374
Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid Bio/Synthetic Biomimetic Aggrecan Macromolecule A Thesis Submitted to the Faculty of Drexel University by Sumona Sarkar in partial fulfillment of the requirements for the degree of Doctor of Philosophy September 2011

Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

Embed Size (px)

DESCRIPTION

PhD Thesis by Sumona Sarkar of Drexel University

Citation preview

Page 1: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid

Bio/Synthetic Biomimetic Aggrecan Macromolecule

A Thesis

Submitted to the Faculty

of

Drexel University

by

Sumona Sarkar

in partial fulfillment of the

requirements for the degree

of

Doctor of Philosophy

September 2011

Page 2: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

© Copyright 2011 Sumona Sarkar. All Rights Reserved.

Page 3: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

ii

Dedications

This work is dedicated to my parents, Ashoke and Namita Sarkar. They have made my

education their first priority, making many sacrifices for me to reach this point. My

Mother has also been an inspiration for my work on this project. Her daily struggle with

low back pain has been a continuous reminder to me of the importance of this work and

its potential benefit in improving the lives of people like her.

Page 4: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

iii

Acknowledgments

There are so many people whose support has made this thesis possible. Firstly I

would like to convey my sincere appreciation to my advisor Dr. Michele Marcolongo.

Her dedication to bringing up the next generations of researchers showed through in

every interaction we had. I have benefited from her guidance in every aspect of this

journey, from project planning to life planning. The number and diversity of

opportunities she has made available to me has made me it possible for me to feel

confident in my new and continuing roles as researcher, mentor and teacher. It is rare to

find such a kind, understanding and strong advisor, and I only hope that I can maintain

the high standards she has set in my future endeavors.

I also thank my committee members Dr. Caroline Schauer, Dr. Edward

Vresilovic, Dr. Kenneth Barbee, Dr. Margaret Wheatley, Dr. Sriram Balasubramanian,

and Dr. Elisabeth Papazoglou for their guidance and support throughout the project. I

especially thank Dr. Vresilovic for his long conversations which helped me to push

forward the research and think outside of the box. I Also thank Dr. Schauer and Dr. Lynn

Penn for patiently introducing me to organic chemistry and pushing me to go beyond my

insecurities in the field. I thank Dolores Conver, Amy Peterson and James Throckmorton

of the Palmses Lab, Chris Winkler of the Taheri Lab, Alex Radin (University of

Pennsylvania) and Joanna McGough of the Risbud Lab (Thomas Jefferson University)

for technical support. I thank the Coulter Foundation for funding for this work.

Next I would like express my unending gratitude to my family, without whose

support I could not have completed this thesis. To my Mother, Namita Sarkar and my

Page 5: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

iv

Father, Ashoke Sarkar who have taught me the value of education and who have always

worked hard to make sure I could pursue my ambitions. To my Aunt and my Uncle who

have stood by me in every pursuit and given me the kind of love and support that I hope

to be able to give to others. To my cousins Soma Kalb and Keya Sau who have served as

wonderful role models for me.

I would especially like to thank Soma for introducing me to biomedical

engineering research and inviting me to have my first research experience in her graduate

lab. My summer at Duke University, working with Soma and her colleagues sparked my

passion for research and motivated me to continue in this field. I also thank Marco

Cannella, my “unintended consequence”. His support from my first day in the lab has

been something I could count on, no matter what the circumstances. Many of my most

critical moments through the degree were resolved after heartfelt conversations with

Marco, I cannot thank him enough.

I would like to express my sincere gratitude to my friends and colleagues at

Drexel University. I thank Valerie Binetti for engaging and motivating technical

conversations which helped me to improve upon my investigations. I also thank her for

her friendship, without which I could not have navigated the graduate experience. I also

thank Chris Massey for being a wonderful collaborator in the lab and friend outside of the

lab. He welcomed me into the lab on my first day and kept the ambiance lighthearted

even on the hardest days. I thank the members of the Biomaterials Lab in the Materials

Science and Engineering Department. In particular I thank Rob Yucha for his daily

support and friendship both in and outside of the lab. I also thank the many students who

have contributed to this project, particularly Nandita Ganesh, Luis Mejias and Sarah

Page 6: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

v

Lightfoot Vidal. I would also like to thank Drs. Mark and Dianne Rothstein for inviting

me to their lab in Prime Synthesis and teaching me organic chemistry techniques and

allowing me to begin my research studies in their lab. I also would like to express my

sincere appreciation to my previous advisors Drs. Tejal Desai and Joyce Wong who saw

the potential in me and gave me a chance to pursue this career path. I also thank the staff

of the School of Biomedical Engineering, Science and Health Systems, and the staff of

the Materials Science and Engineering Department for all of their help in managing the

logistics of graduate school.

Page 7: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

vi

Table of Contents

List of Tables ................................................................................................................... xiii

List of Figures ................................................................................................................... xv

Abstract .......................................................................................................................... xxiii

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

Chapter 2 : Significance ...................................................................................................... 4

Chapter 3 : Background ..................................................................................................... 8

3.1. Anatomy and Physiology of the Spine ................................................................. 8

3.1.1. The Vertebral Column................................................................................... 8

3.1.2. The Intervertebral Disc ................................................................................. 8

3.1.3. The Nucleus Pulposus ................................................................................... 9

3.1.4. The Annulus Fibrosus ................................................................................... 9

3.2. Intervertebral Disc Degeneration ...................................................................... 15

3.3. Treatment Options .................................................................................................. 20

3.3.1. Spinal fusion ................................................................................................... 22

3.3.2. Total Disc Arthroplasty................................................................................... 23

3.3.3. Nucleus Replacement...................................................................................... 24

3.3.4. Tissue Engineering and Regenerative Medicine Approaches to Disc

Degeneration ............................................................................................................. 27

Page 8: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

vii

3.2.4. Summary of current treatment options: a need for an early interventional

strategy ...................................................................................................................... 29

3.4. Aggrecan structure, assembly, and function .......................................................... 34

3.4.1. Proteoglycan Diversity, Structure and Function ............................................. 34

3.4.2. Aggrecan Structure. ........................................................................................ 35

3.4.3. Aggrecan In Vivo Synthesis. .......................................................................... 35

3.4.4. Functional molecular biomechanics of aggrecan ............................................ 36

3.5. Aggrecan Degradation ........................................................................................... 41

3.7. Chondroitin Sulfate ................................................................................................ 47

3.7.1. Localizations and diverse functions of CS...................................................... 47

3.7.2. Conformation and functional mechanics of CS .............................................. 48

3.7.3. Chondroitin Sulfate Stability in Aqueous Solutions ....................................... 50

3.7.4. Age related changes in NP GAG .................................................................... 51

3.7.5. Pharmaceutical application. ............................................................................ 52

3.8. Chondroitin Sulfate networks ................................................................................. 59

3.9. Synthetic bottle brush polymers and aggrecan mimetics ....................................... 61

3.10. Glycopolymers ...................................................................................................... 65

3.11. Summary ............................................................................................................... 68

Chapter 4 : Objective and Specific Aims ......................................................................... 70

Page 9: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

viii

Chapter 5 : Identification of a Chondroitin Sulfate Terminal Handle and the Investigation

of its Utility in the Fabrication of Biomimetic Brush Structures ...................................... 75

5.1. Introduction ........................................................................................................... 75

5.2. Materials and Methods .......................................................................................... 79

5.2.1. Determination of CS primary amine content .................................................. 80

5.2.4. Modification of glass surfaces with Chondroitin sulfate ................................ 82

5.2.5. Characterization of chondroitin sulfate modified surfaces ............................. 83

5.2.6. Chondroitin sulfate-monomer conjugation ..................................................... 83

5.2.7. Proton nuclear magnetic resonance (1H-NMR) and Fourier transform infrared

spectroscopy (FTIR) analysis of AGE-CS conjugation over time ............................ 85

5.2.8. Conjugation of Chondroitin Sulfate to a poly(acrylic acid) polymeric

backbone ................................................................................................................... 85

5.3. Results ................................................................................................................... 94

5.3.1.Primary amine content of CS from various sources ........................................ 94

5.3.2. Attachment of CS to amine-reactive substrates for the “grafting- to” synthesis

strategy ...................................................................................................................... 96

5.3.3. Chondroitin sulfate conjugation to synthetic monomers via the terminal

primary amine. .......................................................................................................... 98

5.3.4. Conjugation of CS to a polymeric backbone for CS-glycopolymer synthesis

via the “grafting-to” strategy .................................................................................. 102

5.4. Discussion ........................................................................................................... 130

Page 10: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

ix

5.4.1. CS Terminal Primary Amine ........................................................................ 130

5.4.2. Surface grafting of CS via the terminal primary amine ................................ 132

5.4.3. Reactivity of the CS terminal primary amine and attachment of a

polymerizable monomer ......................................................................................... 135

5.4.4. Grafting of CS to Poly(acrylic acid) ............................................................. 143

5.5. Conclusion ........................................................................................................... 152

Chapter 6 : Investigation of the Epoxide-CS Primary Amine Reaction for the Synthesis

of CS Brush Structures via the “Grafting-Through” Technique ..................................... 154

6.1 Introduction ........................................................................................................... 154

6.2. Background ......................................................................................................... 155

6.2.1. Epoxide Reactivity ........................................................................................ 155

6.2.2. The Epoxide-Amine Reaction....................................................................... 156

6.2.3. Step-growth polymerization of epoxides and amines ................................... 158

6.3 Methods ................................................................................................................. 168

6.3.1. Reaction of CS to PEG based di-epoxides .................................................... 168

6.3.2. Purification of the CS-DGE reaction ............................................................ 168

6.3.3. Characterization of CS-macromolecule chemical structure ......................... 169

6.3.3. Gel Permeation Chromatography of CS-Macromolecules ........................... 170

6.4 Results ................................................................................................................... 181

Page 11: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

x

6.4.1. Influence of temperature, DGE concentration and DGE molecular weight on

CS amine-DGE reaction kinetics ............................................................................ 181

6.4.2. Purification of CS-macromolecules .............................................................. 183

6.4.3. 1H-NMR and ATR-FTIR analysis of CS-Macromolecule chemical structure

................................................................................................................................. 184

6.4.4. Molecular weight characterization of CS-macromolecules .......................... 186

6.4.5. Stability of amine and epoxide reactants ...................................................... 192

6.4.6. Reactivity of DGEs to the serine amine ........................................................ 195

6.5 Discussion ............................................................................................................. 221

6.5.1. Reaction kinetics of the CS-DGE reaction ................................................... 221

6.5.2. Purification of synthesized CS-Macromolecules .......................................... 225

6.5.3. Chemical structure of the CS-Macromolecules ............................................ 227

6.5.4. CS-Macrmolecule molecular weight distribution ......................................... 229

6.6 Conclusions ........................................................................................................... 237

Chapter 7 : Characterization of CS Macromolecules synthesized via the Epoxy-Amine

Step-Growth “Grafting-Through” technique of Polymer Synthesis ............................... 238

7.1 Introduction ........................................................................................................... 238

7.2 Methods ................................................................................................................. 242

7.2.1. Transmission Electron Microscopy (TEM) Imaging of PEG-CS

Macromolecules ...................................................................................................... 242

7.2.2. Viscosity of CS-macromolecule solutions .................................................... 243

Page 12: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xi

7.2.3. Fixed Charge Density of CS-Macromolecules ............................................. 244

7.2.4. Osmotic pressure of CS-macromolecule solutions ....................................... 245

7.2.4. Fibroblast and Nucleus Pulposus Cytotoxicity studies ................................. 246

7.3 Results ................................................................................................................... 253

7.3.1. TEM Imaging of PEG-CS Macromolecules ................................................. 253

7.3.2. Viscosity of CS-macromolecule solutions. ................................................... 253

7.3.3. Fixed Charge Density of CS-macromolecules. ............................................. 254

7.3.4. Osmotic Pressure of CS-Macromolecule Solutions. ..................................... 254

7.3.5. Cytotoxicity of DGE monomers and CS-macromolecules. .......................... 256

7.4 Discussion ............................................................................................................. 267

7.4.1. Physical structure of CS-macromolecules .................................................... 267

7.4.2. Osmotic Function and Fixed Charge Density of CS-macromolecules ......... 269

7.4.2. Cytotoxicity of CS-macromolecules ............................................................. 275

7.5 Conclusions ........................................................................................................... 283

Chapter 8 : Conclusions ................................................................................................. 285

8.1 Summary ............................................................................................................... 285

8.2 Novel Contributions .............................................................................................. 289

Chapter 9 : Future Work and Recommendations ........................................................... 293

9.1 Recommendations for Biomimetic Aggrecan Synthesis ....................................... 293

9.1.1. “Grafting-To” Synthesis Strategies............................................................... 293

Page 13: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xii

9.1.2. Step-Growth “Grafting-Through” Synthesis Strategies ................................ 294

9.1.3. Alternative “Grafting-Through” Synthesis Strategies: Free-Radical

Polymerization ........................................................................................................ 295

9.1.4. Alternative Biomimetic Aggrecan Structures ............................................... 296

9.2 Recommendations for Biomimetic Aggrecan Characterization............................ 296

9.3 Recommendations for Biomimetic Aggrecan Functional Characterization ......... 297

Appendix 1: Select Literature Review of Key Relevant Technologies ......................... 301

Appendix 2: Alternate views of Chondroitin Sulfate Disaccharide with Linkage Region

and Terminal Amino Acid Residue (demonstrating glycosidic linkages) ...................... 321

List of References ........................................................................................................... 322

Vita .................................................................................................................................. 347

Page 14: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xiii

List of Tables

Table 3.1: Elastomeric Nucleus Pulposus Replacement Devices .................................... 33 Table 3.2: Commonly found chondroitin sulfate isomers ............................................... 57 Table 3.3: Chondroitin sulfate proteoglycans and their function in tendon (90) ............. 57 Table 3.4: Pros and cons of the different routes to bottle brush synthesis ........................ 64 Table 3.5: Key functions of aggrecan and its components .............................................. 69 Table 4.1: Design of Biomimetic Aggrecan .................................................................... 74 Table 5.1: Product Information for Chondroitin Sulfate Samples ................................... 93 Table 5.2: Estimated Number of Primary Amines per Molecule for Control Samples . 128 Table 5.3: Estimated Number of Primary Amines per Molecule for Chondroitin Sulfate Samples ........................................................................................................................... 128 Table 5.4: Theoretical estimation of number of CS chains attached per PAA chain .... 129 Table 5.5: Chain Polymerization Strategies utilized in the fabrication of glycopolymers

......................................................................................................................................... 150 Table 5.6: Estimated cost of CS-PAA macromolecule synthesis .................................. 151 Table 6.1: Poly(ethylene glycol)-diglycodyl ethers of varying molecular weight ........ 179 Table 6.2: Factors involved in GPC and Key Considerations (201) .............................. 180 Table 6.3: Summary of (A) Rate Constants and (B) % Conjugation Plateaus for CS Primary Amine Reaction to DGE of Varying MW ........................................................ 219 Table 6.4: Table of GPC Sample Loading Parameters .................................................. 219 Table 6.5: Molecular Weight distributions calculated relative to polysaccharide standards

......................................................................................................................................... 220 Table 6.6: Molecular weight distributions calculated relative to polystyrene sulfonate standards ......................................................................................................................... 220 Table 7.1: Predicted Restorative Masses of CS-macromolecules and Natural Aggrecan based on FCD of degenerated NP tissues (84, 270-271) ................................................ 281

Page 15: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xiv

Table 7.2: Estimated Cost of CS-macromolecule for therapeutic application in comparison to estimated cost of restoration with natural aggrecan and current treatment option of steroid injections. ............................................................................................. 282 Table 8.1: Properties of “Grafting-to” and “Grafting-through” CS-macromolecules as they relate to the criteria for a functional biomimetic aggrecan macromolecule ........... 291

Page 16: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xv

List of Figures

Figure 2.1: Prevalence of self-reported joint pain ............................................................. 6 Figure 2.2: Lumbar fusion rates by primary diagnosis.(6) ................................................ 6 Figure: 2.3: Seven year trend in total cost for spinal fusion procedures in the United (1) 7 Figure 3.1: Regions of the human spine .......................................................................... 11 Figure 3.2: The Intervertebral disc location ..................................................................... 12 Figure 3.3: Constituents of the nucleus pulposus and subsequent development of hydrostatic pressure (7) ..................................................................................................... 13 Figure 3.4: Load transfer in the IVD and lamellar structure if the annulus fibrosus. (20)

........................................................................................................................................... 14 Figure 3.5: Disc degeneration and mechanical consequences ........................................ 17 Figure 3.6: Changes in major biomacromolecule concentrations in the annulus fibrosus and nucleus pulposus of the intervertebral disc with aging (3, 24-26) ............................. 18 Figure 3.7: Change in water content in the IVD and associated change in glycosaminoglycan content in the IVD with aging and degeneration. (30) ..................... 19 Figure 3.8: SB Charite` III, and radiographic image of the implanted device (41). ........ 31 Figure 3.9: PDN-SOLOTM device in hydrated and dehydrated states. ............................ 31 Figure 3.10: Photograph of the NuCoreTM (Spine Wave) Injectable Disc Nucleus ........ 32 Figure 3.11: Cascade of intervertebral disc degeneration and associated treatment options

........................................................................................................................................... 32 Figure 3.12: Schematic representation of hyalectans found in the intervertebral disc .... 38 Figure 3.13: The bottle brush structure of Aggrecan ....................................................... 39 Figure 3.14: Aggrecan Synthesis in vivo ......................................................................... 39 Figure 3.15: Electrostatic repulsion generated between opposing CS GAG chains.(69) 40 Figure 3.16: Opposing CS and aggrecan mechanics ....................................................... 40

Page 17: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xvi

Figure 3.17: MMP expression in cells of the NP ............................................................. 42 Figure 3.18: Sites of enzymatic cleavage of the aggrecan core protein (7). .................... 43 Figure 3.19: Aggrecan cleavage into smaller molecular weight fragments ..................... 43 Figure 3.20: Nutrient gradient across the NP ................................................................... 46 Figure 3.21: CS repeat disaccharide with possible sulfation points indicated in red. (95)

........................................................................................................................................... 54 Figure 3.22: X-ray diffraction structures of CS-4 ............................................................ 55 Figure 3.23: Ellipsometric measurement of end-grafted GAG height .............................. 56 Figure 3.24: Stability of CS in elevated pH and temperature conditions (92). ................ 58 Figure 3.25: Modified CS for the incorporation in cross-linked CS networks ................. 60 Figure 3.26: Chemically/Physically cross-linked CS networks (105) ............................. 60 Figure 3.27: The three main strategies for the fabrication of bottle brush polymers ....... 63 Figure 3.28: Sugar bearing polymerizable monomers (7) ................................................ 67 Figure 3.29: Cyanoxyl-mediated copolymerization of vinyl-glycomonomers ................ 67 Figure 4.1: Schematic of Biomimetic Aggrecan Design. ................................................ 73 Figure 5.1: Chemical Structure of L-serine and N-acetyl-D-galactosamine .................... 88 Figure 5.2: Chemical structure of the fluorescamine reagent .......................................... 88 Figure 5.3: Aldehyde-amine reaction and reductive amination with sodium cyanoborohydride (150) .................................................................................................... 89 Figure 5.4: Epoxide-amine reaction ................................................................................. 89 Figure 5.5: Carboxyl-amine reaction mediated by EDC and sulfo-NHS (151) ............... 90 Figure 5.6: Example of “grafting-to” strategy for building CS brushes from glass substrate ............................................................................................................................ 91 Figure 5.7: “Grafting-to” polymerization strategy using the carboxylic acid to CS amine reaction .............................................................................................................................. 92

Page 18: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xvii

Figure 5.8: L-serine-fluorescamine standard curve ....................................................... 107 Figure 5.9: Fluorescamine analysis of control samples bovine serum albumin and GalNAc ........................................................................................................................... 108 Figure 5.10: Fluorescamine analysis of chondroitin samples from various sources ..... 108 Figure 5.11: Representative EDAX spectra .................................................................. 109 Figure 5.12: Surface sulfur content on functionalized glass surfaces after treatment with CS .................................................................................................................................... 110 Figure 5.13: Contact angle measurements on CS functionalized glass slides ............... 111 Figure 5.14: Conjugation of CS with amine reactive monomers ................................... 112 Figure 5.15: CS-AGE conjugation over time for varying AA:CS molar ratios with one-phase association exponential curve fit (n=3, 2-way ANOVA with Bonferroni Post-tests)

......................................................................................................................................... 113 Figure 5.16: CS-Acrylic Acid conjugation over time for varying AA:CS molar ratios with one-phase association exponential curve fit (n=3, 2-way ANOVA with Bonferroni Post-test) ......................................................................................................................... 114 Figure 5.17: 1H-NMR spectra of (Top) CS and (bottom)AGE solutions in D2O with spectral assignments. See Appendix 2 for alternate view of CS. .................................... 115 Figure 5.18: CS region of 1H-NMR spectra of AGE-CS conjugate (1000:1 AGE:CS molar ratio) with increasing reaction time ...................................................................... 116 Figure 5.19: Analysis of 1H-NMR integrated area (normalized to GalNAc Methyl proton) of AGE-CS samples over 24h Reaction time ................................................... 117 Figure 5.20: 1H-NMR spectra of vinyl-proton region of AGE-CS (1000:1 molar ratio) samples synthesized over 24h with AGE vinyl proton spectral assignments ................. 118 Figure 5.21: Qualitative analysis of vinyl group incorporation into CS based on 1H-NMR spectral analysis of AGE-CS conjugation over time. Vinyl proton incorporation was based on the integrated area of proton 2 of AGE compared the integrated area of 1 proton, estimated from the methyl peak at 1.9ppm. ....................................................... 119 Figure 5.22: ATR-FTIR spectra and wavenumber assignments for chondroitin sulfate and AGE .......................................................................................................................... 120

Page 19: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xviii

Figure 5.23: ATR-FTIR spectra for AGE-CS conjugates reacted over 24h. Highlighted is the C-O stretching mode associated with the CS disaccharide backbone. The C-O stretching mode is sensitive to modification of the CS hydroxyl ................................... 121 Figure 5.24: CS-PAA conjugation over time ................................................................. 122 Figure 5.25: CS-PAA conjugation with varying temperature (8hr reaction) ................. 123 Figure 5.26: CS-PAA conjugation with varying Na+ Concentration (8hr Reaction) PAA

......................................................................................................................................... 124 Figure 5.27: CS-PAA conjugation with varying CS:PAA molar ratio (8hr Reaction) .. 125 Figure 5.28: Rheological investigation of CS-PAA reaction solutions ......................... 126 Figure 5.29: Specific viscosity over shear rate for a CS-PAA reacted solution ............ 127 Figure 5.30: Grafting-through polymerization of vinyl terminated chondroitin sulfate 148 Figure 5.31: Chain growth polymerization and propagation of the reactive center ..... 149 Figure 5.32: Flow chart of pathways to biomimetic aggrecan based on covalent binding chemistries investigated in this study. ............................................................................ 153 Figure 6.1: Bond lengths and angles in ethylene oxide (the epoxide ring) .................... 162 Figure 6.2: Nucleophile addition to an epoxide ring results in two possible products .. 162 Figure 6.3: Hydrogen bonding in the catalysis of the epoxy-amine reaction ................ 163 Figure 6.4: Possible reactions in an epoxy-amine system. ............................................ 164 Figure 6.5: Example reaction kinetics of an epoxide amine system .............................. 165 Figure 6.6: % Conversion vs molecular weight for typical step-growth polymerization reactions .......................................................................................................................... 166 Figure 6.7: Schemiatic of linear step-growth polymerization of amine terminated CS using the epoxy-amine reaction. ..................................................................................... 167 Figure 6.8: PEG-DGE (top) and Chondroitin-4-sulfate (bottom) 300MHz 1H-NMR spectra ............................................................................................................................. 176 Figure 6.9: FTIR Spectra and spectral assignments for Chondroitin Sulfate and PEG-DGE ................................................................................................................................ 177

Page 20: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xix

Figure 6.10: Schemiatic description of the time course and mechanism of size exclusion chromatography (199) ..................................................................................................... 178 Figure 6.11: Kinetics of CS Primary Amine Reaction with PEG-DGE at varying Temperatures and DGE Concentrations ......................................................................... 197 Figure 6.12: Kinetics of CS Primary Amine Reaction with EG-DGE at varying Temperatures and DGE Concentrations ......................................................................... 198 Figure 6.13: Kinetics of CS Primary Amine Reaction to DGE monomers of varying poly(ethylene glycol) chain length .................................................................................. 199 Figure 6.14: 1H-NMRmonitoring of Day-by-day dialysis of PEG-DGE-CS reaction for purification of Un-reacted PEG-DGE ............................................................................. 200 Figure 6.15: Epoxide chemical shift focus in 1H-NMR of Day-by-day dialysis of PEG-DGE-CS reaction purification ......................................................................................... 201 Figure 6.16: 1H-NMRmonitoring of Day-by-day dialysis of EG-DGE-CS reaction for Purification of Un-reacted EG-DGE ............................................................................... 202 Figure 6.17: Epoxide chemical shift focus in 1H-NMR of Day-by-day dialysis of EG-DGE-CS reaction purification ......................................................................................... 203 Figure 6.18: 1H-NMR analysis of PEG(1K)-CS purification via membrane dialysis ... 204 Figure 6.19: 1H-NMR monitoring of PEG-CS Synthesis over 96h reaction time ......... 205 Figure 6.20: 1H-NMR monitoring of EG-CS Synthesis over 96h reaction time ........... 206 Figure 6.21: Analysis of 1H-NMR peak integrations for PEG-CS (top) and EG-CS (bottom) Synthesis Over 96h reaction time .................................................................... 207 Figure 6.22: 1H-NMR of Purified EG-CS, PEG-CS and PEG(1K)-CS Synthesized at 20mM DGE Concentrations in SBB buffer (pH 9.4, 45̊C, 96h) ..................................... 208 Figure 6.23: Analysis of 1H-NMR peak integrations for CS macromolecules synthesized with varying PEG backbone MW .................................................................................. 209 Figure 6.24: FTIR Spectra for CS macromolecules synthesized with varying PEG backbone MW ................................................................................................................. 210 Figure 6.25: Chromatograms of CS, EG-CS and PEG-CS in relation to polystyrene sulfonate standards .......................................................................................................... 211 Figure 6.26: Polystyrene Sulfonate Standard molecular weight calibration curve ........ 212

Page 21: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xx

Figure 6.27: Example of molecular weight distribution curve and graphical representation of Mn, Mp, Mw, and Mz ......................................................................... 213 Figure 6.28: Concentration extrapolation of CS Intrinsic Viscosity .............................. 214 Figure 6.29: Fluorescamine analysis of CS primary amine stability over time in pH 9.4, 45̊C buffer conditions ..................................................................................................... 215 Figure 6.30: Near-IR Spectra of 1M PEG-DGE epoxide stability in SBB pH 9.4 buffer, 45̊C over 96h ................................................................................................................... 216 Figure 6.31: Near-IR Peak ratio analysis of PEG-DGE (1M concentration) in SBB (pH 9.4, 45̊C) over 96h. ......................................................................................................... 217 Figure 6.32: Fluorescamine investigation of Serine and Glycine versus CS conjugation to DGE ................................................................................................................................ 218 Figure 6.33: Primary and secondary amine reactions of CS with PEG-DGE ............... 234 Figure 6.34: % Yeild of CS-Macromolecules (combined results for EG-CS and PEG-CS)

......................................................................................................................................... 235Figure 6.35: near-IR analysis of PEG(1K)-DGE in DI water over 96h. ......................... 236 Figure 7.1: (Top) Schematic Diagram of Custom Built Membrane Osmometer and Data Acquisition System and (bottom) Image of Assembled Osmometer in Saline Bath ...... 249 Figure 7.2: Pressure Transducer Specifications ............................................................. 250 Figure 7.3: Custom written lab view program for data acquisition from membrane osmometer. Program interfaces with pressure transducer, collects voltage data and converts data to a pressure output. .................................................................................. 251 Figure 7.4: Input and output pane for LabvView program written for data acquisition and processing of membrane osmometer pressure transducer output. .................................. 252 Figure 7.5: TEM of CS (A) and close-up of condensed bead structure of CS (B) ........ 258 Figure 7.6: TEM of (A) Several aggrecan molecules clustered together. (B) Close-up of natural bovine ggrecan where condensed beads are arranged in a chain pattern. ........... 258 Figure 7.7: TEM of PEG-CS macromolecules after 24h of synthesis at lower (A) and higher (B) magnification. PEG-CS appears as small aggregates and individual condensed beads. .............................................................................................................................. 259

Page 22: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xxi

Figure 7.8: TEM of PEG-CS synthesized for 72h at lower magnification (A) demonstrating aggregates of PEG-CS and (B) higher magnification demonstrating beaded chain like structure of PEG-CS. ...................................................................................... 259 Figure 7.9: (Top) Specific Viscosity with shear rate of CS-macromolecules with varying PEG backbone lengths. (bottom) Specific viscosity of CS-macromolecules relative to CS

......................................................................................................................................... 260 Figure 7.10: Fixed Charge Density of CS-Macromolecules as determined by the Blyscan GAG assay. (1-way ANOVA statistical analysis, n=3) .................................................. 261 Figure 7.11: (Top) Validation of membrane osmometer using PEG solutions of known osmotic pressure (20K MW PEG solutions in 0.15M NaCl, room temperature) (244). 262 Figure 7.12: (Top) Representative Osmotic Pressure curves from Membrane Osmometer of CS-macromolecule solutions (50mg/ml, 0.15M NaCl, 25̊C, 1mL Volume). (Bottom) Average Steady State Osmotic Pressure of CS-macromolecule solutions (1-way ANOVA statistical analysis n=3). .................................................................................................. 263 Figure 7.13: Fibroblast cytotoxicity with exposure to varying concentrations of EG-DGE and PEG-DGE as determined by (top) live/dead imaging (Images taken at 10X magnification) and (bottom) MTT metabolic activity assay. (2-way ANOVA statistical Analysis) ......................................................................................................................... 264 Figure 7.14: Live/Dead imaging and % Viability analysis of Fibroblasts dosed with varying concentrations of EG-CS and PEG-CS macromolecules after 48h incubation. 265 Figure 7.15: 72h Cytotoxicity of Fibroblasts and NP Cells dosed with CS-macromolecules at varying concentrations as determined by the MTT assay ................ 266 Figure 7.16: Schematic diagram showing the fixed charges of the NP extracellular matrix and associated mobile counter ions (252). ...................................................................... 279 Figure 7.17: Extrapolated fixed charge density profiles for intervertebral disc tissue of varying degenerative grade (assuming 26yo represents Grade 1 (“healthy”) and 74yo represents Grade 5 discs) (84, 270-271) ......................................................................... 280 Figure 118: Schematics of CS-acrylate synthesis by sequential grafting of adipic dihydrazide and then acrylic acid to CS (293). ............................................................... 314 Figure 119: Synthesis schematic of Styrene sulfonated ethyl ester (SSE) and styrene sulfonated dodecyl ester (SSD) polymer brushes and resulting polymer brush structure as seen with atomic force microscopy imaging. (297) ........................................................ 315 Figure 120: Synthesis of poly-alkyne click-able polymer from protected a protected alkyne functional monomer (trimethylsilyl mechacrylate). (299) ................................. 316

Page 23: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xxii

Figure 121: Synthesis of click-able sugars (sugar containing azide groups) and their “Grafting-to” click conjugation to an alkyne polymeric backbone. (298-299) .............. 317 Figure 122: Synthesis schematic for fabrication of “molecular sugar sticks” using the “grafting from” approach via atom transfer radical polymerization (ATRP) of the protected monomer (bottom) scanning force microscopy height image of molecular sugar sticks. (116) ..................................................................................................................... 318 Figure 123: Synthesis of sulfated and non-sulfated vinyl glycomonomers and subsequent cyanoxyl mediated free-radical polymerization for the synthesis of biomimetic polymers with chemically stable hydrocarbon backbones and biologically active pendant hydrophilic saccharides. (123) ........................................................................................ 319 Figure 124: Polycondensation step-growth polymerization of glycopeptides for the synthesis of sulfated glycopolypeptides as antifreeze glycoprotein analogs (303). ....... 320

Page 24: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xxiii

Abstract Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid Bio/Synthetic

Biomimetic Aggrecan Macromolecule Sumona Sarkar

Advisor: Michele Marcolongo, Ph.D.

