, Thrombosis, and Vascular Biology. 2005;25:1817-23.

[108] Hirano Y, Ishiguro N, Sokabe M, Takigawa M, Naruse K. Effects of tensile and compressive strains on response of a chondrocytic cell line embedded in type I collagen gel. Journal of Biotechnology. 2008;133:245-52.

References

116

[109] Sumanasinghe RD, Bernacki SH, Loboa EG. Osteogenic differentiation of human mesenchymal stem cells in collagen matrices: effect of uniaxial cyclic tensile strain on bone morphogenetic protein (BMP-2) mRNA expression. Tissue engineering. 2006;12:3459-65.

[110] Illi B, Scopece A, Nanni S, Farsetti A, Morgante L, Biglioli P, et al. Epigenetic Histone Modification and Cardiovascular Lineage Programming in Mouse Embryonic Stem Cells Exposed to Laminar Shear Stress. Circulation Research. 2005;96:501-8.

[111] Angele P, Yoo JU, Smith C, Mansour J, Jepsen KJ, Nerlich M, et al. Cyclic hydrostatic pressure enhances the chondrogenic phenotype of human mesenchymal progenitor cells differentiated in vitro. Journal of Orthopaedic Research. 2003;21:451-7.

[112] Angele P, Schumann D, Angele M, Kinner B, Englert C, Hente R, et al. Cyclic, mechanical compression enhances chondrogenesis of mesenchymal progenitor cells in tissue engineering scaffolds. Biorheology. 2004;41:335-46.

[113] Wagner D, Lindsey D, Li K, Tummala P, Chandran S, Smith RL, et al. Hydrostatic Pressure Enhances Chondrogenic Differentiation of Human Bone Marrow Stromal Cells in Osteochondrogenic Medium. Annals of Biomedical Engineering. 2008;36:813-20.

[114] Toyoda T, Seedhom BB, Yao JQ, Kirkham J, Brookes S, Bonass WA. Hydrostatic pressure modulates proteoglycan metabolism in chondrocytes seeded in agarose. Arthritis & Rheumatism. 2003;48:2865-72.

[115] Smith RL, Rusk SF, Ellison BE, Wessells P, Tsuchiya K, Carter DR, et al. In vitro stimulation of articular chondrocyte mRNA and extracellular matrix synthesis by hydrostatic pressure. Journal of Orthopaedic Research. 1996;14:53-60.

[116] Grad S, Lee CR, Gorna K, Gogolewski S, Wimmer MA, Alini M. Surface motion upregulates superficial zone protein and hyaluronan production in chondrocyte-seeded three-dimensional scaffolds. Tissue engineering. 2005;11:249-56.

[117] Kamiya T, Tanimoto K, Tanne Y, Lin YY, Kunimatsu R, Yoshioka M, et al. Effects of mechanical stimuli on the synthesis of superficial zone protein in chondrocytes. Journal of Biomedical Materials Research Part A. 2010;92A:801-5.

[118] Nugent GE, Aneloski NM, Schmidt TA, Schumacher BL, Voegtline MS, Sah RL. Dynamic shear stimulation of bovine cartilage biosynthesis of proteoglycan 4. Arthritis & Rheumatism. 2006;54:1888-96.

[119] Ramage L, Nuki G, Salter DM. Signalling cascades in mechanotransduction: cell–matrix interactions and mechanical loading. Scandinavian Journal of Medicine & Science in Sports. 2009;19:457-69.

[120] Iqbal J, Zaidi M. Molecular regulation of mechanotransduction. Biochemical and Biophysical Research Communications. 2005;328:751-5.

[121] Lee S, Vörös J. An aqueous-based surface modification of poly (dimethylsiloxane) with poly (ethylene glycol) to prevent biofouling. Langmuir. 2005;21:11957-62.

[122] Studer D, Lischer S, Jochum W, Ehrbar M, Zenobi-Wong M, Maniura-Weber K. Ribosomal Protein L13a as a Reference Gene for Human Bone Marrow-Derived Mesenchymal Stromal Cells During Expansion, Adipo-, Chondro-, and Osteogenesis. Tissue Engineering Part C: Methods. 2012;18:761-72.

Chapter 9

117

[123] Livak KJ, Schmittgen TD. Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2−ΔΔCT Method. Methods. 2001;25:402-8.

[124] Mhanna RF, Vörös J, Zenobi-Wong M. Layer-by-layer films made from extracellular matrix macromolecules on silicone substrates. Biomacromolecules. 2011;12:609-16.