Lower back pain resulting from intervertebral disc degeneration is one of the

leading musculoskeletal disorders confronting our health system. In order to

mechanically stabilize the disc early in the degenerative cascade and prevent the need for

spinal fusion surgeries, we have proposed the development of a hybrid-bio/synthetic

biomimetic proteoglycan macromolecule for injection into the disc in the early stages of

degeneration. The goal of this thesis was to incorporate natural chondroitin sulfate (CS)

chains into bottle brush polymer synthesis strategies for the fabrication of CS-

macromolecules which mimic the proteoglycan structure and function while resisting

enzymatic degradation. Both the “grafting-to” and “grafting-through” techniques of

bottle brush synthesis were explored. CS was immobilized via a terminal primary amine

onto a model polymeric backbone (polyacrylic acid) for investigation of the “grafting-to”

strategy and an epoxy-amine step-growth polymerization technique was utilized for the

“grafting-through” synthesis of CS-macromolecules with polyethylene glycol backbone

segments.

Incorporation of a synthetic polymeric backbone at the terminal amine of CS was

confirmed via biochemical assays, 1H-NMR and FTIR spectroscopy, and CS-

macromolecule size was demonstrated to be higher than that of natural CS via gel

permeation chromatography, transmission electron microscopy and viscosity

Page 25: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

xxiv

measurements. Further analysis of CS-macromolecule functionality indicated

maintenance of natural CS properties such as high fixed charge density, high osmotic

potential and low cytotoxicity with nucleus pulposus cells.

These studies are the first attempt at the incorporation of natural CS into

biomimetic bottle brush structures. CS-macromolecules synthesized via the methods

developed in these studies may be utilized in the treatment and prevention of debilitating

back pain as well as act as mimetics for other proteoglycans implicated in cartilage, heart

valve, and nervous system tissue function.

Page 26: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

ii

Page 27: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

1

Chapter 1 : Introduction

Lower back pain resulting from intervertebral disc degeneration is one of the

leading musculoskeletal disorders confronting our health system with one third of the

population experiencing lower back pain annually (1). This irreversible process leads to

loss of mechanical stability with the potential for disc herniation and nerve damage (2).

While disc degeneration is poorly understood, it has been shown that the proteoglycan

content of the nucleus pulposus, or inner region of the disc, decreases linearly with age

and degeneration with almost 40% of proteoglycans lost by the age of 40 (3). This loss of

proteoglycans leads to a loss of hydration of the nucleus manifested by a reduction of

disc height and intradiscal pressure (4). The most common surgical approaches treat the

end-stage of this degeneration with highly costly and highly invasive spinal fusion

surgery while early-stage interventions are aimed at treating the symptoms and not the

causes of back pain (5-6).

In order to mechanically stabilize the disc early in the degenerative cascade and

prevent the need for spinal fusion surgeries, we have proposed the development of a

hybrid-bio/synthetic biomimetic proteoglycan macromolecule for injection into the disc

in the early stages of degeneration, after a first or second episode of back pain. The

biomimetic proteoglycan is based on the major disc proteoglycan aggrecan which is a

bottle brush macromolecule with a protein core and radiating negatively charged

chondroitin sulfate (CS) bristles (7). In the biomimetic approach, the protein core is

replaced with a synthetic polymeric backbone in order to resist enzymatic and hydrolytic

Page 28: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

2

degradation within the nucleus environment while maintaining the hydrating and

biologically active chondroitin sulfate bristles in their natural bottle brush arrangement.

The goal of this thesis was to incorporate for the first time, chondroitin sulfate

chains into bottle brush polymer synthesis strategies for the fabrication of CS-

macromolecules which mimic the aggrecan structure and function. The feasibility of

utilizing both the “grafting-to” and “grafting-through” techniques of bottle brush

synthesis was explored. CS was immobilized onto a model polymeric backbone of

poly(acrylic acid) for investigation of the “grafting-to” strategy via carboxyl-amine

interactions. For investigation of the “grafting-through” strategy of synthesis, an epoxy-

amine step growth polymerization technique was utilized to synthesize macromolecules

with CS bristles and poly(ethylene glycol) backbone segments. Incorporation of a

synthetic polymeric backbone at the terminal amine of CS was investigated via

biochemical assays, proton nuclear magnetic resonance analysis and infrared

spectroscopy, and CS-macromolecule size was demonstrated via gel permeation

chromatography, transmission electron microscopy and viscosity measurements. Further

analysis of CS-macromolecule functionality was conducted in order to characterize

macromolecule fixed charge density, osmotic potential (via membrane osmometry) and

cytotoxicity. These studies are the first attempt at the incorporation of natural CS into

biomimetic bottle brush structures. CS-macromolecules synthesized via the methods

developed in these studies may be utilized in the treatment and prevention of debilitating

back pain. With the high social and economic costs of spine disorders, the development

of an early interventional, minimally invasive treatment of back pain could have far

reaching benefits for our society.

Page 29: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

3

Additionally, macromolecules synthesized in this biomimetic strategy may serve

to mimic other macromolecules of the proteoglycan family. The proteoglycan family

consists of a large array of macromolecules with similar structure, i.e. protein core with

attached glycosaminoglycan (GAG) chains. These proteoglycans differ in core protein

length and amino acid sequence as well as GAG grafting density. The strategies for

biomimetic aggrecan synthesis developed here are transferable to the synthesis of a

family of biomimetic proteoglycan macromolecules that may fill a wide array of

functions in regenerative and restorative medicine (8).

Page 30: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

4

Chapter 2 : Significance

Low Back pain is one of the leading musculoskeletal disorders confronting our

health system with between 30% and 40% of adults in the U.S. reporting to have

experienced low back pain in the previous three months (1). Overall, one in two persons

reports experiencing back pain at least once a year (Figure 2.1). Of the various causes of

low back pain including lumbar back disorders, back injury, and disc disorders,

degenerative changes of the disc are a leading contributor to spinal fusion surgeries with

an increasing trend over the past several decades (Figure 2.2) (6). Lumbar fusion

procedures on average were reported to cost $63,520 in 2004 with an average hospital

stay of 4.7 days. A rising cost of spinal fusion procedures along with a steady increase in

the number of procedures performed has attributed to an estimated cost of $16.9 billion in

2004 for primary spinal fusions, a 215% increase from that reported in 1998 (Figure:

2.3)(1).

The socioeconomic burden of low back pain and degenerative disc disease is

especially striking with 55% to 84% of those individuals aged 44 to 65 reporting a

limited work capacity, attributing their limitation to back pain(1). In 2004 alone, an

estimated 186.7 million work days were lost due to back pain(1). In 2004, the estimated

annual cost for medical care of spine conditions was $193.9 billion, a rise from the

$130.2 billion reported in 1996, an increase of 49%. Earnings loss, due to spine

conditions was estimated at 22.4 billion per year between 2000 and 2004(1). With the

high social and economic costs of spine disorders, the development of an early

Page 31: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

5

interventional, minimally invasive treatment of back pain could have far reaching

benefits for our society.

Earlier stage interventions are currently being explored, such as nucleus pulposus

replacement (9-15). In this procedure the nucleus pulposus is excised and replaced with a

medical device with the goal of restoring disc biomechanics, providing stabilization to

the disc and thereby alleviating lower back pain. This technique however results in

complete removal of the native nucleus tissue, removing the possibility of regeneration

and fully relying on the implanted material for mechanical and biological function.

By mechanically stabilizing the disc early in the degenerative cascade, one may

be able to protect the outer disc tissue, the annulus fibrosus, from an adverse stress state

which may limit or prevent tearing and subsequent herniation of the disc. The approach

proposed here is to develop a hybrid synthetic/bio-based aggrecan-like macromolecule

for injection into discs in the early stages of degeneration, after a first or second episode

of lower back pain. These injections will be minimally invasive, and the implanted

macromolecule will be able to integrate with surrounding tissues working in concert to

maintain disc function.

Page 32: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

6

Figure 2.1: Prevalence of self-reported joint pain by site for persons aged 18 and over in two national health surveys in the United States between

1999-2005. (1)

Figure 2.2: Lumbar fusion rates by primary diagnosis.(6)

Page 33: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

7

Figure: 2.3: Seven year trend in total cost for spinal fusion procedures in the United (1)

Page 34: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

8

Chapter 3 : Background

3.1.

3.1.1. The Vertebral Column

Anatomy and Physiology of the Spine

The human vertebral column, or spine functions to transfer loads and bending

moments of the head, trunk, and any external loads to the pelvis, allow sufficient

physiological movement and flexibility of the upper body and protect the spinal cord

from danger due to motion and trauma (16). It also serves as protection to other vital

internal organs and as a base of attachment for upper-body ligaments, tendons, and

muscles. These functions help explain its form and structure.

The spine consists of a series of individual vertebre, the sacrum and the coccyx

and spans from the skull to the pelvis. Adjacent vertebrae are connected by an

intervertebral disc (IVD). The vertebrae are stacked on top of each other and are grouped

into five distinct regions—there are cervical, thoracic and lumbar vertebrae, as well as a

fused sacral, and fused coccyx (Figure 3.1). From posterior perspective, the spine appears

to be straight, but lateral view shows four normal curves that mechanically serve to give

the spinal column increased flexibility and shock-absorption ability.

3.1.2. The Intervertebral Disc

The intervertebral disc separates adjacent vertebra and acts as a shock absorber,

dissipating energy under loading. The disc is the largest avascular tissue in the human

body requiring nutrients to enter and waste to be removed primarily through diffusion

and fluid flow (16). The disc structure is quite complex combining three distinct regions:

the nucleus pulposus, an inner gelatinous core; the annulus fibrosus, a surrounding

Page 35: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

9

ordered fibrous structure and the bony endplates that are partly cartilaginous and partly

bony and serve as an interface between the disc and the vertebral bodies (Figure 3.2).

The cellular content of the disc is very low, less than ca. 0.25% of tissue volume (17).

The extracellular disc matrix plays the main role in determining the mechanical

properties of the disc.

3.1.3. The Nucleus Pulposus

The central nucleus pulposus (NP) region of the IVD is a gelatinous structure that

is mostly water (77% of wet weight) but also contains a large constituent of

proteoglycans (PGs) (14% of wet weight) which are composed of charged

glycosaminoglycan (GAG) chains covalently attached to a protein core. Randomly

organized collagen fibers are also present in the NP (about 4% of wet weight) and

associate with the proteoglycans to form a loose network (18). Collagen fibers within the

NP provide tensile strength while proteoglycans create a large osmotic swelling pressure

and draw water into the tissue (Figure 3.3). Charged groups on the GAG chains of the

PGs carry with them mobile counter ions like Na+ which draw water is drawn into the

tissue because of the osmotic imbalance generating a hydrostatic pressure within the NP

(7).

3.1.4. The Annulus Fibrosus

The annulus fibrosus (AF) of the intervertebral disc is a structure that serves to

contain the gelatinous nucleus pulposus. The chemical composition of the AF comprises

65-90% wet weight of water, 50-70% dry weight of collagen, 10-20% dry weight of

proteoglycans, and non-collagenous proteins such as elastin (19). The structure of the AF

is laminate in nature, consisting of a minimum of 15 (posterior) and 25 (lateral)

Page 36: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

10

concentric layers. The layers are made up of type I collagen fibers which alternate in

angles from 30o at the peripheral AF to 44o at the central AF with reference to the

transverse plane of the disc (Figure 3.4) (19). Spinal compression generates a hydrostatic

pressure in the NP which places the annulus into tension thereby resisting the applied

load (Figure 2.4.1.) (20). When a compressive load (C) is placed on the IVD, a

hydrostatic pressure (P) is generated in the NP and tensile stress (T) on the annulus. The

AF has a very highly organized structure which results in complex anisotropic behavior

with loading.

Page 37: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

11

Figure 3.1: Regions of the human spine with respect to the vertebrae and corresponding intervertebral discs

(http://healthbase.wordpress.com/2007/06/13/spinal-fusion-surgery/)

Page 38: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

12

Figure 3.2: The Intervertebral disc location between adjacent vertebrae and the individual structures and average dimensions of the IVD (18)

Page 39: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

13

Figure 3.3: Constituents of the nucleus pulposus and subsequent development of hydrostatic pressure (7)

Page 40: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

14

Figure 3.4: Load transfer in the IVD and lamellar structure if the annulus fibrosus. (20)

Page 41: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

15

3.2. Intervertebral Disc Degeneration

The intervertebral disc of the spine is subject to degenerative changes induced by

normal aging as well as injury, loading or genetically induced accelerated disc

degeneration (21). The exact mechanism of IVD degeneration is not well understood;

however possible degenerative cascades have been proposed which lead to eventual

changes in disc mechanical properties (Figure 3.5). Initial changes in the IVD are thought

to be brought about by changes in the nutrient flow to and waste product flow out of the

nucleus. Calcification of the cartilaginous endplates, the major route of nutrient supply to

the disc, leads to harsh cellular environments subsequently reducing matrix synthesis and

increasing enzyme production further reducing proteoglycan, collagen and elastin content

of the disc, in particular in the nucleus pulposus (3, 21-26) (Figure 3.6). The loss of PGs

in the NP has major effects on the load-bearing behavior of the disc, causing the osmotic

pressure of the disc to fall and reducing the disc water content disrupting the NP’s ability

to behave hydrostatically under load (Figure 3.7).

Figure 3.5

Loading can then lead to inappropriate

stress concentrations along the endplate and in the annulus with progressive changes as

disc degeneration proceeds ( ). Changes in stress concentration can lead to

damage of the annulus fibrosus as well as the endplates with stress concentrations being

associated with discogenic pain (18). Several manifestations of disc degeneration and

damage can be visualized with MRI including herniation of the lumbar disc, degeneration

of the lumbar disc, signal changes in the vertebral end plates and annular fissures.

Although these alterations in the IVD are prevalent in the asymptomatic population

Page 42: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

16

(found in 25 to 70% of asymptomatic subjects), disc disruptions have been associated

with back pain and are targets for surgical interventions such as spinal fusion, total disc

replacement, nucleus replacement and annulus repair (27).

Page 43: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

17

Figure 3.5: Disc degeneration and mechanical consequences (A) Cascade of events in IVD degeneration (28) and (B) changes in mechanical properties of the

IVD with degeneration (normal disc to degenerated disc from top to bottom) (29)

Page 44: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

18

Figure 3.6: Changes in major biomacromolecule concentrations in the annulus fibrosus and nucleus pulposus of the intervertebral disc with aging (3, 24-26)

Page 45: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

19

Figure 3.7: Change in water content in the IVD and associated change in glycosaminoglycan content in the IVD with aging and degeneration. (30)

Page 46: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

20

Degeneration of the intervertebral disc is often associated with back pain, making

the IVD a target for surgical intervention. Over 300,000 lumbar discectomy surgeries are

performed annually in the United States accounting for over 34% of all spinal procedures

conducted (1). Removal of disc material can have adverse effects, however, on adjacent

level spinal segments and can lead to instability of the spine (31).

3.3. Treatment Options

The first mode of treatment proposed for the treatment of the majority of patients

who suffer from degenerative disc degeneration is non-operative care. Conservative

treatments include anti-inflammatory medication, physical therapy, and injections. These

treatments are often effective with 70% of patients pain-free after treatment and as low as

25% recurrence. Patients with chronic back pain however are more difficult to treat

successfully.

Of particular interests, as a minimally invasive non-surgical procedure to reduce

back pain is injection therapy. Injection therapies for chronic low back pain include

injections of anti-inflamitory agents, medications, irritants, or proteolytic enzymes into

soft tissues outside or within the spine (32). Epidural steroid injections for managing

chronic low back pain are one of the most commonly performed interventions in the

United States and have indications ranging from sciatica and radiculopathy to herniated

discs, and spinal stenosis (5, 33). While these procedures have shown some short term

(less than 6 weeks) and longer term (6 weeks or longer) effects in providing pain relief,

evidence of the effectiveness of these procedures is conflicting (5, 33). Additionally, the

underlying mechanism of action of epidural steroids is not well understood. One

Page 47: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

21

mechanism is that the resulting neural block interrupts nociceptor input and self-

sustaining activity of neurons as well as patterns of neuronal activity (33). Steroids also

have anti-inflammatory activity by inhibiting release or synthesis of pro-inflammatory

mediator as well as providing local reversible anesthetic effects (33). All of these

mechanisms of action however do not address the root mechanical or biological causes of

pain, and simply address the symptoms. Injections directly into the nucleus region of the

intervertebral disc are also performed to deliver chymopapain to digest nucleus tissue,

thereby reducing the pressure of herniated tissue on the spinal nerves. This procedures

however has not been shown to be more effective than total discectomy (32).

Based on carefully considered guidelines, surgical interventions are recommended

in the event that conservative remedies do not address the clinical needs of the patient.

The following is a review of current surgical and implantable interventions for the

treatment of back pain.

Arthrodesis or spinal fusion is the traditionally accepted surgical intervention

technique advocated in cases of isolated disc degeneration in the cervical or lumbar spine.

The procedure involves the removal of a motion segment through the use of bone grafts

and sometimes through internal fixation. Indications for the procedure are instability of a

motion segment caused by traumatic injury or degeneration (34). A successful fusion

procedure eliminates motion that causes pain and offers the ability to restore

intervertebral height and alignment (35).

Early spinal implants were based on metals due to the utility of metals in other

orthopedic devices. The complex loading nature and function of the spine however

Page 48: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

22

necessitated the development of novel materials and techniques in surgical spine

interventions. Failures of earlier devices have especially expanded the breadth of

materials utilized in spinal interventions. Adjacent level disc degeneration with spinal

fusion, expulsion of implant materials through surgical sites and subsidence of implants

into the endplates have driven the development of new surgical techniques and provided

the impetus for the development of novel spine materials. This is especially illustrated by

the development of in situ curing hydrogel and non-hydrogel nucleus replacements where

materials were designed such that they could be delivered with minimal damage to the

annulus fibrosus. Another drive in materials for spine implants is the utilization of tissue

engineering and molecular approaches.

3.3.1. Spinal fusion

Spinal fusion and stabilization utilizes a wide range of devices such as screws,

wires, artificial ligaments and vertebral cages (36). The instrumentation used in fusion

surgery is designed to stabilize the spine during fusion and promote osseous fusion (37).

Early spinal fusion devices were based on metals commonly utilized in other orthopedic

surgeries such as in hip implants. Metals are still widely used in spine fusion implants

today; however advances have been made implant design and material development in

order to address the unique needs of the spine including maintenance of spine flexibility

and to avoid the occurrence of adjacent area damage.

Spinal Cages and Interbody Spacers. The removal of the intervertebral disc in

order to address discogenic back pain and disc herniation will lead to the loss of disc

Page 49: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

23

height and lead to spinal instability. Cage structures have been developed to assist in

implanting allo- and autologous bone graft interbody spacers. These cages enhance

stabilization of the spinal segment and prevent postoperative collapse. The BAKTM

interbody fusion system for example is a cage structure made of porous Ti-6Al-4V shell

with a cavity for the placement of bone grafts (38).

Spinal fusion has its limitations. No significant improvement has been seen in the

success rate of fusion procedures despite the advances made in terms of surgical

techniques and instruments. Spinal fusion is also associated with accelerated adjacent

level disc degeneration, long post-operative recuperation times and high cost (39).

3.3.2. Total Disc Arthroplasty

The total disc arthroplasty procedure aims to potentially eliminate a painful disc

while preserving/restoring motion (35). The goals of arthroplasty include successful pain

relief and functional recovery, acceptable levels of morbidity associated with the

procedure (equal to or less than those seen in fusion), shorter post-operative recuperation

times and easily implantable surgical techniques (39). The biomechanical objectives of

disc arthroplasty are the restoration of disc function to relieve abnormal stress or strain

caused by degeneration. More importantly, it is essential that the procedure not cause any

detrimental biomechanical effects on surrounding structures within the motion segment

and on adjacent levels(39).

As an example, The Charite` Artifical Disc (marketed by DePuy Spine) was

designed with the goal of replicating the kinematics of the disc. The basic design of the

device incorporates two metallic end plates made of cobalt-chromium alloy (34) which

Page 50: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

24

are attached to adjacent vertebral bodies, and articulate against a central polymeric core

made up of ultra high molecular weight polyethylene (UHMWPE) (40). The Charite`

Aritficial disc has undergone four iterations in design since its inception and offers the

largest and longest clinical trial of any existing artificial disc (34, 41). There have been a

number of clinical studies undertaken to document the performance of the SB Charite`

III (Figure 3.8). Most of these studies have demonstrated that the Charite` artificial disc

has the potential to preserve motion and survive long term implantation in the body with

good to excellent outcomes. However, cohort studies undertaken have shown variability

in the complication and success rates of the device, which raises concerns regarding the

repeatability and reproducibility of the procedure.

3.3.3. Nucleus Replacement

Nucleus replacement offers a less invasive alternative to disectomy with fusion or

total disc replacement. Nucleus replacement devices aim to replace herniated or

surgically removed NP tissue with a prosthesis that can function as a load sharing device

(12-13, 42-43). The goal is to help prevent further collapse and degeneration of the

remaining disc as well as restore the biomechanical function of the annulus by placing the

annular fibers in tension (44).

The original nucleus replacement technologies aimed to primarily mimic the

geometry of the nucleus pulposus. These approaches held the premise that by replacing

the lost nucleus material and restoring disc height, pressure on impinged nerves would be

relieved, alleviating back pain (12). Stainless steel balls and other metal devices were

Page 51: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

25

used for this purpose however, these metal devices, due primarily to their high stiffness,

could not restore normal load distribution and often resulted in implant migration and

subsidence (45). As a result, designs incorporating elastomers and viscoelastic behaviors

have been investigated for nucleus replacement (12, 34, 41, 44-46).

Several requirements have been established for the NP device and its materials.

The implant should be biocompatible without causing local cell death or systemic

toxicity. The device must also be durable and exhibit fatigue resistance as it is likely to

experience loads of over 100 million cycles over 40 years (44). The material should

exhibit low wear characteristics and avoid the formation of particulates. It is also

important that implant have a similar stiffness as the native nucleus tissue to avoid stress

shielding and bone resorption that may lead to endplate subsidence and extrusion of the

implant (44). The device should also fill the disc space to prevent implant movement and

restore disc height while avoiding overfilling the cavity which can result in significant

increases in the stiffness of the joint (13).

Nucleus replacement devices can be subdivided into several categories based on

their primary mode of function (mechanical or elastomeric), material and mode of

delivery. Mechanical nucleus devices can be either one piece or two pieces. The Regain

is a solid 1-piece mechanical nucleus composed of a carbon substrate with a pyrolytic

carbon coating. Its geometry is convex such that it can conform to the natural convexity

of the vertebral endplates. This device is currently in pilot feasibility studies (47). The

Nubac Disc Arthroplasty Device is a two-piece mechanical nucleus and is a PEEK-on-

PEEK articulating device. This device has a ball and socket design with a large surface

contact area designed to distribute stresses and decrease the risk of subsidence. Wear of

Page 52: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

26

the PEEK material is a concern, however over 100 devices have been implanted

worldwide and pilot feasibility studies are underway (45, 47-48).

Elastomeric devices can further be subdivided into hydrogel vs non-hydrorgel

materials, both of which may be delivered either preformed, or in an injectable, in-situ

forming material. Hydrogels are polymers that absorb water and swell when placed in

fluid environments while non-hydrogels polymers are low-friction materials with shock

absorbing capabilities. Some elastomeric devices and their materials are listed in Table

6.1.

Pre-formed Nucleus Replacements. The prosthetic disc nucleus PDN

(HydraflexTM and SOLOTM), was the most widely studied nucleoplasty device with over

5,500 devices implanted world wide (49). The PDN is composed of a preformed

hydrogel core consisting of polyacrylonitrile and polyacrylamide with a polyethylene-

woven jacket (Figure 3.9). This device was implanted dehydrated, and upon implantation

rehydrated, absorbing up to 80% of its own weight in water. This device is no longer

under investigation due to issues with disruption of the woven jacket and subsequent

fragmentation of the hydrogel core (50).

In Situ forming Non-Hydrogel Nucleus Replacements. In efforts to better match

the mechanical needs of a nucleus replacement and to reduce chance for expulsion novel

injectable in situ curing materials are being investigated further. It is especially important

in a nucleus replacement, for the material to resist expulsion from the nucleus cavity

under repetitive loading. In situ curing polymers are therefore of great interest as they

can be administered via minimally invasive techniques, and thereby maintain annulus

Page 53: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

27

integrity and reduce risk of expulsion via the surgical site. In situ forming biomaterials

are designed to change shape or form either during or after implantation (51). In situ

forming biomaterials include polymers or prepolymers that are crosslinked after

implantation. NuCoreTM injectable disc nucleus (Figure 3.10), DascorTM (Disc Dynamics,

Inc. Eden Prairie, MN) prosthetic intervertebral nucleus and Percutaneous Nucleus

ReplacementTM devices are injectable non-hydrogel devices (44). HydrafilTM nucleus

replacement implant and BiodiscTM are injectable hydrogel implants that form in situ (47,

52-53).

These NP replacements are currently at different stages of preclinical and clinical

investigates (Table 3.1). The Biodisc, PDN, Aquarelle, NeuDisc, PNR, Dascor and

Newleus are no longer under clinical investigation due to adverse preclinical trials. The

most common complication of NP replacement is implant extrusion through the insertion

defect that must inevitably be created in the annulus. Endplate failure and subsidence are

also common modes of failure.

3.3.4. Tissue Engineering and Regenerative Medicine Approaches to Disc Degeneration

In addition to the replacement of NP with a prosthetic device, tissue engineering

approaches are also being investigated. The tissue engineering approach provides the

possibility to recover the functionality of the degenerated intervertebral disc. By

manipulating the discs biology to inhibit further degeneration by the introduction of

therapeutic agents such as cells, growth factors and scaffolds, tissue engineering

approaches aim to restore native NP functionality (54). Many scaffold materials have

Page 54: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

28

been investigated to support cell and drug delivery to the NP including

chitosan/hydroxybutyl chitosan, alginate, collagen/atelocollagen, gelatin, hyaluronan,

calcium phosphate, polyl-D, L-lactide (PDLA), demineralized bone matrix (DBM) and

small intestine submucosa (SIS) (54-55). Although NP tissue engineering provides a

promising interventional techinique and alternative to implanted NP devices, its clinical

application is still far away. Although NP cells have shown promise in in vitro cultures

poor nutrient supply in the NP area of degenerated discs will limit cellularity and

functionality of the implanted NP graft material and therefore nutrient availability must

be addressed for successful implantation (55).

Regenerative biological solutions aim to stop disease progression at a cellular

level, addressing the underlying pathology of disc degeneration (56). Disc disease

includes the loss of proteoglycans, increased catabolic enzyme activity, increased cell

senescence and cell death. A major effort in therapeutic strategies for disc degeneration

is the restoration of the disc’s structural morphology by restoring the disc matrix and

reversing the appearance of disc degeneration (57). In order to achieve this, therapeutic

strategies either promote native cells to produce more matrix, or promote cells to reduce

catabolic activity. Earlier strategies investigated promitotic growth factors, however cell

mitosis and an increase in cell number may not be beneficial as there is evidence that

there is an upper limit on cell density in the disc matrix (57).

Gene therapy for example may improve the function and integrity of the

degenerated disc by altering cellular gene expression profiles. Transduction of TGF-β

using adenovirus vectors (in rabbit models) have resulted in increased expression of

TGF-β and proteoglycans without eliciting an immune response (56). In addition, the

Page 55: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

29

introduction of some molecules (bone morphogenic protein (BMP-2, BMP-7) and growth

differentiation factor (GDF-5)) in culture has proven to be highly effective in increasing

disc matrix production (57). BMP-7 and GDF-5 in particular are currently in FDA trials

to test whether single injections of therapeutic molecules can improve disc matrix

appearance on MRI and improve back pain (57). In contrast, up regulation of anti-

catabolic cytokines (IL-1 receptor antibody) may neutralize the increased catabolic action

in the disc (i.e. activity of IL-1 catabolic cytokine), thus correcting the imbalance

between anabolism and catabolism in the disc (56). These strategies however are

similarly limited by their dependence on cellular activity in the nutrient deprived disc

environment.

3.2.4. Summary of current treatment options: a need for an early interventional strategy

In the cascade of disc degeneration, the majority of clinically available treatments

are for the end stages of disc degeneration (Figure 3.11). At this final stage, the IVD can

no longer maintain its height or mechanical and stabilizing properties so it is removed

from the spinal column and replaced by a stabilizing unit. The adjacent vertebrae can be

fused together by various means to stabilize the motion segment, or an articulating total

disc replacement can be used in an attempt to maintain normal spine kinematics. At

stages of moderate degeneration nucleus replacement technologies are being investigated

in both laboratory and clinical settings. In these approaches, the annulus tissue is kept

intact (with the exception of a puncture wound) while the degenerated nucleus tissue is

removed and replaced with a synthetic mechanically functional material. As more

Page 56: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

30

knowledge of the degenerative cascade is gained however the possibility of earlier stage

treatments (after the first or second episode of back pain) which preserve more of the

natural disc biology can be envisioned. Tissue engineering approaches and regenerative

therapies are currently being investigated for these early stage replacements however

limitations with the inherent hostile nature of the nucleus pulposus environment pose a

great challenge to these strategies. An early interventional minimally invasive strategy

which restores mechanical stability to the spinal column and integrates with existing

tissue could be utilized to treat early episodes of back pain while promoting healing

preventing further degeneration (Figure 3.11). Such an interventional strategy must be

robust enough to maintain functionality in the NP environment while integrating with

existing tissue and supporting regeneration.