[125] Tsutsumi S, Shimazu A, Miyazaki K, Pan H, Koike C, Yoshida E, et al. Retention of Multilineage Differentiation Potential of Mesenchymal Cells during Proliferation in Response to FGF. Biochemical and Biophysical Research Communications. 2001;288:413-9.

[126] Kuo CK, Ma PX. Ionically crosslinked alginate hydrogels as scaffolds for tissue engineering: Part 1. Structure, gelation rate and mechanical properties. Biomaterials. 2001;22:511-21.

[127] Ingar Draget K, Østgaard K, Smidsrød O. Homogeneous alginate gels: A technical approach. Carbohydrate Polymers. 1990;14:159-78.

[128] McAloney RA, Sinyor M, Dudnik V, Goh MC. Atomic Force Microscopy Studies of Salt Effects on Polyelectrolyte Multilayer Film Morphology. Langmuir. 2001;17:6655-63.

[129] Estes BT, Diekman BO, Gimble JM, Guilak F. Isolation of adipose-derived stem cells and their induction to a chondrogenic phenotype. Nat Protocols. 2010;5:1294-311.

[130] Al-Soud WA, Ouis I-S, Li D-Q, Ljungh Å, Wadström T. Characterization of the PCR inhibitory effect of bile to optimize real-time PCR detection of Helicobacter species. FEMS Immunology & Medical Microbiology. 2005;44:177-82.

[131] Hübsch E, Ball V, Senger B, Decher G, Voegel J-C, Schaaf P. Controlling the Growth Regime of Polyelectrolyte Multilayer Films:  Changing from Exponential to Linear Growth by Adjusting the Composition of Polyelectrolyte Mixtures. Langmuir. 2004;20:1980-5.

[132] Liu Y, He T, Song H, Gao C. Layer-by-layer assembly of biomacromolecules on poly(ethylene terephthalate) films and fiber fabrics to promote endothelial cell growth. Journal of Biomedical Materials Research Part A. 2007;81A:692-704.

[133] Liu Y, He T, Gao C. Surface modification of poly(ethylene terephthalate) via hydrolysis and layer-by-layer assembly of chitosan and chondroitin sulfate to construct cytocompatible layer for human endothelial cells. Colloids and Surfaces B: Biointerfaces. 2005;46:117-26.

[134] Lin Y, Wang L, Zhang P, Wang X, Chen X, Jing X, et al. Surface modification of poly(l-lactic acid) to improve its cytocompatibility via assembly of polyelectrolytes and gelatin. Acta Biomaterialia. 2006;2:155-64.

[135] Boulmedais F, Frisch B, Etienne O, Lavalle P, Picart C, Ogier J, et al. Polyelectrolyte multilayer films with pegylated polypeptides as a new type of anti-microbial protection for biomaterials. Biomaterials. 2004;25:2003-11.

[136] Luescher M, Rüegg M, Schindler P. Effect of hydration upon the thermal stability of tropocollagen and its dependence on the presence of neutral salts. Biopolymers. 1974;13:2489-503.

[137] Staprans I, Felts JM. Isolation and characterization of glycosaminoglycans in human plasma. The Journal of clinical investigation. 1985;76:1984-91.

[138] Wakitani S, Goto T, Young RG, Mansour JM, Goldberg VM, Caplan AI. Repair of large full-thickness articular cartilage defects with allograft articular chondrocytes embedded in a collagen gel. Tissue engineering. 1998;4:429-44.

References

118

[139] Varghese S, Hwang NS, Canver AC, Theprungsirikul P, Lin DW, Elisseeff J. Chondroitin sulfate based niches for chondrogenic differentiation of mesenchymal stem cells. Matrix Biology. 2008;27:12-21.

[140] Kim HK, Shim WS, Kim SE, Lee K-H, Kang E, Kim J-H, et al. Injectable In Situ–Forming pH/Thermo-Sensitive Hydrogel for Bone Tissue Engineering Tissue Engineering Part A. 2009;15:923-33.

[141] Sia SK, Whitesides GM. Microfluidic devices fabricated in Poly(dimethylsiloxane) for biological studies. ELECTROPHORESIS. 2003;24:3563-76.

[142] Bridges AJ, Conley C, Wang G, Burns DE, Vasey FB. A Clinical and Immunologic Evaluation of Women with Silicone Breast Implants and Symptoms of Rheumatic Disease. Annals of Internal Medicine. 1993;118:929-36.

[143] Okada T, Ikada Y. Modification of silicone surface by graft polymerization of acrylamide with corona discharge. Die Makromolekulare Chemie. 1991;192:1705-13.