Page 57: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

31

Figure 3.8: SB Charite` III, and radiographic image of the implanted device (41).

Figure 3.9: PDN-SOLOTM device in hydrated and dehydrated states. Porous polyethylene weave allows fluid to pass into the hydrophilic core allowing the device to

expand vertically and horizontally (arrow).(44)

Page 58: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

32

Figure 3.10: Photograph of the NuCoreTM (Spine Wave) Injectable Disc Nucleus (red) after injection into a nucleated cadaveric spinal motion segment (lateral cross-section) (58).

Figure 3.11: Cascade of intervertebral disc degeneration and associated treatment options

Page 59: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

33

Table 3.1: Elastomeric Nucleus Pulposus Replacement Devices

Delivery Device Material Status

Hydrogel Injectable Biodisc; Cryolife Albumin and glutaraldehyde

No Longer Under Investigation

Hydrafil; Synthes polyvinyl alcohol/polyvinyl pyrollidine copolymer

In Preclinical Studies

Pre-formed

PDN Hydraflex and SOLO; Raymedica

hydrolyzed polyacrylonitrile No Longer Under Investigation

Aquarelle; Stryker polyvinyl alcohol No Longer Under Investigation

NeuDisc; Replication Medical

partially hydrolyzed acrylic copolymer

No Longer Under Investigation

Non-Hydrogel

Injectable NuCore; Spine Wave

rDNA-based synthetic protein copolymer of silk/elastin

In Pilot Clinical Studies

PNR; TranS1 silicone No Longer Under Investigation

Dascor; Disc Dynamics

polyurethane No Longer Under Investigation

Pre-formed

Newleus; Zimmer polyurethane No Longer Under Investigation

Page 60: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

34

3.4.1. Proteoglycan Diversity, Structure and Function

3.4. Aggrecan structure, assembly, and function

Proteoglycans (PGs) are defined as proteins that have one or more

glycosaminoglycan (GAG) chains attached (59). Matrix proteoglycans can be separated

into three groups which vary in structure and function: basement membrane

proteoglycans, proteoglycans interacting with hyaluronan (HA) and lectins (hyalectans),

and small leucine-rich proteoglycans (60). Structurally, proteoglycans differ in the

sequence and size of their protein core as well as the number and types of GAGs

attached. The major classes of associating GAGs are chondroitin sulfate, heparin sulfate,

keratan sulfate, and dermatan sulfate (60). GAG chains are attached to PGs via

glycosidic linkages joining the terminal end of the GAG to an amino acid residue of the

protein backbone. Proteoglycans have diverse functionality. They act to structurally

organize tissues, influence cell growth, act as growth factor and ion reservoirs (playing a

role as biological filters and modulating growth factor activity, and imparting osmotic

pressure), regulate collagen fibrillogenesis and skin tensile strength and influence corneal

transparency as well as neurite growth (8, 60). In the intervertebral disc, the main

proteoglycans are aggrecan, versican, decorin, and biglycan, some of which are

respresented as schematics in Figure 3.12. The most abundant disc proteoglycan is

aggrecan and will be the focus of this study due to its critical role in maintaining disc

hydration and mechanical function (61).

Page 61: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

35

3.4.2. Aggrecan Structure.

Aggrecan, an aggregating proteoglycan, is the major proteoglycan of the

intervertebral disc. Aggrecan consists of a protein core (approximately 300kDa and

400nm in contour length) to which glycosaminoglycan chains are covalently attached

resulting in a macromolecule with total molecular mass approximately 2,200kDa (62-

63). The protein core consists of several domains which allow for the molecules

flexibility (IGD) attachment to hyaluronic acid (G1 globular domain) attachment of

chondroitin sulfate(CS) (CS1 and CS2 domains) and keratan sulfate (KS) (KS domain)

and cell signaling (G3 globular domain) (Figure 3.13). N- and O-linked oligosaccharides

also decorate the aggrecan core protein throughout the CS and KS regions (64).

Approximately 100 CS glycosaminoglycan (GAG) chains are covalently attached to the

core protein in the CS region with a grafting density of approximately 0.25 to 0.5 nm-1.

Each CS chain which consists of 10-50 repeating disaccharide units of glucuronic acid

(GlcUA) and n-acetylgalactosamine (GalNAc) and is approximately 20kDa with a length

of 40nm(65). The KS region of the aggrecan core protein is smaller with only ~30 KS

chains attached. KS is a smaller GAG chain of 5-15kDa. Approximately 8000-10000

negatively charged groups are present in the aggrecan bottle brush via the charged sulfate

and carboxylic acids of the attached CS and KS chains (64, 66).

3.4.3. Aggrecan In Vivo Synthesis.

Aggrecan is assembled intracellularly through a complex series of enzymatic

reactions. Synthesis begins with translation of the core protein followed by the addition

Page 62: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

36

of a xylose sugar unit to the core protein, where sequence motifs of Ser-Gly-X-Gly or

(Asp/Glu)-X-Ser-Gly serve as sites for xylosylation on the serine residue and a

glycosidic bond is formed between xylose and the hydroxyl group of serine (Figure 3.14)

(67). Additional CS linkage sugars are then added to the xylose sugar. Subsequent CS

disaccharide polymerization and sulfation occurs on the growing sugar chains building

the CS bristles out from the core protein resulting in a densely packed bottle brush

structure (67). After intracellular assembly, the aggrecan macromolecule is secreted to

the extracellular space where it is able to assemble into larger aggregate structures with

hyaluronic acid and link protein (67). Although aggrecan is able to associate with HA

and link protein extracellularly, large HA-aggrecan aggregates are only predominant in

infancy such that by 6 months of age only approximately 30% of the NP is in the

aggregate form, levels as low as 10% aggregation are seen in the adult NP(68).

3.4.4. Functional molecular biomechanics of aggrecan

Aggrecan has two main mechanical functions in the disc: 1) it allows water

uptake by the nucleus due to sulfated groups in the chondroitin and keratin sulfate rich

regions which, in part, provides intradiscal pressure (20) and 2) it provides electrostatic

repulsion due to the elegant 3D macromolecular bottle brush structure, which contributes

to intradiscal pressure and disc height (69-71). A main constituent of aggrecan is the

GAG chains which are covalently attached to the aggrecan protein core and comprise

approximately 80% of the total weight of the molecule. These GAG chains, in particular

in the CS rich region are arranged in closely packed arrays creating a bottle brush

structure. Along with the hydrating properties of the GAG chains imparted by the

Page 63: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

37

charged groups along the molecule, electrostatic forces between closely packed GAG

chains provide a mechanical resistance to applied force. Electrostatic forces between CS

chains account for 50% (290kPa) of the equilibrium compressive elastic modulus as

predicted by theoretical modeling (70, 72). These electrostatic repulsion forces however

will only occur when intermolecular distances are ~5 Debye lengths or less (CS

concentration of 30mg/mL, 2-4nm spacing between chains)(69). GAG chains may come

into close proximity via several modes including adjacent chains on the aggrecan core

protein, interdigitation of opposing GAGs, and steric interactions between opposing

GAGs (Figure 3.15). Interactions between opposing GAG chains have been

experimentally demonstrated to resist force at short distances, however, when opposing

aggrecan chains are brought into contact, they exhibit enhanced force resistance at larger

distances demonstrating the benefit of the more ordered arrangement of CS chains seen in

the aggrecan bottle brush (Figure 3.16) (73-74).

Page 64: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

38

Figure 3.12: Schematic representation of hyalectans found in the intervertebral disc Ig, immunoglobulin-type repeat; LP, link-protein type module; GAG, glycosaminoglycan-binding

domain, EG, EGF-like module; Lectin, C-type lectin-like module; CR, complement regulatory protein; KS, keratan sulfate-attachment module. (59)

Page 65: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

39

Figure 3.13: The bottle brush structure of Aggrecan (top) and the chemical structure of chondroitin sulfate (bottom). Up to 130 CS chains are

covalently bound to the aggrecan core protein.(69)

Figure 3.14: Aggrecan Synthesis in vivo Aggrecan Synthesis in vivo is an intracellular process where the core protein is first translated. GAG chains are then attached to the core protein via a glycosidic bond formed between xylose and the hydroxyl group of serine. The aggrecan bottle brush is then secreted from the cell and

associates with HA and other ECM components. (67, 75)

400nm

40nm

GlcUA GalNAc

Page 66: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

40

Figure 3.15: Electrostatic repulsion generated between opposing CS GAG chains.(69)

Figure 3.16: Opposing CS and aggrecan mechanics Opposing CS and aggrecan mechanics as investigated by AFM showing increased interactions for

the macromolecular interactions. (74)

Page 67: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

41

Enzymatic degradation of aggrecan and other proteoglycans and collagens in the

NP allow for the turnover of matrix material (7). In addition to matrix synthesis, an

interplay exists between the secretion of matrix metalloproteinases (MMPs) and a

disintegrin-like and metaloprotease with thrombospondin motifs (ADMATS) by nucleus

pulposus cells and the secretion of tissue inhibitors of metalloproteases (TIMPs).

However, in the IVD where nutrition is limited, NP cells are in a state of senescence.

The reduced number of active NP cells leads to a decrease in the production of growth

factors required to stimulate matrix synthesis and are unable to produce aggrecan at the

necessary rates to maintain normal aggrecan concentration (76-77). With aging and

degeneration, enzymatic activity in the disc increases (

3.5. Aggrecan Degradation

Figure 3.17), resulting in the

presence of smaller aggrecan fragments and the loss of overall aggrecan

concentration(78). Matrix MMPs and ADMATS such as aggrecanases are the main

enzymes that contribute to the degradation of aggrecan. In particular MMPs 1, 3, 7, 9,

and 13 have showed increased activity with degeneration as well as aggrecanase 1,4,9,5,

and 15 (22, 79-80). TIMPs have also shown upregulation with disc degeneration

however several MMPs that see upregulation in the NP such as MMP 3 and MMP7

demonstrate resistance to the upregulated TIMPS 1 and 2. The major ADAMTS inhibitor

TIMP 3, does not show a concordant upregulation with degeneration, thus, for several

ECM enzymes, there is an imbalance between enzyme and inhibitor resulting in

increased aggrecan breakdown (81). Enzymatic cleavage of aggrecan is targeted to the

core protein of the molecule (Figure 3.18) (7). Several cleavage points exist throughout

the aggrecan core protein resulting in varying sized fragments of aggrecan that vary in

Page 68: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

42

functional capacity as well as ability to migrate out of the nucleus pulposus lowering the

overall tissue osmotic potential, which is dependent in part, on the macromolecular

concentration (Figure 3.19).

Figure 3.17: MMP expression in cells of the NP as measured by the percentage of immunopositive cells (22).

Page 69: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

43

Figure 3.18: Sites of enzymatic cleavage of the aggrecan core protein (7).

Figure 3.19: Aggrecan cleavage into smaller molecular weight fragments that may diffuse from the tissue. (82)

Page 70: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

44

3.6. The environment of the IVD

As the IVD grows, its relative vascularity decreases limiting its nutrient supply.

In an adult disc, a major source of nutrition is through the cartilage endplates superior and

inferior to the disc, via the vascular supply of the vertebrae. As the disc ages, endplate

calcification occurs, further compromising the already limited nutrient supply and waste

removal routes for the disc and especially, the NP. The decreased nutrient supply as well

as changes in disc pH limits the ability of disc cells to synthesize matrix and maintain cell

density. The reduced oxygen supply that accompanies lack of vasculature causes the disc

cells to undergo anaerobic metabolism, thereby increasing the production of lactic acid

and decreasing the pH of the disc (Figure 3.20) (68, 83). The pH of the disc varies with

degeneration. In discs with no visible pathologic changes, pH range is observed between

6.8-7.4, however prolapsed and scarred tissue pH has been observed as low as 5.7 to 6.3

(23, 84). The ionic concentration in the IVD is higher than surrounding tissues due to

sodium and other ions associated with immobile charged GAG chains in the NP matrix.

Typical NaCl concentration in the NP is 0.3M while that in other tissues is generally

0.15M (4, 85).

Transport through the disc endplates largely governs the disc environment.

Studies into the transport properties of cartilaginous endplates reveals a dependence of

molecule size, conformation (globular or long chain) and charge on the ability of

molecules ,to diffuse through the endplate (23). Smaller molecules (i.e. 100d) can enter

the matrix to a greater extent than larger ones (i.e. 10kd). Long chain conformations (i.e.

different MW PEG chains were investigated) are more restricted from entering the matrix

than globular conformations (i.e. starch). Dextrans of MW ranging from 4 to 7 kDa were

Page 71: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

45

also investigated for the diffusion into disc matrix through the endplate, and it was found

that distance traveled by dextran molecules per unit time decreased, again indicating the

greater mobility of smaller molecules (86). Due to the constituents of the NP and

endplate, charge also affects solute mobility where cations penetrate the NP to a greater

extent than anions (23). Endplate calcification, integrity and biochemical composition

can also alter transport where calcification can lead to build up of waste and decrease in

nutrition, but losses in endplate integrity can cause the loss of larger matrix molecules

such as proteoglycans. Hydration of the IVD matrix has also shown to be beneficial in

promoting solute movement through the IVD.

Solute transport through the disc is largely governed by diffusion, however

convective transport is also important for larger molecules. Cells rely on diffusion for

their supply of essential nutrients such as oxygen and glucose. Convective transport

arises from the diurnal loading cycle where approximately 25% of the fluid of the disc is

lost and gained over a day and night period. Convective transport is outward rather than

inward during the day’s activity and may contribute to the removal of larger species such

as TIMP-1 (28kDa), cytokines (10-40kDa) and proteoglycans (>40kDa) as investigated

theoretically. Transport in and out of the IVD is difficult to measure due to the lack of

appropriate animal models and difficulties in measurements in humans. Theoretical

modeling has given some insight, however the complexity of factors influencing solute

transport has made it difficult to validate a reliable model (23).

Page 72: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

46

Figure 3.20: Nutrient gradient across the NP a) Nutrient gradient across the NP from endplate to endplate. (b) sagital section through a human

lumbar disc with the direction of the gradient shown in (a) (83)

Page 73: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

47

Chondroitin sulfate is a member of the glycosaminoglycans (GAGs) family which

are carbohydrate polymers of repeating disaccharide units. Six different types of GAG

are known: heparin (H), heparin sulfate (HS), chondroitin sulfate, dermatan sulfate (DS),

keratin sulfate (KS) and hyaluronate (or hyaluronic acid, HA). The differences between

the molecules arise from the combination of hexosamine (D-galactosamine (GalN) or D-

glucosamine (GlcN)) and hexuronic acids (D-glucuronic acid (GlcA) or L-iduronic acid

(IdoA)) in the repeat disaccharide units. CS is comprised of a repeating disaccharide of

glucuronic acid (GlcA) and N-acetly-galactosamine (GalNAc). In CS, sulfate groups

replace one or more of the OH groups on C4 and C6 of GalNAc and C2 and C3 of GlcA

(

3.7. Chondroitin Sulfate

Figure 3.21). These modifications can generate sixteen isomers of the repeating

disaccharide, however commonly found CS are listed in Table 3.2.

3.7.1. Localizations and diverse functions of CS

Chondroitin sulfates are present as a part of PGs in several tissue locations. They

are present in intracellular granules of cells such as mast cells, at the cell membrane and

in the extracellular matrix. The function of CS and other similar GAGs is closely related

to their localization. Chondroitin sulfates and their proteoglycans have several diverse

functions including structural (69, 73) (charge repulsions), hydrating (7) (increasing

osmotic potential due to charged moieties), biological (87) (chondro-protective and anti-

inflammatory), and signaling (88) (stimulating matrix synthesis and cell differentiation).

In the matrix, GAGs contribute to the organization of the ECM and tissues while in the

Page 74: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

48

cell membrane they generally function as receptors and co-receptors and participate in the

regulation of several cellular events. The interactions depend on specific oligomeric

GAG domains of approximately five to eight monosaccharides (89). Chondroitin sulfates

are attached to proteins in various amounts and in varying ratios to other GAGs such as

DS and KS. CS proteoglycans include decorin, biglycan, fibromodulin, lumican,

aggrecan, versican, brevican, and neurocan. Several of these CS PGs are found in load

bearing tissues such as tendon, cartilage and the intervertebral disc and serve varying

functions (Table 3.3) (90). Versican in particular, which has core proteins ranging from

74kDa to 370kDa and approximately 10-30 CS/DS chains sparsely attached is shown

along with aggrecan to decrease in the NP with increasing degenerative grade (91).

3.7.2. Conformation and functional mechanics of CS

The x-ray structure of the sodium salt of pure CS-4 based on fiber

diffraction patterns is available in the protein database (PDB)

(http://www.rcsb.org/pdb/) (Figure 3.22). CS takes on a fairly extended

conformation and adopts a regular three fold helix (the calcium salt takes on a more

extended two-fold helix conformation). Intramolecular hydrogen bonds stabilize

the structure and closely packed CS chains may also interact via intermolecular

hydrogen bonds (89). The conformation of CS in solutions of varying pH and ionic

concentration have been investigated via the end-immobilzation of CS onto

substrates and subsequent analysis of CS mechanics and height using high resolution

force spectroscopy techniques and variable angle spectroscopic elliposmetry (VASE).

Page 75: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

49

The conformation of CS as well as force generated by opposing GAG layers was found to

be modulated by both solution pH and ionic concentration (Figure 3.23) (73).

CS extension (height) was observed to vary with NaCl concentration which is

consistent with the expected behavior of CS and the effects of salt screening and

molecular conformation. In DI water, the acidic groups of GAGs end tend to remain

protonated (uncharged), thus, the non-ionized GAGs do not attain an extended

conformation at very low salt concentrations and their height remains low. As NaCl

concentration increases, Na+ ions can exchange with dissociated H+ ions and ionization of

GAG charge groups can occur freely. In these conditions, CS chains tend to be in a more

extended, rigid rodlike conformation because the intramolecular electrostatic repulsion

between neighboring charges on the CS-GAG chains outweighs the entropic forces

driving the chains to a random coil (69). At very high NaCl concentrations (> 1M) the

chains collapse into a more random coil-like configuration as the intra- and

intermolecular electrostatic repulsions are screened by the salt causing a decrease in

height. The effect of charge density on GAG height is also demonstrated via changes in

solution pH where changing pH from 3 to 7 doubled the height, reaching a value close to

the CS contour length (~45nm), indicating that the CS chains were fully extended due to

increased repulsive interactions between charge groups at pH 7.

There is also a strong dependence of the CS-CS interaction forces on ionic

strength and pH suggesting that electrostatic repulsions dominate these interactions.

Repulsion between opposing GAG chains occurs over a long range, starting at ~ 150nm

distance, more than twice the CS contour length. Theoretical modeling has predicted that

~50% of the total compressive modulus of cartilage is attributable to electrostatic

Page 76: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

50

interactions while 10-15% is attributable to steric interactions of GAG chains (70, 72-73).

Lateral interactions between neighboring CS chains also contribute to the mechanical

response of CS (69).

3.7.3. Chondroitin Sulfate Stability in Aqueous Solutions

CS stability has been investigated in aqueous solutions with varied pH and

temperature (Figure 3.24) (92). CS was found to be very stable in neutral conditions and

low temperatures (30 °C), remaining stable up to 1000h while at elevated temperatures

(60 °C), CS decomposition becomes apparent around 500h. When CS was exposed to

highly acidic environments (0.1M HCl, pH 1.5), molecular mass decreased and the

number of reducing end groups increased. In addition a decrease in CS sulfation was

detected. MW decrease in acidic condition is likely due to hydrolysis of the

polysaccharide linkages. In very basic conditions (0.1M NaOH, pH 12.10), a decrease in

CS MW was also seen as well as an increase in degradation products (detected by an

increase in absorbance at 232nm corresponding to unsaturated hexuronic acid). No loss

in sulfation was detected in basic conditions. Degradation and decrease in CS MW in

basic conditions is likely due to breakdown of the glycosydic linkages due to β-

elimination reactions (β-elimination reactions form an unsaturated hexuronic acid at the

non-reducing end of each newly formed chain). The combination of elevated

temperature and extreme pH results in accelerated degradation of CS. In acidic

conditions at 30̊C, no depolymerization of CS is seen up to 960h while total degradation

Page 77: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

51

occurred after 96h in 60̊C acidic conditions. Similarly combined basic treatment and

elevated temperature (60̊C) resulted in CS degradation after approximately 190h. (92)

3.7.4. Age related changes in NP GAG

CS, KS, and HA are major GAG constituents of the NP. Aggrecan has both KS

and CS regions, and the aggregation of aggrecan on HA is mediated by link protein. The

ratio of CS to KS changes in the IVD with aging and degeneration. The very large

aggregating GAG, HA also shows a marked increase with respect to total GAG ratio

which may have implications in matrix properties (93). Younger discs have a greater

presence of CS while older discs lose CS and gain KS (almost 60% in older discs) (93).

CS and KS are based on the same polymer, and their secondary structures are

very similar in solution, however the synthesis of these distinct GAGs follows two

different metabolic routes. KS and CS are both produced from glucose. Biosynthesis of

KS converts glucose to galactose with no consumption of oxygen while formation of CS

requires the oxidation of glucose to glucuronic acid. It is suggested that KS may be

preferentially synthesized in aging discs as a functional substitute for CS where oxygen

supplies are reduced. Higher lactate concentrations in the disc due to disrupted nutrient

and waste transport pathways may also change the balance of CS and KS synthesis (93).

KS however is shorter in length than CS and the function of this GAG in maintaining the

mechanical properties of the NP is unknown.

In addition to changes in KS to CS ratio, the structure of CS in aged NP also

changes. A decrease in average chain size occurs between fetal and early postnatal ages

Page 78: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

52

and at maturity, the CS chain size can decrease from 20kDa to as low as 8kDa. In

cartilage, the ratio of CS-4 to CS-6 is also observed to change with an increase in CS-6 in

older tissue (63). Age-related changes in GAG content may affect both the hydrating and

mechanical properties of this molecule and may also affect its interaction with other NP

ECM molecules.

Changes in GAG content in the NP is likely due to changes in synthesis and not

enzymatic or hydrolytic degradation of the GAG chains themselves. Enzymatic

degradation of GAGs are largely mediated by glycosidases (i.e. hyaluronidases),

hydrolytic enzymes present within lysosomes (89). In mammals, the enzymes are found

in the liver, spleen, kidney, testes, urine, plasma and skin. Hyaluronidases cleave bonds

within the polysaccharide chain and produce oligosaccharides of varied length, however,

lysosomal enzymes function only in acid pH ranges (3.3-5.5) (94). The GAGs are

generally stable at physiological pH and the slightly acidic environment of the NP.

3.7.5. Pharmaceutical application.

Chondroitin sulfate is a multi-functional biomolecule utilized in various

pharmaceutical applications (89, 95-96). Currently, CS is recommended by the European

League against Rheumatism (EULAR) as a Symptomatic Slow Acting Drug for

Osteoarthritis (SYSADOA) drug in Europe in the treatment of knee osteoarthritis (OA)

and is often used in conjunction with glucosamine as a dietary supplement as a safe and

effective option for the treatment of OA (96).

As a pharmaceutical agent, CS is usually ingested as an intact polymer. CS of

various types and sources have varying degrees of bioavailability, where for example,

Page 79: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

53

tracheal CS causes a rapid increase in plasma CS levels with a peak within 1-5h

following administration, and returning to baseline after ca. 10h. Shark CS on the other

hand has a slower uptake with peak levels not occurring until ca. 16h following

administration (95). Although bioavailability data of orally administered CS is difficult

to acquire, the overall bioavailability of exogenously delivered CS is estimated around

10-20% (95).

The direct treatment of symptomatic disc degeneration with glucosamine and CS

supplementation has been limited likely due to the reduced vascularity of the NP of the

IVD (the largest avascular structure of the body). However, in one reported case study, a

56 year old man who presented with more than 15 years of frequent/recurrent low-back

pain was treated for 2 years with a glucosamine/CS supplement and observed to have

increased water retention (and less bulging/protrusion) in partially degenerated discs

while fully degenerated discs showed no change (97). Along with the MR imaging (T2

weighted MR imaging) the patient reported a gradual improvement in the range of

motion and functioning of his back as well as an increased capability of withstanding

heavier workloads without pain (97). Although this study is not conclusive, it

demonstrates the possible efficacy of administering CS for the rehydration and functional

recovery of mildly degenerated discs, and their monitoring using T2 MR imaging. The

direct administration of organized CS structures may provide added, less transient and

more direct mechanical benefits.

Page 80: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

54

Figure 3.21: CS repeat disaccharide with possible sulfation points indicated in red. (95)

Page 81: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

55

Figure 3.22: X-ray diffraction structures of CS-4 (A) side and (B) end views shown in stick representation with carbon in green, oxygen in red,

nitrogen in blue and cation in purple. Water in the crystal structure and sodium ions are shown as star shapes. The helix makes a turn every three disaccharides. (89)

Page 82: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

56

Figure 3.23: Ellipsometric measurement of end-grafted GAG height (A) Ellipsometric measurement of end-grafted GAG height versus NaCl concentration at pH

~5.6 for two different grafting densities (s). (B) end-grafted GAG height at pH 3 and pH 7 ([NaCl] = 0.015M) at a grafting density of 6.5nm. (C) Force versus separation distance for GAG-functionalized tip versus GAG-functionalized substrate (3 = 6.5nm) at pH 5.6 and varying NaCl

concentration (0.0001M, 0.001M, 0.01M, 0.1M, 1M). (D) Force versus separation distance at 0.015M NaCl and varying pH. (45nm full extended contour length, repulsive forces observed to

start at ~150nm) (73).

Page 83: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

57

Table 3.2: Commonly found chondroitin sulfate isomers

Name Site of Sulfation Systematic Name Chondroitin sulfate A carbon 4 of GalNAc sugar chondroitin-4-sulfate (CS-4)

Chondroitin sulfate C carbon 6 of the GalNAc sugar

chondroitin-6-sulfate (CS-6)

Chondroitin sulfate D carbon 2 of the glucuronic acid and 6 of the GalNAc sugar

chondroitin-2,6-sulfate

Chondroitin sulfate E carbons 4 and 6 of the GalNAc sugar

chondroitin-4,6-sulfate

Table 3.3: Chondroitin sulfate proteoglycans and their function in tendon (90)

Page 84: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

58

Figure 3.24: Stability of CS in elevated pH and temperature conditions (92). (A) Absorbance at 232nm and (B) reducing end group as a function of time for CS treated under basic and acidic conditions. (C) Abosrbance at 232nm and (D) reducing end groups as a function of time of CS under neutral conditions (aldehyde detected by reducing sugar assay indicating CS de-polymerization, and absorbance at 232nm detects degradation product, unsaturated hexuronic

acid)

Page 85: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

59

The hydrating and biological functions of CS has made it attractive for the use in

many applications including tissue engineering (98), tissue adhesion (99), drug delivery

(100) and wound healing (101). Several biologic and synthetic based hydrogels with

immobilized CS have been formed by the modification of the CS disaccharide.

Incorporated cross-linkable groups include methacrylates (98, 102), NHS (99), and

dialdehydes (via periodate oxidation) (101) (

3.8. Chondroitin Sulfate networks

Figure 3.25). CS has also been physically

associated into hydrogel networks (100, 103). The introduction of chemical and physical

crosslinks into CS allows the cross-linking of CS to other natural or synthetic polymer

systems to form a hydrogel with enhanced swelling and biological function (Figure 3.26).

As an example, methacrylates were introduced into CS by the modification of the CS

hydroxyl with glycidyl methacrylate (reacted at room temperature for 15 days, pH 7.4)

(102). Methacrylated CS was subsequently photopolymerized using the UV

photoinitiator (Irgacure 2959) and 365-nm light at ~10mW/cm2. Hybrid hydrogels were

also formed by the reaction of the methacrylated CS with poly(ethylene oxide)-diacrylate

(PEODA). Chondrocytes were encapsulated in the gel matrix and found to remain viable

and hydration of the polymer system was enhanced by CS incorporation. Although cross-

linked CS hydrogel systems show enhanced swelling and biological function, they do not

incorporate the ultra-structural organization of CS in the aggrecan bottle brush structure

which has been shown to be critical in the realization of the molecules mechanical

function.

Page 86: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

60

Figure 3.25: Modified CS for the incorporation in cross-linked CS networks (A) dialdehyde (101), (B) CS-NHS (99), (C) CS-methacrylate (104)

Figure 3.26: Chemically/Physically cross-linked CS networks (105)

Page 87: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

61

The synthesis of synthetic-based bottle brush polymers or “molecular bottle

brushes” has been extensively studied and reviewed (106-107). The three main methods

are the grafting-to, grafting- through and grafting-from methods of synthesis (Figure

3.27). In grafting-to bottle brush bristles in the form of a monotelechelic polymer is

attached to a functionalized polymeric backbone (108). In the grafting-through strategy,

a macromonomer is combined with initiator in order to induce polymerization of the

polymerizable end of the macromonomer building the polymeric backbone as the

macromonomers are joined together, often via free-radical polymerization(109). In a

third strategy, grafting-from, a macroinitiator polymeric backbone is combined with

monomer which is subsequently polymerized off of the backbone via initiation and

propagation of the free-radical generated by the initiator. Each of these strategies has its

limitations and benefits and controls different structural parameters such as chemical

composition, grafting density, degree of polymerization of side chains, and degree of

polymerization of the backbone, allowing for the fabrication of a family of bottle brush

molecules with varying structures and properties (

3.9. Synthetic bottle brush polymers and aggrecan mimetics

Table 3.4). In addition to densely

packed brushes, sparse brushes, stiff brushes, flexible brushes, multi grafted brushes,

gradient brushes, stars and networks may also be formed. The Ortiz group, for example

(110-112), fabricated a family of end functionalized polymer brushes of poly(2-

hydroxyethyl methacrylate-g-ethylene glycol) with varying backbone length, and brush

grafting density and demonstrated their mechanical properties in the presence of various

stimuli. Synthetic glycopolymer brushes have been fabricated with short monosaccharide

or oligosaccharide side chains which impart biological function to the polymers, however

Page 88: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

62

side chain length is short in comparison to the CS chains of aggrecan inhibiting their

mechanical function(113-115). Attempts have also been made for the fabrication of

proteoglycan-like cylindrical glycopolymer brushes (116) as well as brushes with charged

sulfonate bristles (117). These brush structures emulate the architecture of the aggrecan

brush structure, however, the fully synthetic systems are not able to mimic the biological

activity of the natural biomolecule.

Page 89: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

63

Figure 3.27: The three main strategies for the fabrication of bottle brush polymers (106).

Route Method

“grafting-to” Chemical modification of pre-formed polymer with a sugar-containing reagent

“grafting-through”

Sugar-bearing monomers are polymerized into linear polymers with pendant sugar

moieties

“grafting-from” Polymeric backbone used as a solid support for enzymatic or chemo-enzymatic

oligosaccharide synthesis.