[144] Abbasi F, Mirzadeh H, Katbab A-A. Modification of polysiloxane polymers for biomedical applications: a review. Polymer International. 2001;50:1279-87.

[145] Amma H, Naruse K, Ishiguro N, Sokabe M. Involvement of reactive oxygen species in cyclic stretch-induced NF-κB activation in human fibroblast cells. British Journal of Pharmacology. 2005;145:364-73.

[146] Lawton RA, Price CR, Runge AF, Doherty Iii WJ, Saavedra SS. Air plasma treatment of submicron thick PDMS polymer films: effect of oxidation time and storage conditions. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2005;253:213-5.

[147] Morra M, Occhiello E, Marola R, Garbassi F, Humphrey P, Johnson D. On the aging of oxygen plasma-treated polydimethylsiloxane surfaces. Journal of Colloid and Interface Science. 1990;137:11-24.

[148] Bodas D, Khan-Malek C. Hydrophilization and hydrophobic recovery of PDMS by oxygen plasma and chemical treatment—An SEM investigation. Sensors and Actuators B: Chemical. 2007;123:368-73.

[149] Richert L, Lavalle P, Payan E, Shu XZ, Prestwich GD, Stoltz J-F, et al. Layer by Layer Buildup of Polysaccharide Films: Physical Chemistry and Cellular Adhesion Aspects. Langmuir. 2003;20:448-58.

[150] Kujawa P, Moraille P, Sanchez J, Badia A, Winnik FM. Effect of Molecular Weight on the Exponential Growth and Morphology of Hyaluronan/Chitosan Multilayers:  A Surface Plasmon Resonance Spectroscopy and Atomic Force Microscopy Investigation. Journal of the American Chemical Society. 2005;127:9224-34.

[151] Rossetti FF, Reviakine I, Csúcs G, Assi F, Vörös J, Textor M. Interaction of Poly(L-Lysine)-g-Poly(Ethylene Glycol) with Supported Phospholipid Bilayers. Biophysical Journal. 2004;87:1711-21.

[152] Granéli A, Edvardsson M, Höök F. DNA-Based Formation of a Supported, Three-Dimensional Lipid Vesicle Matrix Probed by QCM-D and SPR. ChemPhysChem. 2004;5:729-33.

[153] Wallace DG. The relative contribution of electrostatic interactions to stabilization of collagen fibrils. Biopolymers. 1990;29:1015-26.

Chapter 9

119

[154] Jokinen J, Dadu E, Nykvist P, Käpylä J, White DJ, Ivaska J, et al. Integrin-mediated Cell Adhesion to Type I Collagen Fibrils. Journal of Biological Chemistry. 2004;279:31956-63.

[155] Humphries JD, Byron A, Humphries MJ. Integrin ligands at a glance. Journal of Cell Science. 2006;119:3901-3.

[156] Grande DA, Southerland SS, Manji R, Pate DW, Schwartz RE, Lucas PA. Repair of articular cartilage defects using mesenchymal stem cells. Tissue engineering. 1995;1:345-53.

[157] Lim H-C, Bae J-H, Song S-H, Park Y-E, Kim S-J. Current Treatments of Isolated Articular Cartilage Lesions of the Knee Achieve Similar Outcomes. Clin Orthop Relat Res. 2012;470:2261-7.

[158] Vasiliadis H, Wasiak J, Salanti G. Autologous chondrocyte implantation for the treatment of cartilage lesions of the knee: a systematic review of randomized studies. Knee Surg Sports Traumatol Arthrosc. 2010;18:1645-55.

[159] Tallheden T, Bengtsson C, Brantsing C, Sjogren-Jansson E, Carlsson L, Peterson L, et al. Proliferation and differentiation potential of chondrocytes from osteoarthritic patients. Arthritis Research & Therapy. 2005;7:R560 - R8.

[160] Giannoni P, Cancedda R. Articular Chondrocyte Culturing for Cell-Based Cartilage Repair: Needs and Perspectives. Cells Tissues Organs. 2006;184:1-15.

[161] Hauselmann HJ, Fernandes RJ, Mok SS, Schmid TM, Block JA, Aydelotte MB, et al. Phenotypic stability of bovine articular chondrocytes after long-term culture in alginate beads. Journal of Cell Science. 1994;107:17-27.

[162] Hauselmann HJ, Masuda K, Hunziker EB, Neidhart M, Mok SS, Michel BA, et al. Adult human chondrocytes cultured in alginate form a matrix similar to native human articular cartilage. American Journal of Physiology - Cell Physiology. 1996;271:C742-C52.

[163] Kiefer MC, Stephans JC, Crawford K, Okino K, Barr PJ. Ligand-affinity cloning and structure of a cell surface heparan sulfate proteoglycan that binds basic fibroblast growth factor. Proceedings of the National Academy of Sciences. 1990;87:6985-9.

[164] Rodgers KD, San Antonio JD, Jacenko O. Heparan sulfate proteoglycans: A GAGgle of skeletal-hematopoietic regulators. Developmental Dynamics. 2008;237:2622-42.

[165] Mariappan MR, Alas EA, Williams JG, Prager MD. Chitosan and chitosan sulfate have opposing effects on collagen–fibroblast interactions. Wound Repair and Regeneration. 1999;7:400-6.

[166] Freeman I, Kedem A, Cohen S. The effect of sulfation of alginate hydrogels on the specific binding and controlled release of heparin-binding proteins. Biomaterials. 2008;29:3260-8.

[167] Re’em T, Kaminer-Israeli Y, Ruvinov E, Cohen S. Chondrogenesis of hMSC in affinity-bound TGF-beta scaffolds. Biomaterials. 2012;33:751-61.

[168] Fan L, Jiang L, Xu Y, Zhou Y, Shen Y, Xie W, et al. Synthesis and anticoagulant activity of sodium alginate sulfates. Carbohydrate Polymers. 2011;83:1797-803.

[169] Ronghua H, Yumin D, Jianhong Y. Preparation and in vitro anticoagulant activities of alginate sulfate and its quaterized derivatives. Carbohydrate Polymers. 2003;52:19-24.

[170] Knudson CB, Knudson W. Cartilage proteoglycans. Seminars in Cell & Developmental Biology. 2001;12:69-78.

References

120

[171] Murray G. Anticoagulant therapy with heparin. The American Journal of Medicine. 1947;3:468-71.

[172] Andersson LO, Barrowcliffe TW, Holmer E, Johnson EA, Sims GEC. Anticoagulant properties of heparin fractionated by affinity chromatography on matrix-bound antithrombin III and by gel filtration. Thrombosis Research. 1976;9:575-83.

[173] Galtrey CM, Fawcett JW. The role of chondroitin sulfate proteoglycans in regeneration and plasticity in the central nervous system. Brain Research Reviews. 2007;54:1-18.

[174] Campo GM, Avenoso A, Campo S, Ferlazzo AM, Micali C, Zanghı ̀L, et al. Hyaluronic acid and chondroitin-4-sulphate treatment reduces damage in carbon tetrachloride-induced acute rat liver injury. Life Sciences. 2004;74:1289-305.

[175] Freeman I, Cohen S. The influence of the sequential delivery of angiogenic factors from affinity-binding alginate scaffolds on vascularization. Biomaterials. 2009;30:2122-31.

[176] Kim KS, Lee JW, Cho SH. Anticoagulation Activities of Low Molecular Weight Sulfated Chitosan and Sulfated Sodium Alginate. Polymer Korea. 2003;27:583-8.

[177] Kasai Y, Akahira A, Kakuta S, Abudula A, Urayama K, Takigawa T. Preparation and Electrochemical Properties of Alginate Sulfate Electrolyte Membranes. Kobunshi Ronbunshu. 2008;65.

[178] Hintze V, Moeller S, Schnabelrauch M, Bierbaum S, Viola M, Worch H, et al. Modifications of Hyaluronan Influence the Interaction with Human Bone Morphogenetic Protein-4 (hBMP-4). Biomacromolecules. 2009;10:3290-7.

[179] Becher J, Möller S, Riemer T, Schiller J, Hintze V, Bierbaum S, et al. Sulfated Glycosaminoglycan Building Blocks for the Design of Artificial Extracellular Matrices. ACS Symposium Series. 2012;1107:315-28.

[180] Wang G, Woods A, Sabari S, Pagnotta L, Stanton L-A, Beier F. RhoA/ROCK Signaling Suppresses Hypertrophic Chondrocyte Differentiation. Journal of Biological Chemistry. 2004;279:13205-14.

[181] Phillips JA, Bonassar LJ. Matrix metalloproteinase activity synergizes with α2β1 integrins to enhance collagen remodeling. Experimental Cell Research. 2005;310:79-87.

[182] Yamaoka H, Asato H, Ogasawara T, Nishizawa S, Takahashi T, Nakatsuka T, et al. Cartilage tissue engineering using human auricular chondrocytes embedded in different hydrogel materials. Journal of Biomedical Materials Research Part A. 2006;78A:1-11.

[183] van Susante JLC, Buma P, van Osch GJVM, Versleyen D, van der Kraan PM, van der Berg WB, et al. Culture of chondrocytes in alginate and collagen carrier gels. Acta Orthopaedica. 1995;66:549-56.