Page 90: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

64

Table 3.4: Pros and cons of the different routes to bottle brush synthesis

Page 91: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

65

The fabrication of synthetic and natural based glycopolymers is also under

considerable investigation (113-114, 118-119). Although glycopolymers refer to a wide

range of materials including natural and artificial carbohydrate-containing polymers, as

well as synthetically modified natural sugar-based polymers, glycopolymers can be more

narrowly defined as polymers with pendant saccaride residues. The fabrication of

glycopolymers can also follow the “grafting-to”, “grafting-through” and “grafting-from”

strategies. In the “grafting-to” technique glycopolymers are formed by the chemical

modification of preformed polymers or polymeric substrates with sugar-containing

reagents. Although this technique is more simple, it often results in less regular

structures with limited glycosylation.

3.10. Glycopolymers

In the “grafting-through” technique, sugar-bearing monomers are polymerized

into linear polymers with pendant sugar moieties (Figure 3.28). Polymerization can be

achieved through several mechanisms including ring-opening metathesis polymerization

(ROMP) (120), reversible addition–fragmentation chain transfer (RAFT) polymerization

(121), atom transfer radical polymerization (ATRP) (122) and cyanoxyl-mediated free-

radical polymerization (Figure 3.29) (123-126). The “grafting-through” synthesis of

glycopolymers however has added complexity compared to the fabrication of fully

synthetic bottle brush polymers. Carbohydrates contain reactive groups such as

carboxylic acids and charged sulfates that often require protection during the

polymeriziation reaction to prevent side reactions. The final glycopolymer then requires

further de-protection to return the carbohydrate side chains to their natural state. The

Page 92: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

66

solubility of large carbohydrates in organic solvents is also limited, limiting the types of

polymer solvent systems that can be utilized in glycopolymer synthesis.

In a “grafting-from” technique, synthetic saccharide chains can be enzymatically

grown from a polymeric substrate using the polymeric backbone as a solid support for

enzymatic or chemo-enzymatic oligosaccharide synthesis. Two classes of biocatalysts

are available for the enzymatic synthesis of glycosides, glycosidases and

glycosyltransferases and the formation of glycosidic bonds (127-128). Stereoselectivity

is a concern in the growth of oligosaccharides and care must be taken in to account to

form the desired alpha or beta linkages. Glycosidic linkages must also be formed with

high regioselectivity which requires the use of protecting groups (127). Although this

method is most reminiscent of natural proteoglycan and GAG synthesis the degree of

polymerization of the sugar side chains may be limited and the sugars will be synthetic in

nature.

Several strategies have been developed to overcome the complexities associated with

glycopolymer synthesis, and many families of glycopolymers have been formulated with

mono- and disaccharide side chains (123-125), however natural CS has yet to be

synthesized into a bottle brush polymeric structure. See Appendix 1 for a detailed

literature review.

Page 93: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

67

Figure 3.28: Sugar bearing polymerizable monomers (7) and examples of acrylate and methacrylate polymerizable sugar monomers. (113)

Figure 3.29: Cyanoxyl-mediated copolymerization of vinyl-glycomonomers with acrylamide as reported by Grande et al (125)

Page 94: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

68

In order to design a minimally invasive early interventional stragey for the

treatment and prevention of discogenic back pain, with lasting mechanical and biological

effects, it is important that the delivered material be resistant to the native IVD

environment and function within the slightly degenerated environment. It is also critical

that the macromolecular ultra-structural organization of CS into the bottle brush structure

be maintained. The state of the healthy and degenerated intervertebral disc has been

thoroughly reviewed, and the mechanical contribution of the major biomacromolecules

aggrecan and CS assessed (

3.11. Summary

Table 3.5). Based on our understanding of these systems we

have proposed the synthesis of a biomimetic hybrid synthetic/bio based CS brush

structure for the replacement of aggrecan in the degenerating disc that provides a

structural improvement over currently synthesized CS architechtures. Following is the

contribution of this thesis to the design and development of biomimetic aggrecan

macromolecules with CS bristles.

Page 95: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

69

Table 3.5: Key functions of aggrecan and its components

Page 96: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

70

Chapter 4 : Objective and Specific Aims

Several parameters have been set forth for the design of a biomimetic aggrecan

macromolecule that can restore the function of the nucleus pulposus in the early stages of

intervertebral disc degeneration (Table 4.1). In order to address these design criteria, the

fabrication of an enzymatically resistant, hybrid synthetic/bio-based biomimetic

aggrecan macromolecule, for use in the treatment of intervertebral disc degeneration

has been investigated (Figure 4.1). Polymer synthesis and bioconjugation techniques for

the fabrication of the macromolecule followed by functional characterization of the

macromolecules is reported. A key component of this project is the incorporation of

chondroitin sulfate in biomimetic aggrecan biomacromolecule synthesis. In order to

achieve the objective set forth in this proposal, several specific aims were established:

Specific Aim 1 (Chapter 5)

Identified a handle for the terminal end functionalization of chondroitin sulfate

and investigated its utility in the fabrication of biomimetic brush structures. The

fabrication of bottle brush polymer with natural chondroitin sulfate side chains requires

the terminal-end immobilization of CS. Commercially available natural CS was

investigated for a terminal primary amine (PA) that may be present as a result of CS

isolation from donor tissues (89, 129-130).

The terminal PA was then investigated for its reactivity to different functional

groups for the formation of covalent linkages with synthetic components. The feasibility

of utilizing both the “grafting-to” and “grafting-through” techniques of bottle brush

synthesis was explored. CS was immobilized onto model surfaces as well as a model

Page 97: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

71

polymeric backbone for investigation of the “grafting-to” strategy. For investigation of

the “grafting-through” strategy of synthesis, vinyl monomers were conjugated to the

terminal primary amine of CS for the synthesis of terminal-end polymerizable CS and

options for the free-radical polymerization of vinyl-terminated CS were discussed. The

terminal primary amine of CS has been used previously for the immobilization of CS onto

solid substrates (69, 130), however its use in polymer synthesis strategies has not yet

been explored.

Specific Aim 2 (Chapter 6)

Utilizing the CS terminal amine-epoxide reaction investigated in Specific Aim 1,

explored the fabrication of CS-macromolecles via the “grafting-through” strategy. The

step-growth reaction between epoxides and amines was investigated for the synthesis of

CS-macromolecules. Since Step-growth polymerization can be conducted in aqueous

media and under moderate reaction conditions and does not require additional initiators

for reaction, this reaction strategy may provide a facie route to the synthesis of

biomimetic aggrecan. CS was reacted with di-epoxides and reaction kinetics monitored

using the fluorescamine assay. Resulting reactions were purified and characterized for

their chemical composition (via 1H-NMR, and ATR-FTIR) and their molecular size via

GPC.

Specific Aim 3 (Chapter 7)

Characterized the functional properties of the synthesized CS-macromolecules.

In order to characterize the functional material properties of the synthesized

biomacromolecule the physical structure and fixed charge density of the CS-

Page 98: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

72

macromolecules was investigated and viscosity and osmotic pressure of the polymer

solutions were compared to that of CS. In order to investigate the biological

compatibility of the synthesized biomacromolecules, fibroblast and nucleus pulposus cell

cultures were incubated with CS-macromolecule solutions and investigated for cell

viability.

Page 99: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

73

Figure 4.1: Schematic of Biomimetic Aggrecan Design. See Appendix 2 for alternate view of the CS terminal handle.

Page 100: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

74

Table 4.1: Design of Biomimetic Aggrecan

Function Mode of Action Biomimetic Aggrecan Criterion

Structural

Should be retained by NP matrix Large Size

Should be stable in the NP environment Enzymatic/Hydrolytic Resistance

Mechanical Should induce electrostatic repulsions

Close Proximity of Charged Groups ~2-4nm spacing between adjacent

GAG chains

Hydrating Should generate an osmotic potential

High Fixed Charge Density High Osmotic Potential

Biological Should maintain a regenerative environment Biocompatible

Clinical

Should be minimally invasive Injectable

Should be cost effective Therapeutic dosage reasonably priced for multiple treatments

Page 101: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

75

Chapter 5 : Identification of a Chondroitin Sulfate Terminal Handle and the Investigation of its Utility in the Fabrication of Biomimetic Brush Structures

Discogenic lower back pain is one of the leading musculoskeletal disorders

confronting our health system (131). Intervertebral disc (IVD) degeneration is an

irreversible process, which leads to a loss in mechanical stability with the potential for

neurologic compromise. The most common surgical approaches treat the end-stage of

this degeneration and include fusion and total disc replacement (132). Earlier stage

interventions, which address degeneration of the nucleus pulposus (NP), or inner region

of the IVD may be employed to mechanically stabilize the IVD and alleviate lower back

pain (9-15).

5.1. Introduction

With age and degeneration, a loss of nutrient supply to the disc results in a

reduction of the synthesis of aggrecan, a major proteoglycan of the nucleus pulposus

which generally comprises approximately 14% of the NP wet weight (18, 61). In

addition, there is a concordant increase in enzymes in the NP that attack the aggrecan

core protein, resulting in fragmentation of the aggrecan molecule (22, 79-80, 133). It has

been shown that the proteoglycan content of the NP decreases linearly with age (134). In

addition, other investigators have localized a decrease in sulfated chondroitin in the

nucleus pulposus portions of the disc (135). This loss of proteoglycans leads to a loss of

Page 102: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

76

hydration of the nucleus manifested by a reduction of disc height and intradiscal pressure

(136).

Aggrecan, an aggregating proteoglycan has two main mechanical functions in the

disc: 1) it allows water uptake by the nucleus due to sulfated groups in the chondroitin

(CS) and keratin sulfate rich regions which, in part, provides intradiscal pressure (7) and

2) it provides electrostatic repulsion due to the elegant 3D macromolecular bottle brush

structure, which contributes to intradiscal pressure and disc height (69-71, 75, 137-138).

Approximately 100 CS glycosaminoglycan chains are covalently attached via a

glycosidic bond to the core protein in the CS region with a grafting density of

approximately 0.25 to 0.5 nm-1 resulting in a densely packed bottle brush structure(67).

Each CS chain, which consists of 10-50 repeating disaccharide units of glucuronic acid

(GluUA) and n-acetylgalactosamine (GalNAc), is approximately 20 kDa (65, 139).

CS has both mechanical and biological function in the intervertebral disc, as well

as other tissues such as cartilage (7, 69, 73, 140-141). Strategies for the replacement of

CS in various formulations have been previously investigated including solutions of CS

with hyaluranon and the modification of hydrogels with CS (98, 101, 105, 142). The

incorporation of CS into hydrogel systems often requires the modification of the CS

disaccharide allowing for the attachment of the CS chain to the hydrogel matrix

throughout numerous points of the CS backbone (101). Biological function is enhanced

by hydrogel delivery of CS but these strategies do not recapitulate the ultra-structural

bottle brush organization of CS found to be mechanically important in native tissue(143).

Page 103: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

77

The fabrication of synthetic and natural based glycopolymers is also under

considerable investigation (113-114, 118-119). Several glycopolymers have been

formulated with mono- and disaccharide side chains (123-125). Glycopolymers have

been demonstrated to have biological function, however their use as a mechanically

stabilizing material has not been investigated and synthesis has been limited to short

sugar side chains. Glycopolymers can be synthesized via the “grafting-to” or “grafting-

through” techniques. In the “grafting-to” technique glycopolymers are formed by the

chemical modification of preformed polymers or polymeric substrates with sugar-

containing reagents. Although this technique is straightforward, it often results in less

regular structures with limited glycosylation. In the “grafting-through” technique, sugar-

bearing monomers are polymerized into linear polymers with pendant sugar moities.

Polymerization can be achieved through several mechanisms including ring-opening

metathesis polymerization (ROMP) (120), reversible addition–fragmentation chain

transfer (RAFT) polymerization (121), atom transfer radical polymerization (ATRP)

(122) and cyanoxyl-mediated free-radical polymerization (123-126). The majority of

glycopolymers have been synthesized by the polymerization of sugar-containing vinyl

monomers, however the preparation of the polymerizable sugar-monomers is often a

complex and multi-step process that requires chemical modification of the sugar moiety.

Natural CS has yet to be synthesized into a bottle brush polymeric structure.

All glycosaminoglycans, with the exception of HA, are naturally synthesized

intracellularly, as proteoglycans, i.e. with a covalent linkage to a core protein (144). CS

is linked to the aggrecan core protein at its L-serine residue where a glycosidic bond is

formed between the linkage CS saccharide xylose and the hydroxyl group of serine (67)

Page 104: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

78

In the isolation of CS, the proteoglycan core protein can be broken down by either

enzymatic digestion or alkaline treatment releasing the GAG chains and leaving the CS

chain with its linkage region intact and still connected to an amino acid residue (89, 129-

130) (see Appendix 2 for alternate view of CS and the CS terminal amino acid residue).

The resulting terminal protein residue on the CS chain has been utilized to immobilize CS

onto agarose membranes using the 1-cyano-4-(dimethylamino)pyridinium

tetrafluoroborate (CDAP) activating agent to form an isourea bond (130). CS isolated

from rat chondrosarcoma cell cultures and digested with proteinase K, to maintain the

terminal amino acid residue, was also utilized to end-graft CS by first reacting the

terminal reactive amine group with dithiobis(sulfosuccinimidyl propionate (DTSSP) then

reducing it to a thiol group for coupling to gold substrates (69, 73). These studies

demonstrate the utility of the CS-terminal amino residue.

Commercially available CS however is inhomogeneous in its physical and

chemical properties. Depending on the isolation and purification techniques used, CS has

variable structure, properties, purity, molecular weight and molecule integrity (89, 95).

Parameters reported in association with CS production are limited and mainly consist of

elementary chemical composition, impurities and general ranges of MW (139). No

information is provided on the terminal end of CS or if the competence of the linkage

region of CS after the isolation procedure.

In this work we investigated commercially available CS for a terminal amino acid

handle. A terminal handle may be utilized in order to immobilize CS into brush

structures which requires the single point immobilization of CS versus the multi point

immobilization of CS utilized in network formation. The terminal primary amine of CS

Page 105: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

79

was then investigated for its utility in the immobilization of chondroitin sulfate on model

surfaces, to polymerizable monomers and to a synthetic polymeric backbone. The

terminal primary amine of CS was utilized to immobilize CS onto functionalized glass

surfaces to investigate the utility of the CS terminal primary amine in the immobilization

of CS via epoxide-amine and aldehyde-amine reactions. CS was also modified at the

primary amine via amine-epoxide, amine-aldehyde, and amine-carboxylic acid

interactions for the synthesis of polymerizable vinyl-CS chains. The carboxylic acid-

amine reaction was investigated for the synthesis of CS-brush structures via the “grafting-

to” technique using a poly(acrylic acid) polymer backbone. The studies described here

demonstrate a facile route to the end functionalization and immobilization of

commercially available CS. Based on these studies, several routes to the synthesis of

chondroitin sulfate based biomimetic aggrecan brush structures have been proposed.

Chondroitin-4-sulfate, Chondroitin-6-sulfate and mixed CS powders were

purchased from two different vendors and originated from different animals and tissues

(

5.2. Materials and Methods

Table 5.1). L-serine, bovine serum albumin and N-acetyl-D-galactosamine were

purchased from Sigma Aldrich and used as control samples. Fluorescamine and all other

reagents were purchased from Sigma Aldrich and used as received.

Page 106: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

80

5.2.1. Determination of CS primary amine content

A fluorescamine assay was developed and validated for use in the determination

of the primary amine (PA) content of CS. Fluorescamine, a fluorometric reagent (Figure

5.2), reacts directly with primary amines to yield highly fluorescent derivatives (390nm

excitation, 475 nm emission) whose resulting fluorescence is proportional to the amine

concentration (145). Reactions of fluorescamine with primary amines proceed at room

temperature with a half-time of a fraction of a second while excess reagent is destroyed

with a half-time of several seconds (146). The fluorescamine reagent is particularly

suitable for the detection of a single PA in CS as it is sensitive to PAs in the picomolar

range(145).

In a general procedure, 150ul of sample solution is plated in triplicate in a flat

bottom black 96-well plate. Samples must be prepared in an amine free buffer; buffer

type and pH may effect fluorescence readings. An alkaline pH range of 8-9.5 is generally

ideal for the fluorescamine-amine reaction where lower pH may result in protonated

amines, making them unavailable for reaction and high pH may cause hydrolysis of the

fluorescamine reagent before the fluorogenic reaction can occur (147). Fluorescamine

solution is prepared at 3mg/ml in DMSO at a minimum of 15mg in 5ml. Fluorescamine

is prepared in a non-aqueous solvent in order to fully solublize the reagent and avoid

immediate hydrolysis (146). A 50µl aliquot of the fluorescamine solution is added to the

sample solutions then allowed to react for 5min at room temperature with constant

agitation on a shaker plate. Sample fluorescence was measured on an Infinite M200

TECAN spectrophotometer with excitation/emission of 365/490nm and gain of 80

(fluorescent sweeps were used to optimize excitation and emission wavelengths for

Page 107: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

81

fluorescamine fluorophor detection). Sample free buffer solution with fluorescamine

reagent was used as a blank for fluorescence readings.

Detection of primary amines with fluorescamine was calibrated with solutions of

known amine content. L-serine (Figure 5.1, MW, 105.09) which is the attachment site for

CS to the aggrecan core protein, and contains only one PA per molecule, was used to

establish a fluorescence vs. [PA] standard curve. The number of PA/molecule of a

sample can be calculated from the linear region of the L-serine standard curve. Bovine

serum albumin (MW, 66,432 Da) (BSA) was used as a positive control biomolecule with

a known high PA content and N-acetyl-D-galactosamine (Figure 5.1, MW, 221.2 Da) was

used as a negative control as it is a secondary amine containing component of CS but

contains no PAs (Figure 4.1). Samples were solublized in sodium borate buffer (SBB,

0.1M, pH 9.4) then diluted in a serial dilution and assayed with the fluorescamine

reagent. Fluorescence was normalized to SBB blanks and samples were taken in

triplicate. The number of primary amines per molecule was calculated at each

concentration from the linear region of the L-serine standard curve.

CSs from various sources were assayed for their primary amine (PA) content

using the established fluorescamine assay. Chondroitin-4-sulfate sample MW as

estimated to be 22 kDa while primarily chondroitin-6-sulfate samples were estimated to

have a MW of 65 kDa (137, 148).

CS was first solublized in sodium borate buffer (0.1M SBB, pH 9.4) at various

concentrations. Then, 150µl samples of CS solution were added to a 96-well plate and

incubated with 50µl of 3mg/mL fluorescamine solution (in DMSO) with constant

Page 108: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

82

agitation for 5min. Fluorescence was normalized to SBB blanks and samples were taken

in triplicate.

5.2.4. Modification of glass surfaces with Chondroitin sulfate

Epoxide or aldehyde functionalized glass slides (Genetix) were utilized for the

immobilization of CS (Figure 5.6). Upon reaction of an aldehyde with a primary amine in

alkaline conditions (pH greater than 9.0) and with reductive amination with sodium

cyanoborohydride, a stable secondary amine bond will form (Figure 5.3) thus

immobilizing CS to the aldehyde glass surface. Epoxides will react with amines also at

alkaline pH via the opening of its oxirane ring (Figure 5.4) (149) immobilizing CS to the

epoxide surface via epoxide amine reactions. Epoxides may also react with other

functional groups such as hydroxyls or sulfyhryls, however, a pH range of 9.0 to 11 is

more chemoselective for amine reactions (149). Functionalized glass slides as well as

non-functionalized glass slides were cut into 1cm x 1cm samples, placed in 6-well plates

and soaked in 3mL of CS solution in SBB pH 9.4 at varying CS concentration. Sodium

cyanoborohydride was added to aldehyde slide samples as described previously. Well

plates were placed on a shaker plate and samples were allowed to incubate with constant

agitation for 4h at room temperature. Slides were subsequently rinsed thoroughly (3 X in

SBB) then placed in 3mL of SBB and allowed to rinse for an additional 3h with constant

agitation to remove any non-covalently bound CS. Slides were rinsed an additional three

times in SBB then allowed to air dry. Non-functionalized glass slides were similarly

prepared as control samples.

Page 109: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

83

5.2.5. Characterization of chondroitin sulfate modified surfaces

CS modified glass slides were prepared for analysis with Energy-dispersive X-ray

(EDAX) microanalysis using an environmental scanning electron microscope (ESEM,

FEI XL30 with EDS). Glass slides were sputter coated with carbon (25sec) then imaged

with back scatter in the ESEM at (voltage, 6 keV, spot size 6, working distance 12 mm,

magnification 1000X corresponding to an area of 121µm x 95µm). Element peaks were

identified in the spectra corresponding to C, O, Na, Mg, Al, Si, S, K, and Ca and spectra

were automatically base-lined and peak areas quantified (spectra collected for a duration

of 210 live seconds). Changes in sample sulfur content (detected at 2.31 keV) with CS

modification was investigated as CS has a large sulfur component. In a semi-quantitative

analysis, the net integrated area for sulfur was normalized to the Si net integrated area

where Si is a major component of the glass slides and should not change greatly with CS

modification. Separate samples were analyzed for their hydrophilicity using a contact

angle goniometer with water as a medium. Three readings from triplicate samples of

epoxide, aldehyde and un-functionalized slides soaked in CS and buffer solutions were

taken. A decrease in contact angle is indicative of an increase in surface hydrophilicity

imparted by charged CS molecules covalently attached to the glass surfaces.

5.2.6. Chondroitin sulfate-monomer conjugation

Several amine reactive chemistries were investigated for the conjugation of

monomers to the CS chain at its terminal primary amine end. The monomers acrolein,

allyl glycidyl ether (AGE) and acrylic acid (Sigma Aldrich) were used. Acrolein contains

Page 110: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

84

an amine reactive aldehyde functionality and a vinyl group, which is polymerizable via

free radical polymerization (149). AGE, is an epoxide and vinyl containing monomer,

which will react with amines at alkaline pH (149). Acrylic acid is a polymerizable

monomer with carboxylic acid functional groups. The carboxylic acid group of acrylic

acid can be activated to be highly reactive with amines using well characterized

EDC/sulfo-NHS coupling reactions (Figure 5.5) (149).

For the conjugation of CS to acrolein and AGE, solutions of monomer at various

concentrations in 0.1M SBB, pH 9.4, were mixed with solutions of CS in 0.1M SBB pH

9.4 to achieve varying monomer:CS molar ratios. For acrolein-CS samples, sodium

cyanoborohydride (5M in 1N NaOH, Sigma) was added at 20µL/mL in order to stabilize

the formed Schiff base to a secondary amine bond (149). Acrylic acid was first activated

with EDC/sulfo-NHS (2 mM EDC and 5 mM sulfo-NHS in MES buffer (0.05M MES,

0.5M NaCl), pH 6.0) for 15 min followed by quenching of excess EDC with 2-

mercaptoethanol (10 min at a final concentration 20 mM). Activated acrylic acid was

then combined with CS (CS in phosphate buffered solution, pH 7.4) at varying

monomer:CS molar ratios. All monomer-CS solutions were placed on a rotator

(Thermolyne Labquake) and allowed to mix at least 4h. CS-Acrylic acid solutions were

then filtered using a sephadex G-25 pre-packed desalting column equilibrated in PBS

(PD-10, GE Healthcare) to remove excess reactants. 150ul samples of each solution were

taken in triplicate and assayed with the fluorescamine assay as described in section 2.1

for their PA content. The percentage of PAs in the CS sample conjugated to monomer is

indicated by the percent decrease in PA content which was calculated as:

Page 111: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

85

𝑃𝑒𝑟𝑐𝑒𝑛𝑡 𝐶𝑜𝑛𝑗𝑢𝑔𝑎𝑡𝑖𝑜𝑛 = 100% ∗[𝑃𝐴]𝑖𝑛 𝐶𝑆 𝑤𝑖𝑡ℎ𝑜𝑢𝑡 𝑚𝑜𝑛𝑜𝑚𝑒𝑟 − [𝑃𝐴]𝑖𝑛 𝐶𝑆 𝑤𝑖𝑡ℎ 𝑚𝑜𝑛𝑜𝑚𝑒𝑟

[𝑃𝐴]𝑖𝑛 𝐶𝑆 𝑤𝑖𝑡ℎ 𝑚𝑜𝑛𝑜𝑚𝑒𝑟

A decrease in the PA content of the CS-monomer solutions with respect to CS

without monomer is indicative of binding of the monomer to CS at the primary amine.

Solutions of monomer without CS were also assayed with the fluorescamine reagent and

found to exhibit no appreciable fluorescent signal at the excitation/emission of interest.

5.2.7. Proton nuclear magnetic resonance (1H-NMR) and Fourier transform infrared

spectroscopy (FTIR) analysis of AGE-CS conjugation over time

Samples of CS-AGE conjugate prepared at the 1000:1 ratio were investigated

with 1H-NMR to further demonstrate CS-monomer conjugation. Samples were prepared

as described previously followed by extensive dialysis against DI water (regenerated

cellulose membrane, MWCO 6-8K) in order to remove un-reacted monomer. Purified

samples were lyophilized reconstituted in D2O at approximately 30mg/mL. 1H-NMR

spectra were taken on a 300MHz NMR spectrometer (UNITYNOVA) at 64 scans and at

ambient temperature. Lyophilized samples were also investigated on an attenuated total

reflectance (ATR)-FTIR (64 scans) for chemical composition and compared to that of

natural CS.

5.2.8. Conjugation of Chondroitin Sulfate to a poly(acrylic acid) polymeric backbone

Page 112: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

86

The polymeric backbone of poly(acrylic acid) (PAA) was chosen for preliminary

investigation of the synthesis of CS-glycopolymer structures via the “grafting-to” strategy

via the carboxyl-amine reaction (Figure 5.7). PAA of 250kDa MW was purchased from

Sigma (35wt% in H2O, 1.6mM, Sigma #416002) in order to mimic the MW of the natural

aggrecan protein backbone. For reaction with CS, PAA (33mg/mL, 0.132mM PAA) was

first activated with EDC/sulfo-NHS (2 mM EDC and 5 mM sulfo-NHS in MES buffer

(0.05M MES, 0.5M NaCl), pH 6.0) for 15 min followed by filtering using a sephadex G-

25 pre-packed desalting column (PD-10, GE Healthcare) to remove excess reactants.

Activated PAA was then combined at a 1:1 v/v ratio with a 20mgml solution of CS

prepared in PBS. Solution pH was then adjusted with NaOH to pH 7.4 and the mixture

was allowed to react over time mixed continuously (Thermolyne Labquake). Final PAA

and CS concentrations were 16.5mg/ml (0.066mM) and 10mg/ml (0.45mM) unless

otherwise noted (reactant ratio of 6.8:1 CS;PAA). Reaction conditions were varied

including, time, temperature (4̊C, 21̊C and 37̊C), ionic concentration and CS to PAA

molar ratio. Conjugation of CS to the PAA backbone was monitored using the

fluorescamine assay, where a loss in the PA signal indicates a loss of the PA to the

reaction. Reaction was also monitored by monitoring reaction solution viscosity during

the conjugation period where an increase in solution viscosity is indicative of the

formation of larger molecules. A 20ml reaction solution (10mg/ml CS, 15mg/ml PAA)

was placed in a concentric cylinder configuration rheometer (AR 2000EX, TA

Instruments). Shear rate was logarithmically ramped from 10/s to 1000/s every hour over

a 4 hour period and the viscosity recorded. In a separate study, the viscosity of an

activated-PAA-CS solution was measured after 24h of reaction (10mg/ml CS, 15mg/ml

Page 113: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

87

PAA) and the viscosity measured in a parallel plate configuration with shear rate

logarithmically ramped from 10/s to 1000/s (AR 2000EX, 750um Gap, 1mL solution).

Control solutions of un-activated-PAA-CS mixed solutions, CS solutions and PAA

solutions were also investigated. Specific viscosity was calculated relative to a

PBS/MES solvent system.

Page 114: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

88

Figure 5.1: Chemical Structure of L-serine and N-acetyl-D-galactosamine controls for development of the fluorescamine assay (www.sigma.com).

Figure 5.2: Chemical structure of the fluorescamine reagent and its reaction with primary amino groups to generate a fluorophor (146).

Page 115: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

89

Figure 5.3: Aldehyde-amine reaction and reductive amination with sodium cyanoborohydride (150)

Figure 5.4: Epoxide-amine reaction

Page 116: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

90

Figure 5.5: Carboxyl-amine reaction mediated by EDC and sulfo-NHS (151)

Page 117: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

91

Figure 5.6: Example of “grafting-to” strategy for building CS brushes from glass substrate

“Grafting To” Strategy

Page 118: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

92

Figure 5.7: “Grafting-to” polymerization strategy using the carboxylic acid to CS amine reaction

Page 119: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

93

Table 5.1: Product Information for Chondroitin Sulfate Samples

Vendor Product Number Tissue Source CS Type Estimated MW

Sigma C4384 Shark Cartilage Primarily CS-6 ~65,000

Sigma C9819 Bovine Trachea 60% CS-4, 40% CS-6 ~22,000

Sigma C6737 Bovine Cartilage Primarily CS-4 ~22,000

Calbiochem 230699 Bovine TracheaMix of CS-4, CS-6, CS-

4,6, CS-2,6~22,000

Page 120: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

94

5.3.1.Primary amine content of CS from various sources

5.3. Results

In order to establish a relationship between fluorescamine fluorescence intensity

and primary amine content of a chondroitin sulfate sample, L-serine was used to establish

a standard curve where L-serine is the amino acid residue of CS and contains one primary

amine (PA) per molecule (Figure 5.8)

The presence of primary amines in CS from various sources and vendors was

investigated using fluorescamine. The concentration of [PA] in sample was calculated

based on the L-serine standard curve (Figure 5.8). Fluorescamine signal versus

concentration of L-serine has both a linear region and a non-linear region at higher

concentrations where saturation of the fluorescence intensity can be seen (Figure 5.8).

Fluorescamine signal is linear to approximately 3500 fluorescence units corresponding to

a serine concentration of 0.625mM. Molarity of L-serine was converted to molecules of

primary amine and the linear region of the fluorescence vs. molecules of primary amine

curve was taken from 1.76x1015 to 5.65x1016 molecules of primary amine. This

corresponded to fluorescence intensities of 138 and 3419 normalized fluorescence units

respectively. The linear relationship of :

𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒𝑠 𝑜𝑓 𝑝𝑟𝑖𝑚𝑎𝑟𝑦 𝑎𝑚𝑖𝑛𝑒 = 6.341𝐸 14 ∗ 𝑓𝑙𝑢𝑜𝑟𝑒𝑠𝑒𝑛𝑐𝑒 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 −

was established with an R2 of 0.9854. This linear relationship was used in order to

determine the primary amine content of different biomoleule samples.