[184] Johnstone B, Hering TM, Caplan AI, Goldberg VM, Yoo JU. In VitroChondrogenesis of Bone Marrow-Derived Mesenchymal Progenitor Cells. Experimental Cell Research. 1998;238:265-72.

[185] Sakaguchi Y, Sekiya I, Yagishita K, Muneta T. Comparison of human stem cells derived from various mesenchymal tissues: Superiority of synovium as a cell source. Arthritis & Rheumatism. 2005;52:2521-9.

[186] De Bari C, Dell'Accio F, Tylzanowski P, Luyten FP. Multipotent mesenchymal stem cells from adult human synovial membrane. Arthritis & Rheumatism. 2001;44:1928-42.

Chapter 9

121

[187] Erickson GR, Gimble JM, Franklin DM, Rice HE, Awad H, Guilak F. Chondrogenic Potential of Adipose Tissue-Derived Stromal Cells in Vitro and in Vivo. Biochemical and Biophysical Research Communications. 2002;290:763-9.

[188] Awad HA, Quinn Wickham M, Leddy HA, Gimble JM, Guilak F. Chondrogenic differentiation of adipose-derived adult stem cells in agarose, alginate, and gelatin scaffolds. Biomaterials. 2004;25:3211-22.

[189] Ma H-L, Hung S-C, Lin S-Y, Chen Y-L, Lo W-H. Chondrogenesis of human mesenchymal stem cells encapsulated in alginate beads. Journal of Biomedical Materials Research Part A. 2003;64A:273-81.

[190] Mauck RL, Yuan X, Tuan RS. Chondrogenic differentiation and functional maturation of bovine mesenchymal stem cells in long-term agarose culture. Osteoarthritis and Cartilage. 2006;14:179-89.

[191] Wang Y, Kim U-J, Blasioli DJ, Kim H-J, Kaplan DL. In vitro cartilage tissue engineering with 3D porous aqueous-derived silk scaffolds and mesenchymal stem cells. Biomaterials. 2005;26:7082-94.

[192] Cho JH, Kim S-H, Park KD, Jung MC, Yang WI, Han SW, et al. Chondrogenic differentiation of human mesenchymal stem cells using a thermosensitive poly(N-isopropylacrylamide) and water-soluble chitosan copolymer. Biomaterials. 2004;25:5743-51.

[193] Hwang NS, Varghese S, Zhang Z, Elisseeff J. Chondrogenic differentiation of human embryonic stem cell-derived cells in arginine-glycine-aspartate-modified hydrogels. Tissue engineering. 2006;12:2695-706.

[194] Zhou G, Liu W, Cui L, Wang X, Liu T, Cao Y. Repair of porcine articular osteochondral defects in non-weightbearing areas with autologous bone marrow stromal cells. Tissue engineering. 2006;12:3209-21.

[195] Lin C-C, Anseth KS. Controlling Affinity Binding with Peptide-Functionalized Poly(ethylene glycol) Hydrogels. Advanced Functional Materials. 2009;19:2325-31.

[196] Zhu J. Bioactive modification of poly(ethylene glycol) hydrogels for tissue engineering. Biomaterials. 2010;31:4639-56.

[197] Lutolf MP, Lauer-Fields JL, Schmoekel HG, Metters AT, Weber FE, Fields GB, et al. Synthetic matrix metalloproteinase-sensitive hydrogels for the conduction of tissue regeneration: Engineering cell-invasion characteristics. Proceedings of the National Academy of Sciences. 2003;100:5413-8.

[198] Hwang NS, Varghese S, Elisseeff J. Controlled differentiation of stem cells. Advanced Drug Delivery Reviews. 2008;60:199-214.

[199] Wilson A, Trumpp A. Bone-marrow haematopoietic-stem-cell niches. Nat Rev Immunol. 2006;6:93-106.

[200] Benoit DSW, Schwartz MP, Durney AR, Anseth KS. Small functional groups for controlled differentiation of hydrogel-encapsulated human mesenchymal stem cells. Nat Mater. 2008;7:816-23.

[201] Salinas CN, Anseth KS. The influence of the RGD peptide motif and its contextual presentation in PEG gels on human mesenchymal stem cell viability. Journal of Tissue Engineering and Regenerative Medicine. 2008;2:296-304.

References

122

[202] Lin Y-C, Brayfield CA, Gerlach JC, Peter Rubin J, Marra KG. Peptide modification of polyethersulfone surfaces to improve adipose-derived stem cell adhesion. Acta Biomaterialia. 2009;5:1416-24.