Page 121: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

95

Bovine serum albumin (BSA) and N-acetyl-galactosamine (GalNAc) were

investigated as positive and negative controls for primary amine concentration

respectively. As the concentration of BSA was increased, the normalized fluorescence

for bound fluorescamine detected at excitation/emission of 365/490nm increased linearly

(R2 > 0.99) indicating a clear presence of PAs within the samples (Figure 1). No

correlation was seen between fluorescence from N-acetyl-D-galactosamine (GalNAc)

samples and GalNAc concentration (R2 = 0.08) and fluorescence signal was low

indicating no PAs with the CS saccharide. A strong linear relationship (R2 > 0.99)

between observed fluorescence and CS concentration for all CS samples indicates the

fluorescamine assay as an appropriate means to measure small fractions of PA in the

GAG samples.

The number of PAs per biomolecule was calculated for each concentration of the

biomolecule based on the L-serine standard curve and averaged. The protein BSA was

used as a positive control and had a high primary amine content (19.74 +/- 4.45 primary

amines per molecule) as expected from the many primary amine containing amino acid

residues in the BSA protein backbone. GalNAc had an average of 0.0 primary amines

per molecule indicating that the amide of GalNAc in the CS disaccharide will not

contribute to the CS primary amine signal (Figure 5.2).

CS-6 from shark cartilage (Sigma) had the highest primary amine content of all

chondroitin sulfate samples tested with an average number of primary amines per CS

chain of 10.52 +/- 1.26 (Table 5.3). CS from bovine trachea from both Calbiochem and

Sigma had primary amine contents greater than one primary amine per CS chain with

7.16 +/- 0.95 and 3.08 +/- 0.38 primary amines per CS chains respectively. CS-4 from

Page 122: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

96

bovine cartilage supplied by Sigma had an average number of primary amines per CS

chain of 1.15 +/- 0.7 (Table 5.3). For the application of CS to the synthesis of biomimetic

aggrecan brush structures, the single point, terminal end immobilization of CS is

required.

For this purpose a single primary amine in CS was investigated. CS-4 C6737

from Sigma demonstrated a single primary amine per CS molecule making it an

appropriate candidate CS for use in CS brush synthesis. In addition this CS is from a

mammalian cartilage source, and CS-4 has been widely used in therapeutic settings with

demonstrated anti-inflammatory and anti-oxidant effects (95). CS-4 C6737 from Sigma

will be utilized for all future studies and referred to simply as CS.

5.3.2. Attachment of CS to amine-reactive substrates for the “grafting- to” synthesis

strategy

5.3.2.1. EDAX analysis of CS modified glass surfaces. Semi-quantitative

elemental analysis of CS modified glass slides was conducted using EDAX in order to

monitor CS attachment to functionalized surfaces. Epoxy and aldehyde glass surfaces

were investigated for their ability to bind primary-amine terminated CS chains. Element

peaks from EDAX spectra were identified corresponding to C, O, Na, Mg, Al, Si, S, K,

and Ca and spectra were automatically base-lined corrected for semi-quantitative

analysis. Sulfur was detected at 2.31 keV (Figure 5.11). Sulfur was present in all surfaces

investigated in trace (wt % < 1%) amounts. As seen in Figure 5.11, non-functionalized

Page 123: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

97

glass surfaces have a small peak in the sulfur region (Figure 5.11) while the epoxy coated

slide in 20mg/mL CS has a slightly larger peak in the sulfur region (Figure 5.11).

EDAX spectra were semi-quantitatively analyzed for their sulfur content, where

an increase in sulfur signal with CS modification is an indication of CS attachment.

Sulfur net integrated ratio was measured and normalized to Si net integrated ratio for

non-functionalized and functionalized glass slides coated with 0, 0.05, 2 and 40mg/mL

CS (Figure 5.12).

Epoxy and aldehyde functionalized slides had higher sulfur content after

treatment with 2 mg/mL CS and 40 mg/mL CS with the greatest sulfur content seen on

2mg/ml CS treated slides. Non-functionalized glass slides showed minimal increase in

sulfur content with all CS treatments. Although a general trend in the increase in surface

sulfur content with increased CS treatment concentration was seen, there was no

significant difference (p> 0.05) between slide types or treatment groups (Figure 5.12).

Due to the trace amounts sulfur detected in these samples and the depth of detection of

EDAX (~1-2nm) only qualitative and semi-quantitative analysis can be conducted with

this technique and results may be masked by background sulfur signals. Background

sulfur signals likely arise from trace amounts of sulfur in the glass itself or present as

residue from manufacturing of the slides.

5.3.2.2. Contact angle analysis of CS modified epoxy-functionalized surfaces. To

further investigate CS binding to amine-reactive surfaces via the CS terminal primary

amine functionality, epoxy-functionalized, aldehyde-functionalized and non-

functionalized (control) glass slides were treated with CS solutions of varying

Page 124: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

98

concentrations and CS attachment determined via contact angle measurements. A

decrease in contact angle is expected as CS is deposited onto the substrates since CS is a

charged molecule and will attract water making the surface more hydrophilic. CS

concentrations of 0, 0.125 and 2 mg/mL were investigated. Slides were rinsed

thoroughly after treatment with CS to remove any non-covalently bound CS. On epoxy-

functionalized slides, contact angle without CS was measured at 76.6 +/- 2.8o. With the

addition of CS at 0.125mg/ml contact angle was reduced to 63.4 +/- 9.5o and further

reduced to 38.2 +/- 6.6o with the addition of CS at 2 mg/ml indicating an increase in

hydrophilicity associated with the deposition of charged CS on the epoxy-functionalized

surface (Figure 5.13). Contact angle was significantly different on epoxy slides treated

with CS with a significant decrease in contact angle seen as CS concentration increased

(p < 0.01). Untreated glass slides did not have a significant change in contact angle with

CS treatment. Aldehyde-treated slides had a slight decrease in contact angle with CS

treatement but changes were not statistically significant.

5.3.3. Chondroitin sulfate conjugation to synthetic monomers via the terminal

primary amine.

5.3.3.1. Fluorescamine analysis of CS-monomer conjugation. Three amine

reactive chemistries were utilized to conjugate various monomers to the CS terminal

primary amine. The monomers acrolein, allyl glycidyl ether (AGE) and acrylic acid were

investigated. Varying monomer:CS molar ratios were reacted for 4h and investigated for

PA content of CS before and after conjugation as determined using the fluorescamine

Page 125: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

99

assay. Acrolein was investigated at 17.7:1, 177:1, and 1770:1 molar ratio with CS. A

decrease in CS PA signal is seen with the addition of increasing amounts of monomer

corresponding to an increase in the percent PAs conjugated to monomer as the amount of

acrolein is increased reaching a maximum of 80% at an acrolein:CS molar ratio of 1770:1

(Figure 5.14). A similar trend is seen for CS-AGE solutions and CS-acrylic acid

solutions. The effect is less pronounced for acrylic acid solutions, with a maximum of

26% conjugation seen at a 1000:1 acrylic acid:CS molar ratio. AGE-CS solutions

showed the highest conjugation reaching 99% at a 1000:1 AGE:CS molar ratio (Figure

5.14). Due to the toxic nature of acrolein as well as cyanoborohydride used in order to

stabilize the schiff’s base formed in the aldehyde-amine reaction, the acrolein-amine

reaction was not investigated further, however these studies indicate the ability to

conjugate CS via the terminal primary amine to synthetic monomers using the aldehyde-

amine reaction.

The AGE-CS and acrylic acid (AA)-CS reactions were investigated further for

temporal dependence. AGE Conjugation was dependent both on concentration and

reaction time. For the 1:1 AGE:CS molar ratio, no significant increase in percent

conjugation is seen after 4h however for the higher ratios of 10:1 and 100:1 AGE:CS,

percent conjugation significantly increased after up to 72h of reaction. For 1000:1

reactions, near complete reaction (~99% conjugation) was seen after 24h of reaction

(Figure 5.15). Reaction rate of the CS primary amine to AGE increased with increasing

AGE:CS ratio and 100:1 and 1000:1 AGE:CS ratios followed one-phase association

kinetics (Figure 5.15). Conjugation of CS to acrylic acid was also monitored over time.

No clear relationship was found between reaction time and conjugation (Figure 5.16).

Page 126: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

100

Conjugation significantly increased after 6hr of reaction for all ratios but decreased again

after 24h of reaction. Further reaction time did not increase conjugation significantly

from the 6hr reaction time. These results indicate that the CS-acrylic acid conjugation

cannot be significantly increased with increased reaction time.

5.3.3.2. Qualitative 1H-NMR and ATR-FTIR study of vinyl incorporation in AGE-

CS conjugates. Monomer conjugation to CS was further investigated by 1H-NMR and

ATR-FTIR for the AGE monomer at a 1000:1 AGE:CS molar ratio over time. Spectra

were acquired for both CS and AGE conjugates at 300MHz (Figure 5.17). Several classes

of protons were resolved and assigned on the basis of their chemical shifts for CS (Figure

5.17, Top) and on the basis of previous literature findings for CS-4 (152-153) . A good

correspondence was seen between literature CS-4 spectra and that obtained in these

studies. Chemical shifts were also assigned to protons of the AGE monomer (Figure

5.17).

1H-NMR spectra spectra for AGE-CS samples reacted at 0, 2, 4, 6, and 24h were

investigated for incorporation of the AGE monomer in CS. 1H-NMR spectra for AGE-

CS samples matched well with that of natural CS. Peaks corresponding to CS protons

maintained their position (Figure 5.18) as well as integrated area (Figure 5.19) over the 24h

reaction period indicating stability of the overall CS structure throughout the reaction. In

addition, no epoxide peaks in the 3.0-2.5ppm region were resolved, indicating thorough

purification of un-reacted AGE monomer (Figure 5.18). Inspection of the 7.0-5.0 ppm

region of the AGE-CS spectra reveals the progressive incorporation of peaks

corresponding to the vinyl AGE protons, AGE-1 and AGE-2 (Figure 5.20).

Page 127: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

101

In a qualitative analysis, a relationship was suggested between the AGE vinyl

monomer and CS molecule in the sample. The peak at 2.0 ppm was assigned to the

acetamide methyl protons of the GalNAc sugar (GalNAc-7). Integration of this peak was

set to 100 and corresponds to three protons for the methyl group. For CS with MW

22,000g/mole as assumed in this study, there are approximately 44 dimers (MW per

dimer is 502g/mole). Each dimmer has one methyl group arising from the GalNAc,

therefore, the integrated area of peak GalNAc-7 is attributed to approximately 44 methyl

groups, resulting in an integrated area of approximately 2.27 per methyl. As methyl has 3

protons, this corresponds to an integrated area of 0.75 corresponding to one proton. To

estimate AGE vinyl group incorporation into CS, the integrated area of the AGE-2

proton, attributed to one proton, was ratioed to that of a single proton arising from CS

(GalNAc-7, calculated at 0.75 integrated area for 1 proton) (Figure 5.21). With increasing

reaction time, vinyl:CS ratio increases from 0.51:1, 0.67:1, 0.77:1, 0.84:1 and 1.72:1 for

0, 2, 4, 6, and 24h reaction times respectively. A ratio above 1, as seen for the 24h

reaction, indicates the incorporation of more than one vinyl to each CS chain.

In order to further investigate the incorporation of AGE into CS and possible side

reactions, ATR-FTIR spectra were obtained for AGE:CS 1000:1 molar ratio reactions

over a 24h reaction period. Wavenumber assignments for key peaks corresponding to CS

(154-157) and AGE (158) are presented in Figure 5.22. Of particular interest is the C-O

stretching mode in CS where it has been demonstrated that in the C-O stretching mode,

shifting occurs with changes to the hydroxyl group orientation (156). The most likely

side reaction to occur would be with the primary hydroxyl of the CS GalNAc

disaccharide (159), however this group is not clearly resolved in the 1H-NMR spectra

Page 128: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

102

(combined in the 3.58 and 3.86ppm band), necessitating a secondary means of detection.

ATR-FTIR spectra for AGE-CS at all reaction time points were consistent with that of

natural CS (Figure 5.23). No strong peaks corresponding to the AGE monomer were

detected in the AGE-CS spectra. The FTIR peak centered at 1033cm-1 which arises from

the C-O stretching mode in CS was not significantly shifted in 2, 4, and 6h reaction

samples (ranging from 1031-1029cm-1), however, after 24h of reaction, a larger shift was

seen in this peak, with peak maxima shifting to 1024cm-1. Spectral shifts in this region

would be expected to occur with changes to the hydroxyl group orientations in CS (156),

and therefore, the shift in this spectral region for 24h samples indicates possible CS

hydroxyl group side chain reactions at this time point.

5.3.4. Conjugation of CS to a polymeric backbone for CS-glycopolymer synthesis via the

“grafting-to” strategy

The carboxylic acid-primary amine reaction was utilized in order to investigate

the “grafting-to” method of brush polymer synthesis. Commercially available

poly(acrylic acid) (PAA, 250,000MW) was utilized due to its water soluble,

biocompatible and stable nature. Time, temperature, buffer ionic concentration and

reactant ratio were investigated for their effect on CS conjugation to PAA. The

conjugation of CS to PAA was measured by a change in the fluroescamine signal of CS-

PAA solutions in comparison to control solutions of CS alone. Reactions were also

followed rheometrically.

Page 129: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

103

5.3.4.1. Influence of time on CS-PAA reaction. CS was reacted with activated

PAA and the reaction monitored over short (< 24h) and long (>24h) time frames. Percent

conjugation reached ~77% within 2h and remained fairly constant through 24h of

conjugation (Figure 5.24). A maximum of 79% conjugation was reached at 8h of

conjugation, however, conjugation did not change significantly over the 24hr period

investigated (p > 0.05). Over a longer time frame, conjugation decreased to 39% after

120h. This decrease may reflect some occurrences of non-covalent bonding or

protonation of the CS primary amine over the reaction period. Our results indicate that

the majority of the CS-PAA reaction occurs in the first few hours of the reaction with no

increased conjugation over long reaction times. This may be due to the EDC/sulfo-NHS

activation life which is generally around 4h caused by hydrolysis of the EDC/sulfo-NHS

complex (149).

5.3.4.2. Influence of temperature on CS-PAA reaction§. The effect of reaction

solution temperature on the conjugation of CS to PAA was also investigated.

Conjugation was monitored after 8h of reaction for reactions carried out at 4̊C, 21̊C

(ambient temperature) and 37̊C (Figure 5.25). Conjugation reached 79.90%, 79.67%, and

81.13% respectively. No statistical difference was seen between different reaction

temperatures (n=3, p > 0.05). Ambient temperature was used for all further CS-PAA

conjugation studies.

5.3.4.3. Influence of ionic concentration on CS-PAA reaction§. PAA (15mg/mL,

final concentration) was prepared similarly and conjugated with CS (10mg/mL) at

varying Na+ concentrations (various dilutions of phosphate buffered solution buffered to

pH 7.4) over 8h at room temperature (Figure 5.26). CS was prepared in PBS buffer

Page 130: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

104

solutions of varying concentration and pH adjusted to pH 7.4 with NaOH. Total Na+

concentrations were 0.5669, 0.6962 and 1.8847 for 0.0001X, 1X, and 10X PBS solutions

respectively. Ion concentration may affect the conformation of CS chains in solution and

may play a role in the interaction of CS and PAA which are both charged molecules at

neutral pH. Ionic concentration of the reaction medium had a significant effect on the

conjugation observed via the fluorescamine assay (n = 3, p < 0.05 as determined by one-

way ANOVA and Tukey’s multiple comparison post hoc analysis for comparisons

between 0.696M Na+ and 1.885M Na+ and, between 0.567M Na+ and 0.696M Na+) .

Highest average conjugation (77%) was seen for conjugation in a 0.696M Na+ which

corresponds to a 1X PBS buffer solution. Average conjugation was decreased for the

0.567 M Na+ solutions (60%) as well as for 1.885M Na+ (56%). 1X PBS buffer solution

was used for all further CS-PAA conjugation studies.

5.3.4.4. Conjugation of CS to PAA at varying CS:PAA molar ratio§. Varying CS

to PAA molar ratio solutions were prepared and conjugated for 8h. Ratios of 20:1, 10:1

and 1:1 CS:PAA were investigated in addition to the original 6.8:1 CS: PAA molar ratio.

Conjugation was lower for the 20:1 and 10:1 molar ratios at 47% and 53% average

conjugation respectively. The 1:1 CS:PAA molar ratio formulation had the highest

conjugation reaching 97%. All conjugations were statistically different as determined by

a one-way ANOVA and Tukey’s multiple comparison post-test (n = 3, p < 0.001) except

for the comparison between 20:1 and 10:1 CS:PAA molar ratios.

Based on the number of carboxylic acid sites on a single PAA chain and the %

conjugation of primary amines, the theoretical number of CS chains attached to each

PAA chain was calculated. For these theoretical calculations it was assumed that each

Page 131: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

105

CS chain had 1 primary amine and that a decrease in primary amine content as detected

by the fluorescamine assay was a result of the CS primary amine conjugating with a

carboxylic acid of PAA and remaining bound. The maximum number of active sites

(activated acrylic acid (AA) monomers) on the 250kDa PAA is calculated to be 3470

given an acrylic acid MW of 72Da. The number of CS bound acrylic acid sites to un-

bound acrylic acid sites was calculated as follows from the CS:PAA ratio and resulting %

conjugation:

3470 ∗ [𝑃𝐴𝐴]

[𝐶𝑆] ∗ % 𝐶𝑜𝑛𝑗𝑢𝑔𝑎𝑡𝑖𝑜𝑛100

And the number of CS chains attached to each PAA chain can be estimated as follows:

[𝐶𝑆] ∗ % 𝐶𝑜𝑛𝑗𝑢𝑔𝑎𝑡𝑖𝑜𝑛100

[𝑃𝐴𝐴]

The number of CS chains attached to each PAA chain was calculated for each CS:PAA

ratio and are reported in Table 5.4. The number of CS chains attached to each PAA

polymer ranges from 1 in the 1:1 CS:PAA formulation to 9 in the 20:1 CS:PAA

formulation. For 6.8:1, 7.5:1 and 10:1 formulations theoretical calculations estimate 5

CS chains per PAA chain. All investigated formulations suggest sparsely grafted

structure.

§These studies were conducted as a part of the Master’s Thesis entitled, “Biomimetic Aggrecan Based on a Poly Acrylic Acid 9PAA) Core and Chondroitin Sulfate (CS) Bristles” (2010, Drexel Univeristy) by Nandita Ganesh

Page 132: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

106

5.3.4.5. Rheological assessment of the CS-PAA reaction. The rheological

properties of CS-PAA solutions were investigated during the reaction period in order to

monitor the formation of CS-PAA reaction products. A rheometer with a concentric

cylinders sample set-up was utilized in order to measure the viscosity of CS-PAA

reaction solutions periodically over a 4h reaction period. Solution viscosity increased

from a specific viscosity of 1.194 to 1.229 over the 4h period and followed pseudo-first

order association kinetics (R2 = 0.99) (Figure 5.28). Specific viscosity was also

investigated for CS-PAA reacted solutions after a 24h reaction period (Figure 5.29). CS-

PAA reacted solution had a specific viscosity of approximately 1.35 after 24h of reaction

which was higher than that of un-reacted CS-PAA solutions (a solution of CS and un-

activated PAA simply mixed immediately before rheological testing) which had a

specific viscosity of approximately 0.82. The specific viscosity of PAA and CS solutions

were also investigated at the concentrations used in the reaction and found to be 0.59 and

0.51 respectively. All solutions maintained a steady-state specific viscosity with respect

to shear rate over a shear rate range from 50/s to 200/s.

Page 133: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

107

Fluorescamine Standard Curve for L-Serine

0.000 0.002 0.004 0.0060

2000

4000

6000

8000

10000

Concentration of L-Serine (M)

Fluo

resc

ence

(Nor

mal

ized

Flu

ores

cenc

e U

nits

)

Fluorescamine Standard Curve for Primary Amines

0 2.0×101 6 4.0×101 6 6.0×101 60

1000

2000

3000

4000

y = 6.341E-14xR2 = 0.9854

Molecules of Primary Amine

Fluo

resc

ence

(Nor

mal

ized

Flu

ores

cenc

e U

nits

)

Figure 5.8: L-serine-fluorescamine standard curve and linear region of fluorescence v.s. sample primary amine content based on the L-serine

standard curve.

Page 134: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

108

Fluorescamine Analysis of Primary AminesIn Control Substances

0.0000 0.0002 0.0004 0.00060

2000

4000

6000L-SerineBovine Serum AlbuminGalNAc

Concentration [M]

Fluo

resc

ence

(Nor

mal

ized

Flu

ores

cenc

e U

nits

)

Figure 5.9: Fluorescamine analysis of control samples bovine serum albumin and GalNAc for validation of the fluorescamine assay.

Fluorescamine Analysis of Chondroitin SulfatePrimary Amine Content

0.0000 0.0001 0.0002 0.0003 0.0004 0.00050

1000

2000

3000

4000 Sigma CS-6Shark CartilageSigma CS-4/CS-6Bovine TracheaSigma CS-4Bovine CartilageEMD CS MixBovine Cartilage

Concentration [M]

Fluo

resc

ence

(Nor

mal

ized

Flu

ores

cenc

e U

nits

)

Figure 5.10: Fluorescamine analysis of chondroitin samples from various sources

Page 135: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

109

Figure 5.11: Representative EDAX spectra (a) non-functionalized glass slide in 0mg/mL CS and (b) epoxy functionalized glass slides in

20mg/mL CS solutions. Spectra were collected at 6keV auto base-lined and semi-quantitatively analyzed for elemental analysis. Elements C, O, Na, Mg, Al, Si, S, K and Ca were detected with

special attention to Si and S for analysis. Trace amounts of S are detected at 2.31 keV.

Page 136: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

110

CS Attachment to FunctionalizedGlass Slides: EDAX Analysis

Glass S

lide

Epoxy Slid

e

Aldehyd

e Slid

e0.0

0.5

1.0

1.5

2.00mg/ml CS0.05mg/ml CS

40mg/ml CS2mg/ml CS

Nor

mal

ized

S/S

i Rat

io

Figure 5.12: Surface sulfur content on functionalized glass surfaces after treatment with CS EDAX analysis of functionalized glass surfaces modified with varying concentrations of CS (6 keV, spot size 6, working distance 12mm, analysis area 121µm x 95µm). Spectra were semi-

quantified and the ratio of S to Si net integrated area was determined and normalized to samples modified with buffer solutions alone. (n = 3, 2-way ANOVA with Bonferroni Post-tests, no

statistical differences observed)

Page 137: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

111

Chondroitin Sulfate FunctionalizedGlass Slides

Un-Treated

Epoxy-Trea

ted

Aldehyd

e-Trea

ted0

20

40

60

80

1000 mg/ml CS0.125 mg/ml CS2 mg/ml CS

Glass Slide Treatment

Con

tact

Ang

le (°)

0 mg/ml CS vs 0.125 mg/ml CS

0 mg/ml CS vs 2 mg/ml CS

0.125 mg/ml CS vs 2 mg/ml CS

Un-Treated Glass Slide ns ns ns Epoxy Treated Glass Slide ** *** *** Aldehyde Treated Glass Slide ns ns ns ** P < 0.01; ***P < 0.001

Figure 5.13: Contact angle measurements on CS functionalized glass slides (n>4, 2-way ANOVA with Bonferroni Post-tests)

Page 138: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

112

Figure 5.14: Conjugation of CS with amine reactive monomers Change in primary amine content, as detected by the fluorescamine assay, with the addition of amine reactive monomers acrolein, allyl glycidyl ether, and acrylic acid. % PA conjugated to

monomer increases with increasing monomer:CS molar ratio. A higher degree of conjugation is realized for acrolein and AGE than for acrylic acid.

Page 139: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

113

CS-AGE Conjugation Over Time

1 10 100

0

50

1001:110:1100:11000:1

Time (hrs)

% C

onju

gatio

n

One-phase association: Y=Plateau*(1-exp(-K*x)) AGE:CS Molar Ratio 1:1 10:1 100:1 1000:1 Plateau (% Conjugation) 12.48 58.8 81 96.26 K (1/hr) 0.124 0.01874 0.2031 1.32 Tau (hr) 8.068 53.36 4.924 0.7575 95% Confidence Intervals Plateau (% Conjugation) 7.126 to 17.83 32.35 to 85.25 76.59 to 85.41 93.65 to 98.88 K (1/hr) 0.0 to 0.3901 0.001883 to 0.03560 0.1547 to 0.2515 1.155 to 1.485 Tau (hr) 2.563 to +infinity 28.09 to 531.1 3.977 to 6.464 0.6735 to 0.8655 R square 0.3991 0.8143 0.9304 0.9357

Figure 5.15: CS-AGE conjugation over time for varying AA:CS molar ratios with one-phase association exponential curve fit (n=3, 2-way ANOVA with Bonferroni Post-tests)

1:1 0 2 4 24 48 72 96 100:1 0 2 4 24 48 72 960 02 ns 2 ns4 ** ns 4 *** *24 *** *** ns 24 *** *** ***48 ** ns ns * 48 *** *** *** ns72 *** ** ns ns ns 72 *** *** *** ** ns96 *** ** ns ns ns ns 96 *** *** *** *** ns ns

10:1 0 2 4 24 48 72 96 1000:1 0 2 4 24 48 72 960 02 ns 2 ***4 * ns 4 *** ns24 ns ns ns 24 *** * ns48 *** *** *** *** 4872 *** *** *** *** ns 7296 *** *** *** *** * ns 96

P value P > 0.05 P < 0.05 P<0.01 P<0.001

Summary ns * ** ***

Page 140: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

114

CS-Acrylic Acid Conjugation Over Time

20 40 60-5

0

5

10

15

201:110:1100:1

Time (hrs)

% C

onju

gatio

n

1:1 2 6 24 48 10:1 2 6 24 48 100:1 2 6 24 48 2 2 2 6 *** 6 *** 6 *** 24 ns *** 24 ns *** 24 ns ** 48 * ns * 48 ** ns * 48 *** ns ***

One-phase association: Y=Plateau*(1-exp(-K*x)) 1:1 10:1 100:1 Plateau (% Conjugation) 2.731 4.936 11.71 K (1/hr) 0.4182 0.4229 0.4512 Tau (hr) 2.391 2.365 2.216 95% Confidence Intervals Plateau (% Conjugation) -0.1446 to 5.606 2.048 to 7.825 8.774 to 14.65 K (1/hr) 0.0 to 2.302 0.0 to 1.485 0.0 to 0.9493

Tau (hr) 0.4344 to +infinity

0.6732 to +infinity

1.053 to +infinity

R square 0.08588 0.1815 0.3374

Figure 5.16: CS-Acrylic Acid conjugation over time for varying AA:CS molar ratios with one-phase association exponential curve fit (n=3, 2-way ANOVA with Bonferroni Post-test)

P<0.001

ns * ** ***

P value

Summary

P > 0.05 P < 0.05 P<0.01

Page 141: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

115

Figure 5.17: 1H-NMR spectra of (Top) CS and (bottom)AGE solutions in D2O with spectral assignments. See Appendix 2 for alternate view of CS.

Page 142: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

116

Figu

re 5

.18:

CS

regi

on o

f 1 H-N

MR

spec

tra o

f AG

E-CS

con

juga

te (1

000:

1 A

GE:

CS

mol

ar ra

tio) w

ith

incr

easi

ng re

actio

n tim

e

Page 143: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

117

Figu

re 5

.19:

Ana

lysi

s of 1 H

-NM

R in

tegr

ated

are

a (n

orm

aliz

ed to

Gal

NA

c M

ethy

l pro

ton)

of A

GE-

CS

sam

ples

ove

r 24h

Rea

ctio

n tim

e

Page 144: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

118

Figu

re 5

.20:

1 H-N

MR

spec

tra o

f vin

yl-p

roto

n re

gion

of A

GE-

CS

(100

0:1

mol

ar ra

tio) s

ampl

es sy

nthe

size

d ov

er

24h

with

AG

E vi

nyl p

roto

n sp

ectra

l ass

ignm

ents

Page 145: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

119

AGE:CS 1000:1 ReactionOver Time

0.0 0.5 1.0 1.5 2.0

CS

0h

2h

4h

6h

24h 1.72:10.84:1

0.77:1

0.67:10.51:1

0.07:1

Vinyl Group:CS Chain Ratio(As estimated by 1H-NMR analysis)

Rea

ctio

n Ti

me

(h)

Figure 5.21: Qualitative analysis of vinyl group incorporation into CS based on 1H-NMR spectral analysis of AGE-CS conjugation over time. Vinyl proton incorporation was based on the

integrated area of proton 2 of AGE compared the integrated area of 1 proton, estimated from the methyl peak at 1.9ppm.

Page 146: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

120

Figu

re 5

.22:

ATR

-FTI

R sp

ectra

and

wav

enum

ber a

ssig

nmen

ts fo

r cho

ndro

itin

sulfa

te a

nd A

GE

Page 147: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

121

Figu

re 5

.23:

ATR

-FTI

R sp

ectra

for A

GE-

CS

conj

ugat

es re

acte

d ov

er 2

4h.

Hig

hlig

hted

is th

e C-

O st

retc

hing

mod

e as

soci

ated

with

the

CS

disa

ccha

ride

back

bone

. Th

e C

-O st

retc

hing

mod

e is

sens

itive

to m

odifi

catio

n of

the

CS

hydr

oxyl

Page 148: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

122

CS-PAA ConjugationOver a Short Time Frame

0 10 20 300

20

40

60

80

100

Time (h)

% C

onju

gatio

n

CS-PAA ConjugationOver a Long Time Frame

0 50 100 1500

20

40

60

80

100

Time (hr)

% C

onju

gatio

n

Figure 5.24: CS-PAA conjugation over time (Top) Short term conjugation of CS-PAA with linear regression (PAA Concentration 16.5mg/ml

(0.066mM), CS concentration 10mg/ml (0.45mM)) (Bottom) Long term conjugation of CS-PAA with non-linear one-phase exponential decay (PAA

Concentration 15mg/ml (0.06) CS concentration 10mg/ml (0.45mM)); Room temperature reaction in PBS.

Page 149: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

123

Figure 5.25: CS-PAA conjugation with varying temperature (8hr reaction) PAA Concentration 16.5mg/ml (0.066mM), CS concentration 10mg/ml§

§Data presented previously in Master’s Thesis, “Biomimetic Aggrecan Based on a Polyacrylic Acid (PAA) Core and Chondroitin Sulfate (CS) Bristles” (2010, Drexel University) by Nandita Ganesh (160).

CS-PAA Conjugation with VaryingReaction Temperature

4 C 21 C

37 C

0

20

40

60

80

100

Temperature (C°)

% C

onju

gatio

n

Page 150: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

124

CS-PAA Conjugation with VaryingNa+ Concentration

0.566

9

0.696

2

1.884

7 0

20

40

60

80

100 * **

* p<0.05; **p<0.01Na+ Concentration (M)

% C

onju

gatio

n

Figure 5.26: CS-PAA conjugation with varying Na+ Concentration (8hr Reaction) PAA Concentration 16.5mg/ml (0.066mM), CS concentration 10mg/ml (0.45mM) §

§Data presented previously in Master’s Thesis, “Biomimetic Aggrecan Based on a Polyacrylic Acid (PAA) Core and Chondroitin Sulfate (CS) Bristles” (2010, Drexel University) by Nandita Ganesh (160).