[203] Liu SQ, Tian Q, Wang L, Hedrick JL, Hui JHP, Yang YY, et al. Injectable Biodegradable Poly(ethylene glycol)/RGD Peptide Hybrid Hydrogels for in vitro Chondrogenesis of Human Mesenchymal Stem Cells. Macromolecular Rapid Communications. 2010;31:1148-54.

[204] Salinas CN, Anseth KS. The enhancement of chondrogenic differentiation of human mesenchymal stem cells by enzymatically regulated RGD functionalities. Biomaterials. 2008;29:2370-7.

[205] Chiu L-H, Chen S-C, Wu K-C, Yang C-B, Fang C-L, Lai W-FT, et al. Differential effect of ECM molecules on re-expression of cartilaginous markers in near quiescent human chondrocytes. Journal of Cellular Physiology. 2011;226:1981-8.

[206] Wojtowicz AM, Shekaran A, Oest ME, Dupont KM, Templeman KL, Hutmacher DW, et al. Coating of biomaterial scaffolds with the collagen-mimetic peptide GFOGER for bone defect repair. Biomaterials. 2010;31:2574-82.

[207] Raynor JE, Petrie TA, García AJ, Collard DM. Controlling Cell Adhesion to Titanium: Functionalization of Poly[oligo(ethylene glycol)methacrylate] Brushes with Cell-Adhesive Peptides. Advanced Materials. 2007;19:1724-8.

[208] Reyes CD, Petrie TA, Burns KL, Schwartz Z, García AJ. Biomolecular surface coating to enhance orthopaedic tissue healing and integration. Biomaterials. 2007;28:3228-35.

[209] Connelly J, Petrie T, García A, Levenston M. Fibronectin-and collagen-mimetic ligands regulate bone marrow stromal cell chondrogenesis in three-dimensional hydrogels. Eur Cell Mater. 2011;22:168-77.

[210] Liu SQ, Tian Q, Hedrick JL, Po Hui JH, Rachel Ee PL, Yang YY. Biomimetic hydrogels for chondrogenic differentiation of human mesenchymal stem cells to neocartilage. Biomaterials. 2010;31:7298-307.

[211] Jongpaiboonkit L, King WJ, Murphy WL. Screening for 3D environments that support human mesenchymal stem cell viability using hydrogel arrays. Tissue engineering Part A. 2009;15:343-53.

[212] Nuttelman CR, Tripodi MC, Anseth KS. Synthetic hydrogel niches that promote hMSC viability. Matrix Biology. 2005;24:208-18.

[213] Von Der Mark K, Gauss V, Von Der Mark H, Muller P. Relationship between cell shape and type of collagen synthesised as chondrocytes lose their cartilage phenotype in culture. Nature. 1977;267:531-2.

[214] Coates EE, Fisher JP. Phenotypic variations in chondrocyte subpopulations and their response to in vitro culture and external stimuli. Annals of Biomedical Engineering. 2010;38:3371-88.

[215] Aydelotte MB, Kuettner KE. Differences between sub-populations of cultured bovine articular chondrocytes. I. Morphology and cartilage matrix production. Connective Tissue Research. 1988;18:205-22.

Chapter 9

123

[216] Schumacher BL, Hughes CE, Kuettner KE, Caterson B, Aydelotte MB. Immunodetection and partial cDNA sequence of the proteoglycan, superficial zone protein, synthesized by cells lining synovial joints. Journal of Orthopaedic Research. 1999;17:110-20.

[217] Flannery CR, Hughes CE, Schumacher BL, Tudor D, Aydelotte MB, Kuettner KE, et al. Articular Cartilage Superficial Zone Protein (SZP) Is Homologous to Megakaryocyte Stimulating Factor Precursor and Is a Multifunctional Proteoglycan with Potential Growth-Promoting, Cytoprotective, and Lubricating Properties in Cartilage Metabolism. Biochemical and Biophysical Research Communications. 1999;254:535-41.

[218] Khalafi A, Schmid TM, Neu C, Reddi AH. Increased accumulation of superficial zone protein (SZP) in articular cartilage in response to bone morphogenetic protein-7 and growth factors. Journal of Orthopaedic Research. 2007;25:293-303.

[219] Neu CP, Khalafi A, Komvopoulos K, Schmid TM, Reddi AH. Mechanotransduction of bovine articular cartilage superficial zone protein by transforming growth factor β signaling. Arthritis & Rheumatism. 2007;56:3706-14.