Page 151: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

125

CS-PAA Conjugation with VaryingCS:PAA Ratio

20:1

10:1

6.8:1 1:1

0

20

40

60

80

100

CS:PAA Molar Ratio

% C

onju

gatio

n

Figure 5.27: CS-PAA conjugation with varying CS:PAA molar ratio (8hr Reaction) Reaction in 1X PBS buffer at 21̊C§

§Data presented previously in Master’s Thesis, “Biomimetic Aggrecan Based on a Polyacrylic Acid (PAA) Core and Chondroitin Sulfate (CS) Bristles” (2010, Drexel University) by Nandita Ganesh (160).

Page 152: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

126

Time Course Rheometry of CS-PAA Reaction10mg/mL CS, 15mg/ml PAA, Shear Rate 158.5/s

Concentric Cylinders Configuration

0 1 2 3 41.15

1.20

1.25

Time (hrs)

Spec

ific

Vis

cosi

ty

Figure 5.28: Rheological investigation of CS-PAA reaction solutions (7.5:1 CS:PAA Ratio) with one-phase association exponential curve fitting

Page 153: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

127

Viscosity of CS-PAA Solution with Reaction10mg/mL CS, 15mg/ml PAA, 24hr Reaction,

Parallel Plate Configuration, 750um Gap

50 100 150 2000.0

0.5

1.0

1.5

Chondroitin Sulfate

Poly Acrylic Acid

CS-PAAMixed Solution

CS-PAAReacted Solution

Shear Rate (1/s)

Spec

ific

Visc

osity

Figure 5.29: Specific viscosity over shear rate for a CS-PAA reacted solution (7.5:1 CS:PAA Ratio) (no curve fitting)

Page 154: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

128

Table 5.2: Estimated Number of Primary Amines per Molecule for Control Samples Based on the L-serine Fluorescamine Standard Curve

Biomolecule Control Molecular Weight

Average # PA/ Molecule

Standard Deviation

Serine Amino Acid 105.09 1.17 0.12 Bovine Serum Albumin Protein 66,400 23.41 0.60 GalNAc Saccharide of CS 221.2 0.00 0.05

Table 5.3: Estimated Number of Primary Amines per Molecule for Chondroitin Sulfate Samples Based on the L-serine Fluorescamine Standard Curve

Chondroitin Sulfate Tissue Source CS Type Average # PA/CS

Chain Standard Deviation

Sigma C4384 Shark Cartilage Primarily CS-6 10.52 1.26

Sigma C9819 Bovine Trachea

60% CS-4, 40% CS-6 3.08 0.38

Sigma C6737 Bovine Cartilage Primarily CS-4 1.15 0.07

Calbiochem 230699

Bovine Trachea

Mix of CS-4, CS-6, CS-4,6, CS-2,6

7.16 0.95

Page 155: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

129

Table 5.4: Theoretical estimation of number of CS chains attached per PAA chain after 8hr reaction; §Samples reacted for 24h.

CS:PAA Molar Ratio

Average % Conjugation

# AA sites/ # CS bound AA Sites

~ # CS Chains attached per PAA Chain

1:1 97 3577 1 6.8:1 77 663 5 §7.5:1 64 723 5

10:1 54 643 5 20:1 47 369 9

Page 156: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

130

Commercially available CS with a terminal primary amine was investigated for its

utility in the fabrication of biomimetic aggrecan bottle-brush structures. A major hurdle

in the synthesis of polymers with carbohydrate pendant groups (glycopolymer) is the

synthesis and preparation of the reactive carbohydrate or glycomonomer (123-126). The

glycomonomer must either be reactive to polymeric substrates or be polymeriziable into a

polymeric backbone (118). To date, chondroitin sulfate has not been investigated as a

glycol-monomer for glycopolymer bottle brush synthesis. In this report, we have

identified a terminal primary amine in commercially available CS and determined the

ability to covalently link CS to synthetic materials via this this terminal-end functionality.

Three amine-reactive chemistries were investigated for reaction with the terminal primary

amine of CS and the attachment of polymerizable vinyl groups. These chemistries were

further investigated for the immobilization of CS onto amine-reactive surfaces and a

polymeric backbone.

5.4. Discussion

5.4.1. CS Terminal Primary Amine

CS is primarily obtained by isolation from tissue sources. It is inhomogeneous

across tissue sources, as well as across isolation techniques and a minimal amount of data

is provided on the physical and chemical properties of the biomolecule when purchased

(95, 139). For our studies, we required that a primary amine be available on the terminal

end of CS. Although it is likely a serine residue is present based on reported cleavage

techniques to remove CS from the protein backbone, exact isolation technique and

Page 157: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

131

protein residue content is not reported for commercially-available CS (89, 129, 139).

Variations in the isolation procedure may result in the introduction of additional PAs

throughout the CS chains as well as cleavage of the linkage region from the CS main

chain (89). Although these small variations are not generally disruptive to the overall

structure and function of the CS molecule and therefore not reported, they are critical for

our investigation.

CS was investigated from two vendors, Calbiochem and Sigma and three tissue

sources, bovine trachea, bovine cartilage and shark cartilage. The fluorescamine assay,

which is sensitive to primary amines at the picomolar level was used to detect PAs in the

CS macromolecule (145). CS-6 from shark cartilage which is a larger type of CS (~65K

MW) had the highest number of primary amines per chain. Due to the larger size of this

CS as well as its high primary amine content, with CS was not pursued for the synthesis

of biomimetic aggrecan. However, with further purification, and mild isolation

techniques, the utilization of CS-6 in biomimetic aggrecan may allow for a diverse family

of biomimetic aggrecans to be synthesized with varying CS chemistry and molecular

weight. The primary amine content of CS-4 and mixes of CS from bovine trachea varied

between vendors with Calbiochem CS-4 having approximately 7 PAs compared to

approximately 3 for Sigma CS-4 (Table 5.3). The difference in the number of PAs is

indicative of the inhomogeneity of the CS isolation techniques. Calbiochem CS which is

a less pure mix of different CS isomers may have more PAs due to protein impurities

within the CS isolate. Also possible is the deacetylation of the GalNAc of CS. The

amide bond in GalNAc is subject to hydrolysis in alkaline solutions, and processing

conditions may cause breakage of this bond exposing PAs throughout the CS backbone.

Page 158: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

132

This is particularly disruptive to the synthesis of CS bottle brush structures since reaction

chemistries must be limited to a single point on the molecule, preferably on the terminal

end in order to develop a well controlled, organized structure. CS-4 from bovine

cartilage, supplied by Sigma (C6737) was determined to have ~1.15 primary amines per

CS chain. This is indicative of a single primary amine per CS chain, and based on

isolation techniques, the single primary amine is likely at the terminal end of CS and

arises from a serine residue. Sigma C6737 which is the sodium salt of chondroitin sulfate

is marketed as a standard for cetylpyridinium chloride (CPC) titration and is therefore

likely more purified than other marketed CS. This purity may have reduced protein

contamination more apparent in other CS products.

In the calculation of CS primary amine content, polydisperse molecular weight

distributions may affect the calculation of the ratio of PAs to CS molecules. Based on

reported literature values, a CS MW of 22,000 was assumed for CS-4 and mixed

populations of CS (65, 139) and 65, 000 for CS-6 however it is expected that there is a

MW distribution within each CS sample and between vendors’ isolation techniques,

tissue source and age may result in varying CS MW. If MW is greatly different between

products the calculation of the number of PAs per CS molecule may be skewed

contributing to the discrepancy between calculated values. Based on the fluorescamine

analysis of CS primary amine content conducted here, Sigma CS-4 from bovine cartilage,

supplied by Sigma was chosen for further investigation.

5.4.2. Surface grafting of CS via the terminal primary amine

Page 159: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

133

The utility of the CS terminal primary amine in immobilizing CS was investigated

using model surfaces. Model surfaces functionalized with amine reactive groups were

used to investigate the attachment of CS to functionalities that may be incorporated into

polymeric backbones. Epoxy and aldehyde functionalized surfaces were treated with CS

solutions of varying concentration and investigated using EDAX as well as contact angle.

Reaction conditions were such that amine reactions with the surface would be more

favorable than reaction with other functional groups of CS such as carboxylic acids and

hydroxyls. A pH of 9.4 was used in CS attachment to both surfaces, this is particularly

important in CS attachment to the epoxide surfaces where epoxides are reactive to both

hydroxyls and sulfhydryls when pH is high (pH ~ 11) and low (neutral pH) respectively

(149).

Semi-quantitative EDAX analysis showed an overall increase in surface sulfur

content on functionalized slides attributable to the attachment of CS to the surface (Figure

5.12). Surfaces were thoroughly rinsed before analysis to remove any adsorbed CS

chains leaving only covalently bound CS. Non-functionalized glass slides were

investigated as a control and they had no discernable increase in sulfur content with CS

treatment indicating minimal CS signal due to adsorption. Sulfur signal on epoxy and

aldehyde functionalized slides varied with CS concentration, however a relationship

between CS concentration and attachment could not be established since sulfur was only

present in trace amounts and therefore difficult to accurately quantify. Although sulfur

may be present on the surface of the slides at a higher wt% the depth of analysis for

EDAX is approximately 1 -2µm, therefore signal from the glass may overpower surface

signal from sulfur. Depth of analysis of the EDAX for sulfur given 6keV of energy was

Page 160: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

134

predicted to be approximately 0.7µm which may account for the sulfur signal seen on

glass surfaces treated with 0mg/mL CS where sulfur may be present as an impurity in the

glass. The sulfur within the glass slides may have reduced the sensitivity of this method

for detecting change in surface sulfur content with CS treatment however the increase in

sulfur content on functionalized slides suggests CS attachment. A more sensitive method

that may be utilized for the determination of CS surface attachment is x-ray photoelectron

microscopy which can detect surface chemistries with a depth of analysis of only 1-

10nm.

To further investigate CS attachment to amine reactive surfaces, epoxy-and

aldehyde- functionalized surfaces were treated with varying CS concentration of CS and

surface hydrophilicity was measured using a contact angle meter. Due to the charged

nature of CS it is expected that surface hydrophilicity would increase (i.e. contact angle

would decrease) as CS is deposited onto the surface (Figure 5.13). A decreasing contact

angle was seen for epoxy-functionalized surfaces with increased CS concentration. This

indicates a strong relationship between CS concentration and attachment of CS to the

surface suggesting covalent binding of CS. No such relationship was seen for non-

functionalized glass surfaces which is indicative of simply adsorbed CS. Although a

similar trend was seen for aldehyde-functionalized slides, there was no statistical

difference between aldehyde functionalized slides indicating low levels of CS attachment

on these surfaces. These surface studies provide evidence that CS can be immobilized

onto amine reactive surfaces. Such immobilization can be transferred to synthetic

monomers or polymer chains for the synthesis of CS bottle brush polymers via the

“grafting-through” or “grafting-to” methods.

Page 161: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

135

5.4.3. Reactivity of the CS terminal primary amine and attachment of a polymerizable

monomer

Three different amine reactive chemistries were investigated for the

functionalization of CS at its terminal primary amine with a polymerizable monomer.

Aldehyde, epoxide and carboxylic acid reactive groups were investigated by the

conjugation of CS to acrolien, allyl glycidyl ether and acrylic acid monomers

respectively. All three monomers have one amine reactive functionality as well as a

vinyl leaving group. The vinyl leaving group may be used subsequently in order to

polymerize CS into brush structures via the “grafting-through” strategy (113-114, 118).

The attachment of monomer to the CS primary amine was monitored using the

fluorescamine assay, where a drop in fluorescence signal from CS samples would be

expected as monomer binds to the CS PA making it inaccessible to the fluorescamine

reagent. As the molar ratios of monomer:CS was increased, an increase in the percent of

monomer conjugation was also seen (Figure 5.14). In all cases, conjugation was

modulated by monomer:CS molar ratio with the maximum conjugation seen at a molar

excess of monomer at 1000:1. Higher molar ratios could not be achieved due to

insolubility of the monomers at higher concentrations. A large molar excess is likely

required due to the small concentration of PAs available in comparison to the CS

disaccharide concentration.

Acrolein binding to CS via the aldehyde-amine reaction reached 80% at high

monomer:CS ratios and demonstrates the utility of this binding strategy in modifying and

Page 162: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

136

immobilizing CS however, the toxic nature of acrolein and the coupling reagent sodium

cyanaborohydride made this reaction strategy undesirable for further studies in

comparison to AGE which also had high conjugation but does not require additional

coupling reagents. In addition, the low stability of aldehyde functional groups and the

complexities associated with the synthesis of water soluble polymers with pendant

aldehyde groups (161), the aldehyde reaction is less desirable for initial studies into the

synthesis of biomimetic aggrecan.

The carboxylic-acid amine reaction was investigated for its temporal dependence.

Acrylic Acid:CS conjugation only reached a maximum of 15% at a 100:1 molar ratio.

This may be due in part to the several steps required to activate acrylic acid for reaction

with PAs. An EDC/sulfo-NHS coupling reaction was utilized where incomplete

activation of the carboxylic acids in acrylic acid may have led to fewer active monomers

available for conjugation. Carbodiimide coupling of carboxylic acids and amines is often

used in order to promote the reaction of the two functional groups in order to form amide

bonds (162). Amide bonds are very hydrolytically stable having a half life of ca. 600

years in neutral solutions at 25̊C which makes the amide linkage very attractive for

bioconjugation applications (163). In the absence of a coupling agent, carboxylic acids

and amines do not form amide bonds however interactions between the two species can

result in the formation of protonated amines. Protonated amines are not detectable using

the fluorescamine assay. Their formation in the AA:CS reaction medium can result in a

temporary decrease in the fluorescamine signal, transiently increasing the conjugation

detected. This may account for the fluctuations in the conjugation seen over time. No

clear time dependence was seen for the AA-CS reaction. Both linear and one phase

Page 163: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

137

exponential fits were poor fits (Figure 5.16). No significant increase in conjugation was

seen beyond 6h of reaction. This is in good agreement with the window of reactivity of

the EDC/sulfo-NHS activation of carboxylic acids which generally ranges from 2-4h

(149, 162). Modifications of the EDC/sulfo-NHS activation procedure may improve

conjugation of the CS primary amine to carboxyl functional groups. A two-step

activation procedure was utilized in this study in order to avoid activation of carboxylic

acids in CS. Quenching of excess EDC with 2-mercaptoethanol may have reduced

conjugation between CS and acrylic acid. Filtration of excess EDC instead of quenching

may improve the reaction in addition to optimization of EDC/sulfo-NHS reagent

concentrations.

Due to the comparatively low conjugation between acrylic acid and CS seen in

this study, the addition of a vinyl monomer to CS via acrylic acid conjugation is not

optimal. These studies however do demonstrate the feasibility of reacting the CS primary

amine with carboxylic acid functionalities using carbodiimide coupling and this reaction

was investigated further for the conjugation of CS to poly(acrylic acid) in the “grafting-

to” strategy of polymer synthesis (Figure 5.7).

The conjugation of AGE to CS was investigated further and the reaction kinetics

determined for varying AGE:CS molar ratios. At the molar ratio of 1:1 AGE:CS, no

clear time dependence was observed (Figure 5.15). Data could not be accurately fit to

exponential or linear regressions (R2 = 0.4 and 0.2 respectively). At low molar ratios,

reaction between CS and AGE are minimal. This may be due to the large amount of CS

present in the solution with respect to the terminal primary amine reactive group and the

AGE monomer. For higher ratios the CS-AGE reactions progressed with time and the

Page 164: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

138

rate constants increased with increasing AGE:CS molar ratio. The 10:1 AGE:CS molar

ratio reaction progressed with time but did not reach a plateau. Based on an exponential

fit of moderate accuracy (R2=0.8) this reaction is predicted to reach ~58.8% conjugation

with a rate constant of 0.019h-1 however the 95% confidence interval for the rate constant

ranged from 0.002h-1 to 0.036h-1. The extrapolated time constant had a 95% confidence

interval between 28h and 53h and therefore may be beyond the scope of the time points

investigated and thus the rate constant cannot be accurately determined from the time

range investigated in this study. These studies were conducted at a CS concentration of

25mg/ml; increasing the overall concentration of both CS and AGE monomer may aid in

increasing reaction kinetics by reducing interpolymeric repulsion and increasing the coli-

coil overlap time (164). Reaction kinetics at higher concentration however cannot be

monitored using the fluorescamine assay due to the limited linear region of the

fluorescamine signal to amine concentration standard curve.

At the 100:1 AGE:CS molar ratio, a maximum of 89% conjugation was reached

after 96h of reaction. Based on the one-phase association exponential fit (R2 = 0.93), a

plateau of 81% conjugation is predicted for this reaction with a rate constant of 0.2h-1

corresponding to a time constant of 4.9h (95% confidence interval between 4h and 6h).

This indicates that a maximum conjugation was reached within the time window

investigated. These studies at the 10:1 and 100:1 AGE:CS molar ratios demonstrate the

longevity of the CS-AGE reaction. At the 1000:1 ratio the reaction proceeds quickly at a

rate constant of 1.3h-1 reaching a conjugation of over 98% after just 24h. In the synthesis

of vinyl-ended CS chains the highest conjugation is desired in order to synthesis a

majority population of vinyl modified chains, therefore the 1000:1 ratio is most

Page 165: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

139

appropriate for this purpose. Studies at the lower molar ratios however demonstrate the

ability to utilize the epoxide-amine reaction in synthesis techniques that require long

reaction times such as the “grafting-to” method of brush synthesis. In surface brush

synthesis for example, three regimes of brush synthesis kinetics have been observed

(165-167). In the first regime, polymer attachment occurs quickly but results in sparsely

grafted polymers that take on a flattened mushroom like arrangement. In the second,

regime attachment occurs considerably more slowly, on the order of several hours to

days. The second regime of attachment is necessary however in order to generate

densely grafted brushes with extended chain conformations. In a third regime,

attachment rate increases again until saturation occurs. It is therefore necessary, when

considering reactive chemistry options to consider the rate and longevity of the reaction.

The epoxide-amine reaction affords long reaction times and can therefore be useful in the

fabrication of dense structures, especially in the “grafting-to” method of synthesis.

We further investigated the incorporation of AGE monomer into CS (1000:1 AGE:CS

molar ratio) over a 24h reaction period using 1H-NMR (Figure 5.18, Figure 5.20) and

ATR-FTIR (Figure 5.23). Both 1H-NMR and ATR-FTIR analysis of the AGE-CS

macromolecule suggested stability of the CS chemical structure over the reaction period.

Monitoring of vinyl group incorporation into CS suggested progressive incorporation of

the AGE vinyl group onto CS with up to a 0.84:1 Vinyl:CS ratio seen after 6h. This ratio

suggests that approximately 84% of CS molecules have a single vinyl group incorporated

at this reaction time point. In conjunction with fluorescamine data demonstrating

reaction of AGE with the CS primary amine, and ATR-FTIR data indicating minimal side

reactions with CS hydroxyls, this data indicates, selective reaction of AGE with the Cs

Page 166: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

140

terminal primary amine and incorporation of single vinyl groups at the terminal end of

CS after 6h of reaction. If a greater number of side reactions were occurring, the amount

of vinyl proton incorporation into the CS 1H-NMR spectra would be expected to be

considerably higher. An example of this can be seen in the 24h reaction of AGE-CS

(1000:1 AGE:CS molar ratio). After 24h a ratio of 1.72:1 Vinyl:CS was determined

indicating, that after 24h of reaction more than one vinyl group is likely incorporated into

each CS molecule. This would indicate possible side reactions of the AGE monomer

with reactive groups other than the single terminal primary amine. This is further

supported by the relatively large shift seen in the C-O stretching mode peak of the 24h

AGE-CS macromolecule, indicating possible side reactions on the CS hydroxyls. In the

1000:1 AGE:CS molar ratio reaction, the primary amine reaction rate was determined to

be 0.75h (Figure 5.15) indicating that majority of the primary amine reaction should

occur within the first 3h (~95% of the reaction). The plateau % conjugation for this

reaction was high, at 96%, therefore, after the initial hours of reaction, very little primary

amine is available, and excess epoxide may be more likely to react with originally less

reactive species. This phenomenon can be seen in the 24h reaction.

The estimation of vinyl incorporation into the CS macromolecules is semi-

quantitative and sensitive to CS MW, as well as noise in the 1H-NMR signal, signal

baselining and choice of integration region. However, in conjunction with fluorescamine

studies and FTIR analysis, these results provides promising evidence that the AGE

monomer is conjugating with CS specifically at the CS terminal primary amine. CS

spectra after conjugation also matched well with natural CS spectra, indicating no major

modification of the CS main chain with the AGE-CS monomer conjugation technique.

Page 167: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

141

No major peak shifts were detected at the 1033cm-1 peak s of CS-macromolecules up to

6h of reaction, indicating that there was no significant bonding of the CS hydroxyls by

the AGE monomer at these time intervals and that incorporation of the AGE vinyl group

into CS is limited to AGE reaction with the CS terminal primary amine (as detected by

the fluorescamine assay (Figure 5.15).

In order to avoid over incorporation of vinyl groups throughout the CS backbone,

the time of reaction must be monitored and limited. In this study, a 6h reaction time of

1000:1 AGE:CS molar ratio reaction was demonstrated to have ~84% conversion of CS

to vinyl-terminated CS. Further investigation of other AGE:CS molar ratio reactions (i.e.

100:1 AGE:CS) over the time course of reaction may reveal further possible time-

concentration combinations that result in vinyl-terminated CS with minimal side

reactions. The simple modification of CS to incorporate terminal end polymerizable

moieties was demonstrated here, and vinyl-terminated CS may be used in future studies

in order to synthesize CS brush polymers via free-radical “grafting-through”

polymerization strategies.

Vinyl-CS can be utilized in free-radical polymerization reactions for the

fabrication of CS-glycopolymer structures via the “grafting-through” technique (Figure

5.30). In free-radical chain polymerization (Figure 5.31), polymerization is initiated by a

reactive species R* produced from an Initator I (I R*) where the reactive species may

be a free radical, cation or anion (168). R* adds to a monomer molecule by opening the

π-bond to form a new reactive center. In general monomers will have the form

(CH2=CH-X). This process is repeated as more monomer molecules are successively

added to continuously propagate the reactive center (Figure 5.31). Polymer growth is

Page 168: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

142

eventually terminated by the destruction of the active center by an appropriate reaction

depending on the type of reactive center and reaction conditions.

Current relevant synthesis systems that have previously been utilized for the

synthesis of glycopolymers are outlined in (Table 5.5) (113, 118, 123, 125-126).

Polymerization has been achieved through several mechanisms including ring-opening

metathesis polymerization (ROMP) (120), reversible addition–fragmentation chain

transfer (RAFT) polymerization (121), atom transfer radical polymerization (ATRP)

(122), cyanoxyl-mediated free-radical polymerization and classical free-radical synthesis

(123-126). Although several strategies are available that have produced glycopolmyers

with low poly dispersity and high control, many of these methods are not directly

amenable to the fabrication of CS bottle-brushes (113, 118, 123).

The complex nature of CS due to its large molecular weight, charged groups and

functional groups, poses a challenge to the synthesis of CS bottle-brush structures via

conventional methods (115). Methods that require functional group protection are

especially problematic for larger polysaccharides such as CS where many functional

groups are present and the complete de-protection of the molecule may prove difficult.

In addition, systems requiring high purity are not desirable, as the CS being utilized is a

commercially available biomolecule derived from tissue sources and control over

contaminants is limited. Finally, CS is highly insoluble in organic solvents, reducing the

possible polymerization strategies to those amenable to aqueous reactions. Given these

limitations, the “living” radical cyanoxyl-mediated polymerization strategy or the

APS/TMEDA mediated polymerizations may be promising routes to the synthesis of CS

bottle-brush polymers from the vinyl terminated CS synthesized in this study.

Page 169: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

143

5.4.4. Grafting of CS to Poly(acrylic acid)

The carbodiimide mediated carboxylic acid-primary amine reaction was utilized

in order to investigate the “grafting-to” method of brush polymer synthesis due to the

water soluble, stable and biocompatible nature of carboxylic acid polymers. The

polymeric backbone of poly(acrylic acid) (PAA) was chosen for preliminary

investigation of the synthesis of CS-glycopolymer structures via the “grafting-to” strategy

via the carboxyl-amine reaction (Figure 5.7). PAA of 250kDa MW was utilized in order

to mimic the MW of the natural aggrecan protein backbone. PAA is a linear polymer

with pendant carboxylic acid groups and an enzymatically resistant hydrocarbon

backbone. It has been used previously in the hydrogel form with bioactive molecules for

the culture of cells and is shown to be non-toxic in in vitro studies (169).

Chondroitin sulfate was conjugated to the PAA backbone with high percent

conjugation, ranging from 40% to over 90% based on reaction conditions. Reaction

temperature, which can affect the motion of the polymer chains through Brownian

Motion, thereby affecting their collision rate and time was investigated but found to have

no significant effect on conjugation (Figure 5.25).

Ionic concentration was also investigated for its effect on CS-PAA conjugation.

Ionic concentration of the reaction solution is an important factor due to the charged

nature of both the PAA and CS chains. The ionic concentration can affect the

conformation of the molecule making the terminal primary amine more or less accessible

as well as changing the ability of the molecules to interact due to charge repulsions. In

Page 170: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

144

the case of CS, at low salt concentrations (ranging from 0.0001 to 0.01M Na+) and high

salt concentrations (on the order of 1M Na+) CS chains take on a more collapsed

conformation due to protonation of the CS chains and salt screening. At moderate salt

concentrations (~0.1M Na+) CS chains may take on a more extended rod-like

conformation due to intra- and inter-molecular electrostatic repulsion (73). Our results

indicate that Na+ concentration did significantly effect the conjugation of CS and PAA

(Figure 5.26). Low and high salt concentrations had lower conjugation than the salt

concentration of 1XPBS. All solutions were maintained at a pH of 7.4; lower salt

concentrations could not be tested with PBS buffer since buffering of the solution to pH 7

required NaOH which again increases the salt concentration. In a salt concentration of

0.69M (1X PBS) CS chains most likely take on an extended conformation. This

conformation may allow the terminal primary amine to be more accessibly. The

electrostatic repulsions between CS molecules and the charged PAA chain may however

hinder close packing of the chains in brush formation.

The ratio of CS:PAA in the reaction medium played a large role in the synthesis

of CS-PAA conjugates. As expected lower CS:PAA molar ratio reactions yielded higher

conjugations due to the large excess of carboxylic acid functional moieties compared to

primary amines, however, a high conjugation does not always translate to a high ratio of

CS attachment to the PAA backbone. From Table 5.4 it can be seen that lower molar

ratios of CS:PAA had higher conjugation, but the estimated number of CS chains

attached to each PAA chain was 1-5 in comparison to the high CS:PAA molar ratio of

20:1 where conjugation was low (~47%) but yielded ~9 CS chains attached to each PAA

Page 171: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

145

backbone. Higher CS:PAA molar ratios may be investigated in order to achieve higher

grafting density of CS onto the PAA backbone.

Reaction time was also investigated for the CS-PAA reaction. As seen with the

acrylic acid studies, the carbodiimide mediated CS-PAA reaction was not modulated by

time over a 24hr reaction window. Longer reaction times were investigated however and

conjugation was seen to decrease over a 120hr period. A one-phase exponential decay

was fit to the data with an R2 = 0.93. The predicted plateau conjugation from the non-

linear fit was ~21% with a time constant of 103.5h. This is an over 40% reduction in

percent conjugation from the originally detected conjugation of ~64% after 24h. As

described previously, this drop in percent conjugation may be due to non-covalent

interactions between the carboxylic acids of PAA and the amines of CS which can

transiently reduce the fluorescamine signal. Although conjugation is seen to decrease

over time, after 120h ~40% conjugation is maintained.

In addition to fluorometric monitoring of CS-PAA conjugation, rheometric

studies were conducted on CS-PAA samples reacted in optimal conditions (Ambient

temperature, 1X PBS, < 24hr reaction, molar ratio of 7.5:1 CS:PAA) in order to follow

the attachment of CS to the PAA backbone. The high molar ratio of 20:1 CS:PAA was

not investigated in this manner because reactions were not purified, and contamination

with a high percentage of un-attached CS chains may mask rhelogical changes broguh

about by CS-PAA conjugates. CS-PAA reaction solutions were investigated over time

for their viscosity. Viscosity of the reaction solution increased with time and fit well to a

one-phase association exponential curve (R2 = 0.99). The rate constant of the reaction

was 0.78h-1 corresponding to a time constant of 1.3h. Rheometric investigations of

Page 172: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

146

reacting solutions are a good indication of conjugation and polymer formation since

viscosity is proportional to molecular weight by the Mark-Houwink-Kuhn-Sakurada

equation (170):

[𝜂] = 𝐾(𝑀𝜐����)𝑎

Where [η] is intrinsic viscosity, 𝑀𝜐���� is the viscosity average molecular weight and K and

α are constants dependent on polymer conformation. Intrinsic viscosity is the limit of the

reduced specific viscosity as the concentration approaches zero, and is given by the

equation:

[𝜂] = lim𝑐→0

(𝜂𝑠𝑝𝑐

)

Where c is concentration (g/cm3) and ηsp is specific viscosity. Viscosity measurement

were conducted on a mixed solution with possible free PAA chains as well as free CS

chains, therefore viscosity measurements may have been reduced due to lower molecular

weight species within the sample. However, an increase in viscosity for reacted CS-PAA

samples versus simply mixed CS-PAA samples, demonstrates the formation of larger

molecular weight species within the solution. Therefore, rheometric data in conjunction

with fluorometric analysis of the primary amine chemical reaction further demonstrates

the covalent conjugation of CS to the PAA backbone.

Further modifications of the reaction medium as well as the CS:PAA ratio may

aid in increasing grafting density. The use of a reaction scheme that allows for longer

reaction times may also aid in increasing brush density as fabrication of surface brushes

has demonstrated various stages in brush formation before dense grafting can be achieved

Page 173: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

147

(166, 171-173). These studies demonstrate the feasibility of utilizing the CS primary

amine in the “grafting-to” approach of brush synthesis; however other chemistries and

strategies will need to be followed in order to synthesize brush structures with dense

attachment of CS chains. The estimated cost of the current “grafting-to” CS-PAA

synthesis strategy is listed in Table 5.6.

Page 174: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

148

Figure 5.30: Grafting-through polymerization of vinyl terminated chondroitin sulfate

Page 175: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

149

Figure 5.31: Chain growth polymerization and propagation of the reactive center as polymerization proceeds (168)

Page 176: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

150

(113, 118, 123, 125-126)

Tabl

e 5.

5: C

hain

Pol

ymer

izat

ion

Stra

tegi

es u

tiliz

ed in

the

fabr

icat

ion

of g

lyco

poly

mer

s

Page 177: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

151

Table 5.6: Estimated cost of CS-PAA macromolecule synthesis

"Grafting-To" Synthesis Supplies

Estimated Cost to Synthesize

1g MES Buffer $3.12 EDC/NHS $1.19

PD-10 Desalting Column $8.10 Chondroitin Sulfate $54.20

Polyacrylic Acid $27.31 Loss due to Purification Unknown

Total Cost CS-PAA $93.92

Page 178: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

152

The synthesis of a biomimetic aggrecan can have far reaching implications for

restoring the mechanical function of load bearing tissues, such as the intervertebral disc.

The modification and subsequent immobilization of CS at its terminal end is important

for the fabrication of CS brush structures that mimic the bottle brush organization of

proteoglycans such as aggrecan, where the bottle brush structure is critical for mechanical

function. In this work, we have presented a simple method for the end-functionalization

of CS via a terminal primary amine. Functionalization was achieved via three distinct

chemistries and resulted in vinyl-CS macromonomers that may be used in the

polymerization of CS into brush structures via the “grafting-through” technique. CS was

also attached to surfaces and polymeric backbones demonstrating its utility in the

“grafting-to” method of synthesis. From these studies, several pathways for the synthesis

of biomimetic aggrecan can be envisioned as outlined in

5.5. Conclusion

Figure 5.32. Although the

“grafting-to” technique of biomimetic aggrecan via caroboxylic acid-amine reactions was

demonstrated to be feasible in these studies, low conjugation and low grafting density

was associated with this technique. The epoxide reaction chemistry investigated here

resulted in high degrees of conjugation, was demonstrated to be specific to the terminal

primary amine and can react over a long period of time. These characteristics make the

epoxide-amine reaction chemistry promising for the synthesis of biomimetic aggrecan via

both the “grafting-to” and “grafting-through” techniques of brush synthesis. In addition,

the epoxide-amine reaction is unique in that it can be utilized in a step-growth synthesis

strategy for “grafting-through” polymerization.

Page 179: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

153

Figure 5.32: Flow chart of pathways to biomimetic aggrecan based on covalent binding chemistries investigated in this study.

(Red) Pathways not recommended for further investigation. (Yellow) Pathways with some limitations, recommendations have been made for further investigation. (Green) Pathways that

will be investigated further

Page 180: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

154

Chapter 6 : Investigation of the Epoxide-CS Primary Amine Reaction for the Synthesis of CS Brush Structures via the “Grafting-Through” Technique

Degeneration of the intervertebral disc is characterized by the loss of

proteoglycans (PGs), especially aggrecan, in the central nucleus pulposus region of the

disc. Aggrecan is a large proteoglycan (PG) found in load bearing tissues. It is

composed of a protein core to which chondroitin sulfate (CS) is covalently bound in a

very dense array (2-4nm spacing). Charged anionic groups on the CS chains draw water

into the intervertebral disc and electrostatic repulsions generated between closely packed

GAG chains resist deformation thereby allowing the tissue to distribute mechanical

forces(7, 69). Aggrecan degeneration is mediated enzymatically through cleavage of the

core protein and leads to numerous degenerative conditions such as arthritis and

intervertebral disc degeneration (22). The goal of this project is to develop an

enzymatically resistant biomimetic replacement for the large PG aggrecan by replacing

the core protein with a synthetic polymer, while maintaining the biologically active CS

bristles.

6.1 Introduction

To mimic the bottle brush structure of native aggrecan, we have utilized a

terminal primary amine on CS to selectively immobilize CS by its terminal end. The CS

terminal primary amine was investigated previously for reactivity to several functional

groups including aldehydes, carboxylic acids and epoxides. The epoxide-amine reaction

was chosen for further investigation due to its higher reactivity to the CS amine as well as

its relatively long window of reactivity. In addition, the reaction of epoxides to amines

does not require any specific activation. Simple base catalysis was utilized in the

Page 181: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

155

epoxide-amine reaction and the reaction generates no reaction bi-products(174) . A

unique property of the epoxy-amine reaction is the ability of epoxides to react to both

primary amines and the generated secondary amine (174-175). Epoxides and amines can

therefore be utilized in step-growth type polymerization. In the case of a di-epoxide and

a primary amine, step-growth polymerization will result in linear polymers (176).

6.2.1. Epoxide Reactivity

6.2. Background

Epoxide groups are very reactive due to the strain induced by the bond angles of

the epoxide ring (Figure 6.1). The strain energy has been found to be 13kcal/mole for the

epoxide ring. Almost all reactions in solution involve opening of the epoxide ring by a

nucleophile and addition of the molecule. In the general case of the reaction two

products are possible (Figure 6.2). Products resulting from group addition to the carbon

atom with the greater number of free hydrogen atoms are considered normal (Figure 6.2,

XVI) while, products in which group addition occurs at the oxygen atom are considered

abnormal (Figure 6.2, XVII) (159). Under basic or neutral conditions, the normal isomer

is usually the major or only product while in acidic conditions, the abnormal product

becomes more likely (159). For the purposes of our studies we will assume that reaction

products consist only of the normal isomer since reactions are conducted in basic

conditions.

Epoxides are fairly stable at neutral pH or above but can hydrolyze to a diol at

acid pH(149, 177). Epoxides are reactive to many functional groups including amines,

Page 182: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

156

hydroxyls, carboxylic acids, phenols, mercaptans and isocyanates(149, 178-179).

Epoxide reaction with hydroxyls (R—OH) requires high pH conditions, typically in the

pH range of 11-12 resulting in the creation of an ether bond (149). Amine nucleophiles

(R—NH2) react at more moderate alkaline pH values around pH 9.0 to result in

secondary amine bonds (149). Sulfyhryls (R—SH) are highly reactive with epoxides at a

more physiological pH range of 7.5-8.5 and result in thioether bonds(149). In general,

epoxide reactivity to nucleophiles follows the trend of increasing basicity for example,

alcohol > phenol > acid (178) and are most reactive to nucleophiles when base catalyzed

(174, 178, 180). When reacting epoxides with the terminal primary amine of CS it is

important to reduce possible side reactions with other possible nucleophiles of CS, most

notably the hydroxyls and carboxylic acids of the CS disaccharide. Therefore, pH was

controlled for all reactions at pH 9.4 in a sodium borate buffer.

6.2.2. The Epoxide-Amine Reaction

The addition of amines and other nucleophiles to epoxides generally requires

some degree of catalysis by species capable of hydrogen bonding, for example, hydroxyls

in the solvent, other primary amines, and hydroxyl groups produced in the addition

reaction (180-181). Preformed hydrogen-bonded complexes between epoxides and, for

example, hydroxyls, react with amine groups to form a transition state as seen in Figure

6.3. These intermediary transition complexes, formed by hydrogen bonding, help to

accelerate the rate of the epoxide-amine reaction.

In a system of epoxides and amines, several reactions can take place as illustrated

in Figure 6.4. Primary amines react with the epoxide ring to create a β-hydroxyl group

Page 183: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

157

while the amine molecule adds to the epoxide forming a secondary amine linkage. The

secondary amine can then be further reacted with additional epoxide to form a tertiary

amine linkage. The formed hydroxyl as well as other hydroxyls in the system are

reactive to the epoxide in a reaction called etherification, however this reaction is much

slower than the amine addition reactions and the β-hydroxyl formed can actually catalyze

the amine reaction further, allowing for autocatalyzation of the reaction.

The addition reaction of secondary amines and epoxides occurs by the same

mechanism as those seen for the primary amine reaction, however the secondary amine is

more sterically hindered. This results in a negative substitution effect on the reactivity of

the amino hydrogen in secondary amines as compared to the amino hydrogens of the

primary amine(182). In the general case, the rate of reaction of an epoxide group with an

amino hydrogen is slower when substitution on the nitrogen has already occurred. The

degree to which the secondary amine reaction is slower than the primary amine has

showed varied results as it is a difficult properly to measure. The substitution effect is

described by the ratio of the rate constant for secondary amine (k2) and primary amine

(k1) reactions with epoxides, (k2/k1). The (k2/k1) has been seen to range from 0.1 up to

approximately 1.1, and therefore some controversy exists as to the magnitude of the

negative substitution effect (180, 182). In addition the electrophilicity of the amine

molecules may have an effect on primary and secondary amine reactions modulating the

negative substitution effect (183). This effect however is very important in epoxy-amine

linear polymerization and cure systems as the rate of polymerization and structure of

cross-linked networks will vary greatly depending on the degree of the negative

substitution effect (180).

Page 184: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

158

The kinetic relationship between epoxide reactions with primary and secondary

amines is illustrated in the example kinetic curves seen in Figure 6.5 (174). As

demonstrated, primary amines must first react with epoxides to generate a population of

secondary amines. Tertiary amines may form immediately, however the negative

substitution effect may reduce the rate of this reaction. Primary amines begin to react

with epoxides immediately, resulting in a buildup of secondary amines and a reduction in

the epoxide and primary amine populations. The tertiary amine population increases

slowly, and then more quickly as more secondary amines are available for reaction. This

kinetic study suggests that there is not necessarily a step wise reaction that requires the

reaction of epoxides with all of the primary amines in the reaction before secondary

amines can react to form tertiary amines (174). These studies demonstrate the relative

kinetics of the epoxy primary and secondary amine reactions however, they were

conducted in bulk and on a particular epoxy-amine system so results may not be fully

transferable to solvent systems and other epoxy-amine reactions.

6.2.3. Step-growth polymerization of epoxides and amines

The reactivity of epoxides to both protons of the primary amines lends the epoxy-

amine system to use in step growth polymerization. Step-growth polymerization

proceeds via a step-by-step series of reactions between reactive sites. Each step utilizes

two co-reacting sites and generates a new covalent bond between the pair of functional

groups. Step-growth polymerization is a widely used commercial polymer synthesis

technique and today makes up a multi-billion dollar industry (184). The work of

Page 185: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

159

Carothers and his group at DuPont pioneered this technique in the 1950’s and 60’s with

the synthesis of materials such as nylon 6,6 and Poly(ethylene terephthalate) (PET) via

step-growth polymerization (184).

In the case of the epoxy-amine reaction, an epoxide group reacts with an amine

proton to create an amine covalent bond and at each step one epoxide and one proton is

used(175). The number of reactive sites per monomer (referred to as, functionality) and

the molar ratio between co-reactive sites are the main factors that control polymer

structure. In order to obtain linear polymers, both monomers must be bifunctional (175).

In the case of epoxy-amine reactions, a di-epoxide which has two epoxide functionalities

can be reacted with a single primary amine, which has two reactive hydrogen protons, in

order to generate linear polymers. For example, linear poly(aniline) has been synthesized

via the step-growth reaction of diglycidyl ether of bisphenol A (DGEBA) and aniline

(185).

In linear step-growth polymerization, the molar mass of the product grows

gradually, and the polydispersity tends to 2 which represents 100% of monomer

conversion (175). Functional groups react to form dimmers, trimers, etc. and very high

molecular weight products do not appear until the end of the reaction (Figure 6.6). Step

growth polymerization is advantageous in a system which incorporates biomolecules

since it can be conducted in aqueous solutions and does not require initiation and

propagation of a free-radical. Free radical propagation in chain growth polymerization

can be interrupted by other reactive groups in the biomolecule and often requires non-

aqueous reaction conditions. Chondroitin sulfate, which is primarily soluble in water has

Page 186: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

160

limited utility in chain growth polymerization, although some systems (i.e. APS/TMEDA

mediated polymerization) may be explored.

Some general considerations for step growth polymerization are that high %

conversions and long reaction times are necessary in order to achieve high molecular

weight polymers as demonstrated in Figure 6.6. In addition side reactions must be

avoided to avoid branching as well as chain termination (184). Generally a 1:1 molar

ratio of reactant functionalities is required for the synthesis of high molecular weight

polymers, however reactions can proceed even in an excess of one functionality (184,

186). Miller et al found that step-growth polymerization of a di-functionalized peptide

biomolecule and di-functionalized polyethylene glycol monomer in excess of PEG

resulted in % conversions of only 80% but 40% of the resulting polymer population had a

MW of 500kDa, a reaction of approximately 100 monomers (186). These properties of

step-growth polymerization make it a promising option for use with biomolecules and

step-growth polymerization was therefore further investigated for synthesizing

biomimetic chondroitin sulfate brushes.

6.2.4. Step-growth polymerization of CS macromolecules with PEG-DGE

In the previous study (Chapter 5), a primary amine at the terminal end of CS was

identified for commercially available CS (CS-4, Sigma C6737). The epoxide amine

reaction was demonstrated to reach the highest degree of primary amine conjugation in

comparison to both aldehyde-amine and carboxylic-acid amine reactions. In the current

study, the epoxide-amine reaction was studied further for its utility in the linear step

Page 187: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

161

growth polymerization of biomimetic CS-macromolecules using di-epoxide

poly(ethylene glycol) (Figure 6.7). CS was reacted with poly(ethylene glycol)-diglycidyl

ether, and the reaction kinetics investigated. Products were purified and analyzed for

molecular weight and chemical structure.

Page 188: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

162

Figure 6.1: Bond lengths and angles in ethylene oxide (the epoxide ring) (159)

Figure 6.2: Nucleophile addition to an epoxide ring results in two possible products normal, XVI and abnormal, XVII (159)

Page 189: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

163

Figure 6.3: Hydrogen bonding in the catalysis of the epoxy-amine reaction (180)

Page 190: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

164

Figure 6.4: Possible reactions in an epoxy-amine system. (174)

Page 191: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

165

Figure 6.5: Example reaction kinetics of an epoxide amine system (equivalent amounts of n-butyl-amine and phenyl glycidyl ether at 50C̊) (174)

Page 192: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

166

Figure 6.6: % Conversion vs molecular weight for typical step-growth polymerization reactions (187)

Page 193: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

167

Figure 6.7: Schemiatic of linear step-growth polymerization of amine terminated CS using the epoxy-amine reaction.

Page 194: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

168

6.3 Methods

6.3.1. Reaction of CS to PEG based di-epoxides

In order to achieve linear polymerization of primary amine terminated CS, CS

was reacted with several molecular weights of poly(ethylene glycol)-diglycidyl ether

(PEG-DGE) (Table 6.1) where each diglycidyl ether has two epoxide functionalities.

PEG based di-epoxides were choosen due to the biocompatibility and hydrophilic nature

of PEG (188). In a general synthesis procedure CS was reconstituted from a lyophilized

state into 0.1M SBB, pH 9.4 at 25mg/mL (1.13mM). Samples were reacted in 10ml

volumes in 15ml centrifuge tubes. The solution was mixed thoroughly via vortex mixing

to ensure a homogeneous solution. DGE was then added to the CS solution at a

particular molar concentration (i.e. 10, 20, 40 100, or 200mM). The solutions were then

mixed thoroughly and placed into a water bath (25̊C, 37̊C, or 45̊C) with continuous

shaking for 96hrs. Samples were assayed over the reaction period for reaction of the CS

primary amine using the fluorescamine assay described previously in Chapter 5.

Reaction rates of the CS primary amine reaction were determined by fitting a one-phase

association curve to % conjugation data (n >= 4 for each data point) obtained over the

96hr reaction period. The GraphPad Prism 5 data analysis software package was utilized

for curve fitting and data analysis.

6.3.2. Purification of the CS-DGE reaction

Page 195: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

169

Un-reacted DGE monomers were removed from the reaction by extensive

dialysis. Reaction samples in 20ml volumes were loaded into 6-8K MWCO regenerated

cellulose dialysis membranes (40mm flat width, 28cm length, pre-wetted in distilled

deionized (DI) water for 30min). Dialysis was conducted over 96hrs against 1.65L of DI

water with daily water changes. Purified samples are then lyophilized, resulting in a

white cotton-like powder. Typical yield from a 20mL reaction is approximately 100mg.

6.3.3. Characterization of CS-macromolecule chemical structure

Purification and chemical structure of the CS-macromolecules were confirmed via

1H-NMR (300Mhz UnityInova NMR Spectrometer). Samples were prepared at 30mg/ml

in D2O and scanned at ambient temperature. Scans were taken at a spin of 20 and a

number of transients of 64. The ACD/NMR Processor (Academic edition) was used in

order to analyze peak locations and integrated areas. Example 1H-NMR spectra for raw

materials PEG-DGE and chondroitin-4-sulfate are presented in Figure 6.8. Spectral

assignments for PEG-DGE were predicted using a NMR-predictor (189) and found to be

in agreement with previously published NMR assignments for polyethylene glycols and

epoxides (190). CS spectral assignments were established based on several previously

published chondroitin sulfate 1H-NMR studies and are in good agreement with the

current spectra (152-153, 155, 191-195). Purified, lyophilized samples were also

investigated with attenuated total reflectance-fourier transform infrared spectroscopy

(ATR-FTIR) in order to monitor changes in the chondroitin sulfate chemical signature

after conjugation with PEG-based diglycidyl ethers. Scans were taken at number of scans

Page 196: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

170

64 and analyzed with Omnic 8.1.11 software (ThermoFischer Scientific). Example ATR-

FTIR spectra for raw materials chondroitin-4-sulfate and PEG-DGE are presented along

with spectral assignments in Figure 6.9. FITR spectra of PEG-DGE have bands that

correspond both to PEG (196) and epoxide moieties (197) in the mid-IR range.

Chondroitin sulfate bands corresponding to sulfate, amide, sugar and hydroxyl groups of

CS have been previously identified and are in good agreement with the current spectra

(154-157). Primary amine bands (two bands from 3400-3300 and 3330-3250 cm-1) (198)

corresponding to the terminal primary amine of CS could not be distinguished in the CS

spectra due to a broad OH peak between 3600 and 2900cm-1.

6.3.3. Gel Permeation Chromatography of CS-Macromolecules

Size exclusion chromatography (SEC) is a powerful and important technique for

characterizing macromolecules. SEC is a form of liquid chromatographic separation of

macromolecules where separation is based on molecular size. Gel permeation

chromatography (GPC) is a form of SEC and is often taken synonymously with SEC.

The term “gel” refers to the use of a non-rigid or semiriid organic gel stationary phase.

The purpose of SEC is to provide molecular weight distributions (MWD) for polymeric

materials. Although SEC is based on molecular size it relates to molecular weight

information by the relationship between linear dimension and molecular weight in a

freely jointed polymeric chain (random coli). The root mean square end-to-end distance

or the radius of gyration is proportional to the square root of the molecular weight, and

therefore the log of either distance is proportional to ½ the log of the molecular weight.

Page 197: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

171

The GPC technique is comprised of four major components, the sample, a

stationary phase (gel packing material), a liquid mobile phase and a detector. The

stationary phase of the GPC column consists of porous particles with controlled pore size.

This pore size determines the range of molecular weights that can be separated within the

column. In a typical experiment, the sample, contained in the liquid mobile phase is

injected into the column and the mobile phase passes through the column at a fixed flow

rate, developing a pressure gradient across the length of the column (Figure 6.10) (199-

200). The sample polymer molecules pass into the column as a result of the pressure

gradient. The smaller macromolecules are able to penetrate the pores of the stationary

phase as they pass through the column, lengthening their path through the column and

thus slowing their elution from the column. Larger macromolecules are too large to be

accommodated in the stationary phase pores and remain in the interstitial space, giving

them a minimal path through the stationary matrix resulting in earlier elution from the

column. A detector placed at the end of the column responds to the elution of the

macromolecules in the mobile phase by generating a signal (peak) fir each group of

macromolecules that passes through generating a chromatogram (Figure 6.10). The

differently sized macromolecules elute from the column at different times changing the

appearance of the chromatographic bands. Typically, the chromatogram shows a

continuum of molecular weight components that are not resolved within the single peak,

however shifts in the peak and changes in peak shape give rise to different forms of

molecular weight determinations (199-200).

Several parameters of the GPC method must be determined to ensure proper

chromoatographic conditions (Table 6.2). Since SEC is a non-interaction separation

Page 198: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

172

mechanism based on the separation of macromolecules by size factors, therefore, any

mechanism that interferes with a purely size based separation is undesirable. For

example, sample interactions with the column, sample aggregation, shape and charge

interactions can all interfere with the size exclusion mechanism, introducing error into the

molecular weight distribution determination. Parameters that must be considered include

mobile phase composition, stationary phase chemistry, column porosity, flow rate,

dissolution conditions, temperature, sample stability, samples concentration, sample

geometry standard selection and detection method (200-201). In addition to

consideration s that need to be made for the CS chemistry and structure, an added

complexity is introduced with the incorporation of an un-charged synthetic polymeric

component.

Hybrid materials (i.e. copolymers) and branched polymers present increased

complication in SEC analysis (200). Copolymers are characterized by chemical

composition distribution that is represented by the mass fraction of molecules of a given

copolymer composition vs. the copolymer composition. Average composition changes

along the polymerization process and also possibly with hydrodynamic volume. Since

macromolecules in SEC are fractionated according to hydrodynamic volume rather than

molar mass, copolymer composition can have distorted molecular weight distributions

(200). For homopolymers, instantaneous mass is proportional to the differential

refractometer (DR) signal. With copolymers, the specific refractive index increment

depends on the instantaneous composition which may change with hydrodynamic

volume; therefore the instantaneous MWD is not monodisperse. The variability in molar

masses in the detector when a complex polymer is analyzed introduces bias in the MWD

Page 199: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

173

(200). Differential refractometers results are therefore only an estimate for MWD and are

often combined with other detection methods. MWD bias is further magnified when

resolution is poor. Imperfect resolution can occur from a non-exclusion (secondary)

fractionation caused by differential column and solvent interactions of the different

polymers of the copolymer system. Non-exclusion effects may shift and distort the

chromatograms, resulting in both positive and negative molecular weight deviations

(200). Overall, linear copolymers are more simple for MWD determination than

branched copolymers. Branched copolymers of a given molar mass and composition

have smaller hydrodynamic volume than its linear analogue and volume reduction is

more apparent with increasing in branching and again changes in hydrodynamic volume

interrupt the SEC process (200). A difficulty also arises when determining an appropriate

MW standard for copolymers, and therefore triple detection SEC (DR, intrinsic viscosity

and light scattering) is often used to characterize copolymer and branched systems

without using molecular weight calibrations. Certain copolymer conditions have been

identified for the proper determination of MWD with SEC: (1) instantaneous

distributions of the molecular weights and of the chemical compositions are both narrow

and (2) the instantaneous average composition varies monotonically with the molecular

weight (200). For branched polymers it is important also that the average number of

branches per molecules increases monotonically with the molar mass and that the average

composition does not change with molar mass (200). These conditions are often difficult

to meet, especially for hybrid bottle brush type polymers, therefore, viscometric and

imaging techniques (i.e. atomic force microscopy, transmission electron microscopy) are

often looked to provide structural characterization of these materials (106).

Page 200: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

174

The conformation and charges associated with CS and the large size and

aggregation of PGs introduces complexity into the separation of these molecules (202)

however several solvent/medium/column systems have been investigated for their

separation and may be amenable to the separation of the synthesized CS-macromolecules.

Proteoglycans have been separated in HPLC using the ProPac PA1 anion-exhange

column using 8M urea as the solid phase in order to dissociate aggregated PGs (203).

GAG separation and MW calculation has been accomplished using Na2SO4-NaH2PO4,

pH 6 as the mobile phase and Protein Pak 125 and 300 columns (Waters) in an HPLC

system (202). Generally protein standards of varying known molecular mass (670,000 to

24,000) are used to calibrate SEC of proteoglycans, however they have been

demonstrated to have skewed results due to the different chemical and structural nature of

proteins (202-203).

Given these intricacies of GPC analysis and the complications associated with

proteoglycan, copolymer, and brush polymer SEC, an analytical techniques contract lab

was contacted for GPC method development. Method development for GPC was

conducted by Jordi Labs (Bellingham, Massachusetts). Jordi labs is a contract analytical

services laboratory which also manufactures and distributes its own custom built products

and columns for GPC. Jordi labs provides specialized analytical testing of polymers as

well as method development and was contracted for method development for the GPC

analysis of three formulations of CS-macromolecules (PAA based, EG-based and PEG-

based CS macromolecules). Although GPC was attempted on-site at Drexel University,

difficulties were encountered with CS-macromolecule aggregation and identification of

Page 201: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

175

appropriate standards in typical size exclusion columns (Sephacryl 300, molecular weight

range of 10 to 1500kDa).

Page 202: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

176

Figure 6.8: PEG-DGE (top) and Chondroitin-4-sulfate (bottom) 300MHz 1H-NMR spectra with peak assignments (189-190, 193). See Appendix 2 for alternate view of CS structure.

Page 203: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

177

Figu

re 6

.9:

FTIR

Spe

ctra

and

spec

tral a

ssig

nmen

ts fo

r Cho

ndro

itin

Sulfa

te a

nd P

EG-D

GE

Page 204: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

178

Figure 6.10: Schemiatic description of the time course and mechanism of size exclusion chromatography (199)

Page 205: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

179

Table 6.1: Poly(ethylene glycol)-diglycodyl ethers of varying molecular weight

Di-glycidyl Ether Molecular Weight (g/mole)

Mean Square End-to-End Distance (Freely-Rotating Chains Model)

Ethylene Glycol- Diglycidyl Ether (EG-DGE) 174.2 0.5nm

Poly(ethylene glycol)-Diglycidyl-Ether (PEG-DGE) 526 5nm

Poly(ethylene glycol)(1000)-Diglycidyl Ether (PEG(1K)-DGE) 1000 12.5nm

Page 206: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

180

Table 6.2: Factors involved in GPC and Key Considerations (201)

Primary factors in GPC Implications and Concepts to Consider

Mobile Phase Composition Should be a strong Solvent for the polymer being analyzed and Solvent must have greater affinity for solvent than stationary phase.

Column Chemistry Should promote strong interactions between solvent and column and reduce interactions between column and sample

Column Porosity Must cover complete MW range of the sample Flow Rate Can be varied over a narrow range without adversely affecting calculated MW. Typical

flow rates range from 0.5-1.5ml/min.

Dissolution Conditions Conditions used to dissolve sample must avoid sample degradation and aggregation

Temperature Increasing temperature can prevent column sample interactions and result in improved separation. Solvent viscosity also decreases with temperature reducing back pressure.

Sample Stability All conditions must be managed to avoid shear degradation, hydrolytic degradation and thermal degradation. In addition, Sample aggregation and crosslinking can increase MW overtime and must be avoided.

Sample Concentration Column overloading can result in degraded resolutions and introduce significant error in MW calculations. Typical concentrations range from 0.1-4mg/ml for random coil polymers less than 800,000MW to 0.1-0.5mg/ml for high MW polymers while oligomers and monomers can be analyzed at higher concentrations. Generally lowest concentration should be used

Sample Geometry/ Standard Selection

Polymeric molecules may adopt a range of conformations including random coil, rigid rods, or spheres which can alter the way a polymer travels through a column. Standards chosen will ideally have the same chemistry as the sample, but in the case similar chemistry is not available, best standards will have the same geometry of the sample.

Detection Method Many universal concentration detection methods are available including the differential refractive index detector (DRI) and UV/IR detectors. A second/third molecular weight sensitive detector (triple detection)can be added to the SEC system providing direct means of absolute MW determination without external standards (i.e. Low angle/multi-angle/right-angle laser light scattering detection, and viscometric detection)

Page 207: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

181

6.4.1. Influence of temperature, DGE concentration and DGE molecular weight on CS

amine-DGE reaction kinetics

6.4 Results

CS was reacted to di-epoxides over 96hrs and monitored for primary amine

content in order to follow the reaction of the CS primary amine with the epoxides. The

reaction of CS primary amines to epoxides progressed with time and was modulated by

both temperature and di-epoxide concentration (Figure 6.11 and Figure 6.12). One-phase

association curves were fit to the % conjugation versus reaction time plots. For PEG-

DGE reactions to CS at 10mM and 20mM DGE concentrations, one-phase association

were a poor fit for the data at 21̊C (R2 = 0.339 and 0.674 respectively). This may be due

to lack of data at longer time points which would be required to fit one-phase association

curves with slower reaction rates. At higher reaction temperatures however, the one-

phase association curve fit is more accurate for all DGE concentrations investigated.

Reaction rate was increased from 0.032/h to 0.057/h at a 10mM PEG-DGE concentration,

and from 0.056/h to 0.093/h at 20mM PEG DGE concentration. Reactions conducted

with 100mM PEG-DGE resulted in relatively fast reaction rates at all temperatures (K =

0.093/h and K = 0.155/h respectively) with reactions at 45̊C proceeding too quickly to

determine reaction rate in the time frame investigated. In addition increased reaction rate

at the higher temperatures and PEG-DGE concentrations investigated, the predicted %

Conjugation plateaus were also higher at higher reaction temperatures and DGE

concentrations (Figure 6.11). Highest Amine reaction plateaus were predicted to be 81%,

92% and 99% for 10mM 45̊C, 20mM 45̊C, and 100mM 37̊C PEG-DGE-CS reactions

Page 208: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

182

respectively. Similar trends were seen for EG-DGE-CS reactions conducted at 20mM,

40mM, and 200mM EG-DGE concentrations where an increase in temperature increased

reaction rate as did an increase in EG-DGE concentration (Figure 6.12).

Due to the higher % Conjugations of CS primary amines achievable at 45̊C

compared to 21̊C and 37̊C as demonstrated by higher predicted plateau % conjugations,

the 45̊C reaction was investigated further. Three different DGE backbone monomers of

varying molecular weight and epoxide spacing were investigated (Table 6.1) for their

reaction to CS at 45̊C and varying DGE concentrations. Reactions at 1mM DGE

concentration (~ 1:1 CS:DGE ratio) did not progress greatly with time and one-phase

association curves were a poor fit (Figure 6.13). At higher DGE concentrations, reactions

for EG- and PEG-DGE MWs progressed with time and were better fits to the one-phase

association relationship (R2 > 0.9) while PEG(1K)-DGE reactions appeared to react too

quickly to be fit to a one-phase association in the time frame investigated. PEG(1K)-

DGE reactions for concentrations at 10mM or above likely have reaction rates faster than

0.88/h, and Tau of 1.13h. Experimental limitations for collecting reaction data earlier

than 2h, make it difficult to accurately predict one-phase association curves for reactions

occurring with a time constant in the range of PEG(1K)-DGE. Reactions progressed

significantly faster for PEG(1K)-DGE reactions compared to EG-DGE and PEG-DGE

reactions (Table 6.3 A) however, reaction plateaus were similar at 20mM and 100mM

DGE reactions reaching between 86.69%-89.07% and 96.82%-97.18% conjugation

respectively (Table 6.3 B). At 10mM reactions however, EG-DGE reactions had

comparatively lower % conjugation plateau values as well as reaction rates. (Table 6.3).