[220] Lee SY, Niikura T, Reddi AH. Superficial zone protein (lubricin) in the different tissue compartments of the knee joint: modulation by transforming growth factor beta 1 and interleukin-1 beta. Tissue engineering Part A. 2008;14:1799-808.

[221] Wang JG, Miyazu M, Xiang P, Li SN, Sokabe M, Naruse K. Stretch-induced cell proliferation is mediated by FAK-MAPK pathway. Life Sciences. 2005;76:2817-25.

[222] Buschmann MD, Gluzband YA, Grodzinsky AJ, Hunziker EB. Mechanical compression modulates matrix biosynthesis in chondrocyte/agarose culture. Journal of Cell Science. 1995;108:1497-508.

[223] De Croos JNA, Dhaliwal SS, Grynpas MD, Pilliar RM, Kandel RA. Cyclic compressive mechanical stimulation induces sequential catabolic and anabolic gene changes in chondrocytes resulting in increased extracellular matrix accumulation. Matrix Biology. 2006;25:323-31.

[224] Hansen U, Schünke M, Domm C, Ioannidis N, Hassenpflug J, Gehrke T, et al. Combination of reduced oxygen tension and intermittent hydrostatic pressure: a useful tool in articular cartilage tissue engineering. Journal of Biomechanics. 2001;34:941-9.

[225] Saini S, Wick TM. Effect of low oxygen tension on tissue-engineered cartilage construct development in the concentric cylinder bioreactor. Tissue engineering. 2004;10:825-32.

[226] Wernike E, Li Z, Alini M, Grad S. Effect of reduced oxygen tension and long-term mechanical stimulation on chondrocyte-polymer constructs. Cell Tissue Res. 2008;331:473-83.

[227] Klein TJ, Schumacher BL, Blewis ME, Schmidt TA, Voegtline MS, Thonar EJ, et al. Tailoring secretion of proteoglycan 4 (PRG4) in tissue-engineered cartilage. Tissue engineering. 2006;12:1429-39.

[228] Fonseca KB, Bidarra SJ, Oliveira MJ, Granja PL, Barrias CC. Molecularly designed alginate hydrogels susceptible to local proteolysis as three-dimensional cellular microenvironments. Acta Biomaterialia. 2011;7:1674-82.

References

124

125

Curriculum Vitae

Name Rami Mhanna

Date of birth 05. 05. 1982

Nationality Lebanese

Present address Albisstrasse 96, CH-8038 Zurich, Switzerland

Education 2009 – Present PhD Candidate, Swiss Federal Institute of Technology Zurich

(ETH Zürich)

Supervised by Prof. Marcy Zenobi-Wong and Prof. Janos Vörös

Research Topics:

• The effect of the microenvironment on 3D cultured cartilage cells

• The response of cartilage cells to mechanical stimulation • Layer-by-layer films of extracellular matrix molecules • Functionalization of magnetic micro-robots for drug and gene

delivery

2007 – 2008 Masters of Biomedical Engineering, University of Melbourne

• Research Project: On-Chip Emulsions for Drug Delivery Applications, Supervised by Dr. Brigitte Stadler (Score: 89/100)

2001 – 2006 Bachelor of Engineering, Notre Dame University, Lebanon

• Major: Computer and Communication • Senior Project: Wireless Messaging System (Score: 92/100)

Employment History Jan – March 2008 Research Assistant, Department of Chemical and Biomolecular

Engineering,

126

July – Aug 2008 University of Melbourne, Nanostructured Interfaces and Materials Group, Prof. Frank Caruso

• Formation of on-chip emulsions for drug delivery applications • Drug delivery capsules using the layer-by-layer technique

Nov – Dec 2007 Research Assistant, University of Melbourne, Bio21 Institute, Dr. Sally Gras

• Studying the kinetics of collagen fibril formation

2006 – 2007 Technical IT Assistant, Full Time, Telnet, Lebanon

• Worked with a team of engineers in installation of computers, printers, projectors, servers, LANs, WANs and VOIP gateways

• Managed office work and task management for employees

1998 – 2001 Technical Electrician, Part Time, Rashaya Electric, Lebanon

• Construction of PCBs, transformers, UPSs and electrical stabilizers

Publications

Journal Publications:

• Mhanna R, Öztürk E, Vallmajo Martin Q, Millan C, Müller M and Zenobi-Wong M. Biomimetic polyethylene glycol hydrogels functionalized with GFOGER peptides induce chondrogenic differentiation of human mesenchymal stem cells. In preparation • Mhanna R, Kashyap A, Palazzolo G, and Zenobi-Wong M. Chondrocyte Culture in 3D Alginate Sulfate Hydrogels Promotes Proliferation While Maintaining Expression of Chondrogenic Markers. Submitted to Biomaterials • Mhanna R, Öztürk E, Schlink P, and Zenobi-Wong M. Probing the microenvironmental conditions for induction of superficial zone protein expression. Submitted to Osteoarthritis and Cartilage • Rottmar M, Mhanna R, Vogel V, Zenobi-Wong M, and Maniura-Weber K. Interference with the contractile machinery of the fibroblastic chondrocyte cytoskeleton induces re-expression of the cartilage phenotype. Submitted to Journal of Experimental Cell Research • Sugihara K, Delai M, Mhanna R, Kusch J, Poulikakos D, Vörös J, Zambelli T, and Ferrari A. Label-free detection of cell-contractile activity with lipid nanotubes. Integrative Biology 2013;5(2):423-30 • Mhanna RF, Vörös J, Zenobi-Wong M. Layer-by-layer films made from extracellular matrix macromolecules on silicone substrates. Biomacromolecules 2011; 12: 609-616

127

Patents: • Palazzolo G, Mhanna R, Becher J, Möller S, Schnabelrauch M, and Zenobi-Wong M. Modified alginate hydrogels for tissue engineering and regenerative medicine, European Patent EP 12007934.8, November 19, 2012

Conference Contributions:

• Mhanna R, Vallmajo Martin Q, Öztürk E, Millan C, Müller M, and Wong M. Functionalized biomimetic polyethylene glycol hydrogels utilizing GFOGER peptides for cartilage tissue engineering. 3rd TERMIS World Conference, September 2012, Vienna, Austria

• Qiu F, Mhanna R, Zhang L, Ding Y, Tottori S, Sugihara K, Zenobi-Wong M, and Nelson BJ. Artificial bacterial flagella functionalized with liposomes for biomedical applications. 7th MRC Graduate Symposium, June 2012, Zurich, Switzerland • Mhanna RF, Schlink P, Vörös J, and Marcy Wong. The effect of chondrocyte morphology on their response to mechanical compression. Termis NA Conference, December 2011, Houston, USA • Sugihara K, Delai M, Stuck J, Mhanna R, Ferrari A, Vörös J, Zambelli T. The directed-assembly of lipid nanotubes from inverted-hexagonal structures. 2nd Nanotoday Conference, December 2011, Hawaii, USA • Sugihara K, Delai M, Stuck J, Mhanna R, Ferrari A, Vörös J, Zambelli T. Lipids on polyelectrolytes for biosensing and biomaterials. Biological Surfaces and Interfaces - ESF EMBO, July 2011, Sant Feliu de Guixols, Spain • Rottmar M, Mhanna RF, Wong M and Maniura K. Cell shape versus cytoskeleton integrity: probing the re-expression of the chondrogenic phenotype. Biological Surfaces and Interfaces - ESF EMBO, July 2011, Sant Feliu de Guixols, Spain • Mhanna RF, Schlink P, Vörös J, and Marcy Wong. The effect of chondrocyte morphology on their response to mechanical compression. Biological Surfaces and Interfaces - ESF EMBO, July 2011, Sant Feliu de Guixols, Spain • Mhanna RF, Vörös J, and Wong M. Build-up of layer-by-layer films of extracellular matrix molecules on PDMS. 8th International Symposium on Polyelectrolytes, November 2010, Shanghai, China • Mhanna RF, Schlink P, Vörös J, and Wong M. A new design for 3D loading of cells using controlled alginate gelation. 16th Swiss Conference on Biomaterials - SSB 2010 , May 2010, Zürich, Switzerland • Mhanna RF, Vörös J and Wong M. Layer-by-layer assemblies of extracellular matrix molecules and their effect on loaded chondrocytes. The Swiss Society for Biomedical Engineering Annual Meeting - SSBE 2009 , August 2009, Bern, Switzerland • Mhanna RF, Vörös J and Wong M. Chondrocyte response to tensile strain on layer-by-layer films made from extracellular matrix molecules. 22nd European Conference on Biomaterials - ESB 2009, August 2009, Lausanne, Switzerland

128

Research Mentoring

March – Nov 2012 Swiss Federal Institute of Technology Zurich (ETH Zürich)

Aditya Kashyap, Semester Project

Title: Effect of alginate sulfation on the cartilage phenotype

March – Nov 2012 Queralt Vallmajo Martin, Semester Project

Title: Functionalized PEG hydrogels for cartilage tissue engineering

Sept 2009 – Philippe Schlink, Master Thesis

March 2010 Title: New techniques for application of 3D strains on chondrocytes