Page 209: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

183

% Conjugation rate as well as plateau increases with increasing temperature, DGE

concentration as well as DGE MW (PEG(1K)-DGE > PEG-DGE > EG-DGE).

6.4.2. Purification of CS-macromolecules

CS and DGE reaction solutions were purified via extensive dialysis against DI

water. The dialysis process for PEG-DGE and EG-DGE synthesis was monitored daily

via 1H-NMR analysis in order to validate the sample purification process (Figure 6.14 and

Figure 6.16). MWCO for dialysis bags was between 6000-8000 MW. Since CS is

approximately 22,000MW, both CS and synthesized CS macromolecules (PEG-CS, EG-

CS, and PEG(1K)-CS will remain within the sample. DGE monomers on the other hand

are much smaller in MW, ranging from ~172 to ~1000 Da. These DGE monomers can

pass easily across the chosen membranes and therefore be purified out of the sample.

For both samples, distinct contributions from the DGE monomer could be seen in

raw samples (0h of dialysis). Strong ethylene glycol peaks were resolved at 3.6ppm as

well as epoxide peaks at 2.9 and 7.7ppm. CS peaks were also resolved, however, several

CS peaks were distorted by contaminants seen in the spectra for the PEG-DGE raw

material (Figure 6.8). With dialysis, however, CS peaks became better resolved and

peaks corresponding to the DGE monomers were reduced. For both DGE monomers,

epoxide peaks were still present after 24h of dialysis (Figure 6.15 and Figure 6.17) as well

as some residual contaminant peaks in the 3ppm peak region. After 48hrs of dialysis,

PEG-CS and EG-CS 1H-NMR spectra closely resemble that of natural CS, with the

exception of the incorporation of a distinct ethylene glycol peak at 3.66ppm in the PEG-

Page 210: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

184

DGE spectrum. For the EG-DGE sample, only a small change is seen in the spectra in

the 3.66ppm region, however no distinct peak is present. Both EG-CS and PEG-CS

samples were completely purified of un-reacted DGE monomer by 48h of dialysis as

evidenced by loss of the epoxide peaks, loss of peaks corresponding to DGE monomer

contaminant, as well as reduction in the ethylene glycol peaks. After an additional 24h of

dialysis (72h total), minimal changes to the spectra are seen indicating that additional

dialysis time is not required for PEG-CS and EG-CS, however, it is important to note that

additional dialysis does not damage the CS macromolecules. PEG(1K)-CS was dialyzed

according to similar procedures and investigated with 1H-NMR after 48h Figure 6.18.

After 48h of dialysis, small residual epoxide peaks are present in the PEG(1K)-CS

spectra. The larger size PEG(1K)-DGE monomer may require more than 48h of

membrane dialysis for full purification, all samples will therefore be dialyzed for at least

72h in order to ensure complete purification of unreacted DGE monomer.

6.4.3. 1H-NMR and ATR-FTIR analysis of CS-Macromolecule chemical structure

Synthesis of PEG-CS and EG-CS were monitored using 1H-NMR over the 96h

reaction time. Samples were prepared at 20mM DGE concentrations as previously

described, and extensively dialyzed after 24, 48, 72, and 96h of reaction, lyophilized,

then investigated with 1H-NMR (Figure 6.19 and Figure 6.20). In both EG-CS and PEG-

CS synthesis, peaks corresponding to CS were distinctly maintained. For PEG-CS

samples, progressive increase in the 3.66ppm peak height can be seen indicating

increased addition of PEG into the PEG-CS samples with time (Figure 6.19). For EG-CS

Page 211: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

185

samples a distinct ethylene glycol peak was not resolved, this may be due to the relatively

small ethylene glycol sequence in the EG-DGE monomer. It is still likely that EG-DGE

was incorporated into the EG-CS samples, as evidenced by small spectral change in this

region and fluorescamine reaction kinetics data which indicates strong reaction of the CS

primary amine with EG-DGE. In order to better assess the changes to the CS molecule

with the incorporation of PEG-DGE and EG-DGE, 1H-NMR peak integrated areas were

determined for key peaks corresponding to CS (Figure 6.21). For PEG-CS samples,

minimal change was seen in integrated area of peaks corresponding to the GalNAc H1,

GalNAc H2 and H3, and GlcUA H2 protons over the 96h reaction time. A small increase

is seen in the GlcUA H3 proton integrated area. Interestingly, a progressive increase in

the GlcUA H4,5 and GalNAc H5 and H6 regions is seen over the reaction time, this peak

region is also the region which contains the polyethylene glycol peak centered at

3.66ppm. The successive increase of this peak region likely indicates addition of PEG

onto CS over the reaction time. EG-CS samples showed larger variability in peak

integrated areas for all peaks over the 96h reaction time. The peak region containing the

3.66 ethylene glycol signal did not show a progressive increase in integrated area, likely

due to the relatively small portion of ethylene glycol in the EG-DGE monomer. For both

PEG-DGE and EG-DGE all peaks showed slight fluctuations in integrated area with

reaction time, however, the only progressive trend with reaction time is seen with the

3.86-4.05 peak region of PEG-DGE.

Final formulations of EG-CS, PEG-CS and PEG(1K)-CS synthesized with 20mM

DGE concentration were investigated for their chemical composition via 1H-NMR (Figure

6.22 and Figure 6.23)and ATR-FTIR (Figure 6.24) after purification. 1H-NMR Peaks

Page 212: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

186

corresponding to an ethylene glycol component are evident in both PEG-CS and

PEG(1K)-CS samples while all CS peaks are maintained (Figure 6.22). Integration of 1H-

NMR peak areas corresponding to CS protons demonstrates maintenance of the major CS

protons and relative consistency in peak integrated areas for all formulations except in the

3.86 to 4.05 peak region. Peak integrated area in this peak region increases with

increasing DGE MW, likely due to the distinct ethylene glycol peak centered at 3.66 that

occurs in this region. ATR-FTIR spectra for EG-CS, PEG-CS and PEG(1K)-CS

macromolecules were consistent with that of natural CS (Figure 6.24). No strong peaks

corresponding to the EG-DGE, PEG-DGE or PEG(1K)-DGE monomers were detected,

however a change in peak shape is seen for the CS Amide I band. Of note, is the lack of

change in the ATR-FTIR peak centered at 1032cm-1 which arises from the C-O stretching

mode in CS. Spectral shifts in this region would be expected to occur with changes to the

hydroxyl group orientations in CS (156), and therefore, a lack of change in this spectral

region indicates preservation of the CS hydroxyl groups in EG-CS, PEG-CS and

PEG(1K)-CS samples .

6.4.4. Molecular weight characterization of CS-macromolecules

Gel permeation chromatography (GPC) was conducted by Jordi Labs, a contract

laboratory specializing in liquid chromatography. Several attempts were made in order to

determine the molecular weights of CS-macromolecules synthesized with varying

synthetic backbones. The chosen column for GPC size exclusion chromatography was

the JORDI DVB Sulfonated Mixed Bed column, 250mm x 10mm (ID) (204). This

column is base on polymerized divinylbenzene packing material and is utilized in the

separation of water soluble polymers due to hydrophilic group association to the DVB

Page 213: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

187

backbone (204). A sulfonated column was chosen because sulfonated columns have a

negatively charged surface which can separate water soluble polymers with negatively

charged surfaces, for example, organic acids, polysaccharides, starches, cellulose and in

our case chondroitin sulfate. Sulfonated columns also allow for the separation of

negatively charged polymers without the need for high salt concentrations (typical

solvents for this phases include aqueous/organic mixtures such as water/methanol)

making them especially useful in light scattering analysis where high salt concentrations

can compromise the system (204). A mixed bed column was chosen because the sample

molecular weight was unknown, and may have had a broad molecular weight

distribution, which is better detected by the molecular weight range of the mixed bed

column (10,000 – 700,000) (204).

A mobile phase of 85/10/5/0.1% Water/Acetonitrile/Isopropyl

alcohol/Phosphoric acid was chosen for CS and the CS-macrmolecules based on

solubility and column compatibility. In addition, sample preparation included filtration

through a 0.20μm disposable Nylon filters in order to remove debris and aggregates.

Concentrations ranging from 2.0 – 6.7 mg/ml (Table 6.4) were used for sample loading

and a system flow rate of 1.2 ml/min was found to be ideal. Injection size varied from 50

to 200μL for the different sample solutions ( Table 6.4). The column temperature was

maintained at 35° C in order to aid in reducing secondary column associations. Protein

standards were not compatible with the chosen column type, polysaccharide standards

were therefore used as a first attempt in MWD characterization. Polysaccharide

standards with a concentration of 1.0 mg/ml were used with the following molecular

weights: 710K, 380K, 200K, 106K, 45.9K, 22K, 11.2K & 5.6K and an injection size of

Page 214: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

188

100 μl. The sample was monitored at a sensitivity of 8 and a scale factor of 20 with a

WATERS 410 differential refractometer. Data acquisition and handling was made with

Jordi GPC software.

From sample chromatograms, and standard calibration curves, differential weight

fraction molecular weight distribution plots can be generated for the estimation of sample

molecular weight distribution and polydispersity (Figure 6.27). Three different molecular

weight averages were provided by Jordi labs for the characterization of sample MWD:

number average molecular weight (Mn), weight average molecular weight (Mw), and Z

average molecular weight (Mz) and can be calculated as follows (199):

𝑀𝑛 = ∑ 𝑁𝑖 𝑀𝑖𝑖

∑ 𝑁𝑖𝑖=

∑ 𝑊𝑖 𝑖

∑ 𝑊𝑖𝑀𝑖�𝑖

=∑ ℎ𝑖 𝑖

∑ ℎ𝑖𝑀𝑖�𝑖

𝑀𝑤 = ∑ 𝑁𝑖 𝑀𝑖

2𝑖

∑ 𝑁𝑖𝑖 𝑀𝑖=∑ 𝑊𝑖 𝑀𝑖𝑖

∑ 𝑀𝑖𝑖=∑ ℎ𝑖 𝑖 𝑀𝑖

∑ 𝑀𝑖𝑖

𝑀𝑧 = ∑ 𝑁𝑖 𝑀𝑖

3𝑖

∑ 𝑁𝑖𝑖 𝑀𝑖2 =

∑ 𝑊𝑖 𝑖 𝑀𝑖2

∑ 𝑊𝑖𝑖 𝑀𝑖=∑ ℎ𝑖 𝑖 𝑀𝑖

2

∑ ℎ𝑖𝑖 𝑀𝑖

Where Ni and Wi are the number and weight respectively of molecules having molecular

weight Mi and the subscript i is an index representing all of the molecular weights present

in the sample. The third representation (farthest right) represents the method for

obtaining Mn, Mw and Mz from the GPC chromatogram where hi is the height (from

baseline) of the GPC curve at the ith elution increment and Mi is the molecular weight of

the eluting species at this increment (Mi is obtained from the standard calibration curve)

(199). Mn provides information about the lowest molecular weight portion of the sample

Page 215: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

189

while Mw is the average closest to the center of the peak and Mz represents the highest

molecular weight portion of the sample (199). Polydispersity index (PDI) which is

defined as Mw/Mn was also provided by Jordi labs and gives an indication of the

broadness of the molecular weight ranges in the sample.

Results from GPC analysis against polysaccharide standards are presented in

Table 6.5. This standard was chosen due to limited solubility of other charged standards

in the chosen mobile phase. Since, the polysaccharide standards are not charged, it was

expected that molecular weights would be exaggerated for ionic polymers. CS control

sample showed a monomodal normal distribution with a weight average molecular

weight of 295K. Unfortunately, PAA and PAA CS was observed to elute poorly in the

chosen SEC system. PAA did not initially elute through the column however with

replicate injections a peak was observed which increased in intensity with each injection.

This is typical of polymers which are absorbing to the column phase. A steady state was

reached after multiple injections which provided a repeatable result but the weight

average molecular weight was low (53K) (PAA expected MW 250,000 based on retailer

information); this again is consistent with nonspecific interactions with the column phase.

EG-CS and PEG-CS were observed to show repeatable distributions without prior sample

injections with monomodal distributions. The observed molecular weights were higher

than for the CS control by nearly 80K. PAA and PAA-CS samples were not further

investigated due to their inherent incompatibility with the chosen column. A separate

GPC system will likely have to be developed for CS-macromolecules with varying

synthetic polymeric backbone chemical composition.

Page 216: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

190

In order to improve upon these results, polystyrene sulfonated standards were

investigated as possible compatible standards. In a second study samples were prepared

in the same mobile phase, to a concentration of 2.0mg/ml for EG-CS and PEG-CS. CS

sample were prepared at 2.5mg/ml and all samples were filtered using a 0.2μm disposable

Nylon filters and run in duplicate in the same solvent. The system was run at a flow rate

of 1.2 ml/min on a JORDI DVB Sulfonated Mixed Bed column, 250mm x 10mm (ID)

and again, column temperature was maintained at 35° C. Polystyrene Sulfonated

standards with a concentration of 0.5 mg/ml were used at the following molecular

weights: 1020K, 679K, 305K, 148.5K, 65.4K, 33.5K, 15.8K, 6.43K, 3.42K & 891 with

an injection size of 100μl. Samples were similarly monitored at a sensitivity of 8 and a

scale factor of 20 with a WATERS 410 differential refractometer and data acquisition

and handling was conducted with Jordi GPC software.

A chromatogram overlay of polystyrene sulfonate standards and the samples of

CS, EG-CS and PEG-CS is presented in Figure 6.25. Peaks of interest are eluted at

approximately 9min. Negative peaks eluting later in the chromatogram (~15min) are due

to solvent effect and are unrelated to the samples. Similar negative peaks are seen when

comparing standards to the samples, confirming this assumption. A well fit polystyrene

sulfonated calibration curve for molecular weight was established from the standards

(Figure 6.26) and used for generation of the differential weight fraction (Wf ) molecular

weight distribution curves for CS (Figure 6.27) and the samples. Mn, Mw, Mz, and PDI

were calculated from the chromatographs and results are presented in Table 6.6. CS Mw

molecular weight was 29,113, near the expected value of 22K for CS-4 and Mz values for

CS (average Mz = 62,227) are consistent with contamination with larger MW CS-6

Page 217: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

191

which has a MW of ~65K (65, 139). EG-CS had an average Mw of 38,916 while PEG-

CS had an average Mw of 42,464. Polydispersity of all samples was between 2.10 and

2.36. The use of a polystyrene sulfonated standard increased the accuracy of molecular

weight results.

6.4.4.1. Validation of GPC molecular weight estimation of CS and CS-

macromolecules. GPC evaluation of CS molecular weight was also validated by

calculation of the viscosity average MW of CS where the Mark-Houwink-Kuhn-Sakurada

relationship between intrinsic viscosity and molecular weight is (170, 205):

[𝜂] = 𝐾(𝑀𝜐����)𝑎

Where [η] is intrinsic viscosity, 𝑀𝜐���� is the viscosity average molecular weight and K and

α are constants dependent on polymer conformation. Intrinsic viscosity is the limit of the

reduced specific viscosity as the concentration approaches zero, and is given by the

equation:

[𝜂] = lim𝑐→0

(𝜂𝑠𝑝𝑐

)

Where c is concentration (g/cm3) and ηsp is specific viscosity (205). K and α values have

been determined for CS to be 1.7x10-5 and 1.01 respectively (0.2M NaCl, 25̊C) (206).

Viscosity measurements were conducted on serial dilutions of natural CS from 0.006g/ml

to 0.015g/ml (0.2M NaCl solution, 25C̊) on a cone-on-plate rheometer (AR2000ex

rheometer, TA Instruments, 1̊ Cone angle) with shear rate linearly ramped from 50-200/s

(5 points collected) . Concentration range was chosen such that specific viscosity was

between 0.2 and 1.9 for accurate intrinsic viscosity determination (170). Average ηsp/c

Page 218: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

192

was plotted vs concentration, and the zero concentration viscosity, [η], extrapolated from

the linear fit (Figure 6.28). Intrinsic viscosity was determined to be 38.64 ml/g, similar to

literature values for CS-4 intrinsic viscosity (34ml/g (207)). Viscosity average

molecular weight, 𝑀𝜐���� ,for CS was then determined as:

(𝑀𝜐����) =

⎜⎛

[𝜂]100�

𝐾

⎟⎞

1𝛼

where intrinsic viscosity is divided by 100 based on the original calculations of K and α

by Wasteson et al (206). 𝑀𝜐���� for CS was determined to be 20,580Da. The general

relationship of 𝑀𝜐���� to Mn and Mw is, Mn < 𝑀𝜐����< Mw (205). From GPC analysis, Mn was

estimated at 12,360Da and 29,110Da respectively. In these studies, 𝑀𝜐���� falls between the

Jordi Labs analysis of CS Mn and Mw, validating the GPC technique for evaluation of

CS molecular weight and estimation of EG-CS and PEG-CS molecular weight. Possible

problems associated with EG-CS or PEG-CS interaction with the column or solvent as

well as macromolecule aggregation were not observed in the respective chromatograms,

indicating EG-CS and PEG-CS likely flows through the selected column similarly to

natural CS. Similar studies of 𝑀𝜐���� for EG-CS and PEG-CS cannot be conducted because

K and α values have not been determined for these macromolecules.

6.4.5. Stability of amine and epoxide reactants

Page 219: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

193

In order to better understand the lack of high molecular weight polymer products

in EG-CS and PEG-CS samples, the reactive groups of the polymer step-growth polymer

system were investigated for their stability in the reaction conditions. The

fluoresecamine assay was used to investigate the stability of the CS terminal primary

amine in SBB buffer (pH 9.4) at 45̊C over a 96h period in order to mimic the synthesis

environment (CS concentration 25mg/ml). An L-serine standard curve was generated

and the number of primary amines on both CS and L-serine were determined over the

incubation period (Figure 6.29). The average number of primary amines per CS molecule

over the course of 96hrs did not fluctuate significantly (p > 0.05) and the number of

primary amines detected was not significantly different from L-serine (p > 0.05), which is

known to have one primary amine per molecule. It is therefore concluded that CS

primary amine stability is not compromised in the reaction medium and thus will not

hinder polymerization via step-growth synthesis.

Near infrared spectroscopy (near-IR) was used in order to monitor epoxide group

stability in the SBB aqueous environment (pH 9.4, 45̊ ) over a 96h period (similar to a

typical DGE-CS reaction). Distinct epoxide peaks can be resolved in near-IR spectra and

near-IR is often used as a tool to monitor epoxide-amine reaction kinetics (197, 208). We

were unable to use near-IR to monitor the reaction kinetics of CS-DGE reactions due to a

strong broad CS signal in the near-IR spectral region however, this technique can be used

to monitor epoxides of DGE monomers in the absence of CS. In order to investigate

DGE epoxide stability in the reaction conditions utilized in synthesis, PEG-DGE was

used as a model DGE. PEG-DGE samples were prepared at a 1M concentration in SBB

(0.1M, pH 9.4) and loaded into glass capillary tubes which were sealed on both ends

Page 220: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

194

before investigation. Near-IR spectra were collected at 64 scans with an SBB (0.1M, pH

9.4) background. Samples were placed in a water bath at 45̊C between investigations and

temperature of the near-IR sample holder was maintained at 45̊C during investigations.

Spectra were collected every 24h over a 96h period and monitored for changes associated

with the epoxide peaks of PEG-DGE (Figure 6.30).

In the near-IR spectra of PEG-DGE in SBB, peaks centered around 5850cm-1

corresponds to the CH stretching and bending of polyethylene glycol (209-211). This

peak will likely stay constant in the SBB environment since there is no indication for

hydrolysis of these bonds in these conditions. Peaks centered around 6094cm-1

correspond to the epoxide ring stretching vibrations (first overtones) (208). This peak is

not generally used for quantitative kinetic studies of epoxide ring opening due to

significant overlap with absorptions due to C-H and C-H2 stretching vibrations centered

around 6000cm-1 (208). An additional epoxide peak is found centered around 4542cm-1

and is assigned to epoxide ring stretching and bending vibrations. Absorption intensity of

this peak has been shown to decreases systematically during epoxide ring opening

reactions and is therefore often used to monitor epoxide rings in kinetics studies. Near-

IR peak at 4397 is assigned to OH absorption interactions and –OH stretching vibrations

can also be seen around peak wavenumber 7000cm-1. An increase in hydroxyl (OH)

absorptions may indicate hydrolysis of epoxide rings since ring opening of epoxides will

generate hydroxyls, but the OH peaks are susceptible to influence by hydrogen bonding.

Therefore the OH peaks of the near-IR spectrum cannot be used as a quantitative measure

of epoxide reactions. An analysis of near-IR peak height, ratioed to the PEG peak

indicates that there may be some changes in hydroxyl content and hydrogen bonding in

Page 221: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

195

PEG-DGE samples over time in SBB buffer. A drop is also seen in the 6094cm-1 epoxide

peak, however a concurrent decrease in the 4542cm-1 peak is not seen. The 4542cm-1

peak remains fairly steady indicating overall stability of epoxides in the reaction medium,

however, from changes to the OH and 6094cm-1 epoxide peaks, it is possible that some

level of non-specific epoxide ring opening may occur over the CS-DGE reaction period

reducing the reactivity of the DGE monomers. This effect however does not appear to be

significant due to the ability to react amines in DGE solutions over the course of the 96h

reaction (Figure 6.13) as well as the maintenance of the 4542cm-1 epoxide peak.

6.4.6. Reactivity of DGEs to the serine amine

Investigations into CS amine and DGE functional groups demonstrated their

stability in the reaction medium over the course of the reaction. The DGE-amine reaction

was therefore investigated for possible steric interference caused by bulky CS side chains

in order to better understand the limitations of the epoxy-CS-amine step growth

polymerization method. The serine molecule contains the same adjacent functionalities

to the amine as seen in the terminal primary amine of CS (Figure 6.32). Conjugation of

serine to the PEG based diglycidyl ethers should follow the same mechanisms as CS but

without the steric effects of the bulky and charged CS side chains. Reaction of EG-

DGE, PEG-DGE and PEG(1K)-DGE to serine (SBB buffer, pH 9.4, 45̊C), were therefore

monitored using the fluorescamine assay in order to investigate the steric contribution of

the CS chain in conjugation to DGEs (Figure 6.32). Serine-DGE reactions proceeded to a

higher % conjugation than CS at similar concentrations, however, conjugation did not

Page 222: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

196

proceed to completion. Although steric effects may play a part in the lower reactivity of

CS to DGE compared to serine, an overall low reactivity of DGE to the serine amine is

apparent. A further study was conducted for the conjugation of the simplest amino acid

to DGE in order to investigate possible reaction interference from the free OH of serine

which is not available in CS. Glycine was reacted to PEG- DGE similarly to CS and

serine, and % conjugation was not found to differ greatly from serine conjugation.

Therefore, an inherent low reactivity of the CS-primary amine to PEG based DGE

epoxide functional groups may contribute to the low polymerization seen in the step-

growth polymerization of CS and EG, PEG and PEG(1K)-DGEs.

Page 223: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

197

Figu

re 6

.11:

Kin

etic

s of C

S Pr

imar

y A

min

e Re

actio

n w

ith P

EG-D

GE

at v

aryi

ng T

empe

ratu

res a

nd D

GE

Con

cent

ratio

ns

with

One

-Pha

se A

ssoc

iatio

n C

urve

Fitt

ing

Page 224: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

198

Figu

re 6

.12:

Kin

etic

s of C

S Pr

imar

y A

min

e Re

actio

n w

ith E

G-D

GE

at v

aryi

ng T

empe

ratu

res a

nd D

GE

Con

cent

ratio

ns

with

One

-Pha

se A

ssoc

iatio

n C

urve

Fitt

ing

Page 225: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

199

1mM DGE Concentration

1 10 100

0

20

40

60

80

100EG (174g/mole)PEG (526g/mole)PEG 1K (1000g/mole)

Time (h)

% C

onju

gatio

n

10mM DGE Concentration

1 10 100

0

20

40

60

80

100EG (174g/mole)PEG (526g/mole)PEG 1K (1000g/mole)

Time (h)

% C

onju

gatio

n

20mM DGE Concentration

1 10 100

0

20

40

60

80

100EG (174g/mole)PEG (526g/mole)PEG 1K (1000g/mole)

Time (h)

% C

onju

gatio

n

100mM DGE Concentration

1 10 100

0

20

40

60

80

100EG (174g/mole)PEG (526g/mole)PEG 1K (1000g/mole)

Time (h)%

Con

juga

tion

Figure 6.13: Kinetics of CS Primary Amine Reaction to DGE monomers of varying poly(ethylene glycol) chain length

(Data fit to One-Phase Association Exponential Curves)

Best-fit values EG 1mM PEG 1mM PEG 1K 1mM Best-fit values EG 10mM PEG 10mM PEG 1K 10mMPlateau 12.26 9.428 19.45 Plateau 69.46 76.73 75.41

K 0.03917 0.03807 0.1425 K 0.06038 0.07561 1.298Tau 25.53 26.27 7.018 Tau 16.56 13.23 0.7705

R square 0.1781 0.1184 0.1899 R square 0.9375 0.9383 0.02821

Plateau 4.635 5.069 3.409 Plateau 2.06 1.899 1.669K 0.04935 0.06743 0.09338 K 0.006642 0.007056 0.4364

95% Confidence IntervalsPlateau 2.708 to 21.80 -1.014 to 19.87 12.42 to 26.47 Plateau 65.22 to 73.70 72.82 to 80.65 71.97 to 78.85

K 0.0 to 0.1408 0.0 to 0.1770 0.0 to 0.3349 K 0.04670 to 0.07407 0.06107 to 0.09014 0.3989 to 2.197Tau 7.101 to +infinity 5.650 to +infinity 2.986 to +infinity Tau 13.50 to 21.41 11.09 to 16.37 0.4552 to 2.507

Best-fit values EG 20mM PEG 20mM PEG 1K 20mM Best-fit values EG 100mM PEG 100mM PEG 1K 100mMPlateau 86.69 89.07 87.97 Plateau 96.82 97.18 96.84

K 0.1089 0.1328 0.8876 K 0.4903 0.5717 1.222Tau 9.185 7.529 1.127 Tau 2.04 1.749 0.8183

R square 0.9658 0.9356 0.3687 R square 0.9451 0.9281 0.6689

Plateau 1.36 1.654 1.174 Plateau 0.6327 0.5952 0.3837K 0.006409 0.009216 0.1091 K 0.01842 0.02232 0.06675

95% Confidence Intervals 95% Confidence IntervalsPlateau 83.89 to 89.49 85.66 to 92.47 85.55 to 90.38 Plateau 95.52 to 98.13 95.96 to 98.41 96.05 to 97.63

K 0.09567 to 0.1221 0.1138 to 0.1518 0.6628 to 1.112 K 0.4523 to 0.5282 0.5257 to 0.6176 1.085 to 1.360Tau 8.192 to 10.45 6.588 to 8.785 0.8990 to 1.509 Tau 1.893 to 2.211 1.619 to 1.902 0.7355 to 0.9220

Std. Error

One-phase association: Y=Plateau*(1-exp(-K*x))

One-phase association: Y=Plateau*(1-exp(-K*x)) One-phase association: Y=Plateau*(1-exp(-K*x))

One-phase association: Y=Plateau*(1-exp(-K*x))

Std. Error

95% Confidence Intervals

Std. Error

Std. Error

Page 226: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

200

Fi

gure

6.1

4: 1 H

-NM

Rmon

itorin

g of

Day

-by-

day

dial

ysis

of P

EG-D

GE-

CS

reac

tion

for p

urifi

catio

n of

Un-

reac

ted

PEG

-DG

E

Page 227: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

201

Figure 6.15: Epoxide chemical shift focus in 1H-NMR of Day-by-day dialysis of PEG-DGE-CS reaction purification

Page 228: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

202

Fi

gure

6.1

6: 1 H

-NM

Rmon

itorin

g of

Day

-by-

day

dial

ysis

of E

G-D

GE-

CS

reac

tion

for P

urifi

catio

n of

U

n-re

acte

d EG

-DG

E

Page 229: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

203

Figure 6.17: Epoxide chemical shift focus in 1H-NMR of Day-by-day dialysis of EG-DGE-CS reaction purification

Page 230: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

204

Figure 6.18: 1H-NMR analysis of PEG(1K)-CS purification via membrane dialysis with inset of epoxide region.

Page 231: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

205

Figu

re 6

.19:

1 H-N

MR

mon

itorin

g of

PEG

-CS

Synt

hesis

ove

r 96h

reac

tion

time

Page 232: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

206

Figu

re 6

.20:

1 H-N

MR

mon

itorin

g of

EG

-CS

Synt

hesi

s ove

r 96h

reac

tion

time

Page 233: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

207

Figu

re 6

.21:

Ana

lysi

s of 1 H

-NM

R p

eak

inte

grat

ions

for P

EG-C

S (to

p) a

nd E

G-C

S (b

otto

m) S

ynth

esis

Ove

r 96

h re

actio

n tim

e

Page 234: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

208

Figure 6.22: 1H-NMR of Purified EG-CS, PEG-CS and PEG(1K)-CS Synthesized at 20mM DGE Concentrations in SBB buffer (pH 9.4, 45C̊, 96h)

Page 235: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

209

Figu

re 6

.23:

Ana

lysi

s of 1 H

-NM

R p

eak

inte

grat

ions

for C

S m

acro

mol

ecul

es sy

nthe

size

d w

ith v

aryi

ng P

EG b

ackb

one

MW

Page 236: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

210

Figu

re 6

.24:

FTI

R S

pect

ra fo

r CS

mac

rom

olec

ules

synt

hesi

zed

with

var

ying

PEG

bac

kbon

e M

W

Page 237: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

211

Figure 6.25: Chromatograms of CS, EG-CS and PEG-CS in relation to polystyrene sulfonate standards

Page 238: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

212

Figure 6.26: Polystyrene Sulfonate Standard molecular weight calibration curve for JORDI DVB Sulfonated Mixed Bed column (250mm x 10mm (ID)) eluted with 85/10/5/0.1%

Water/Acetonitrile/Isopropyl Alcohol/Phosphoric Acid

Page 239: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

213

Figure 6.27: Example of molecular weight distribution curve and graphical representation of Mn, Mp, Mw, and Mz

Page 240: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

214

Concentration Extrapolation ofCS Intrinsic Viscosity

0.000 0.005 0.010 0.015 0.02030

40

50

60

70

y = 1454x + 38.64R2 = 0.94

Concentration (g/ml)

Red

uced

Spe

cific

Vis

cosi

ty( η

sp/c

) (m

l/g)

Figure 6.28: Concentration extrapolation of CS Intrinsic Viscosity

Page 241: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

215

Figure 6.29: Fluorescamine analysis of CS primary amine stability over time in pH 9.4, 45C̊ buffer conditions

Page 242: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

216

Figu

re 6

.30:

Nea

r-IR

Spe

ctra

of 1

M P

EG-D

GE

epox

ide

stab

ility

in S

BB

pH

9.4

buf

fer,

45C̊

ove

r 96h

Page 243: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

217

Figu

re 6

.31:

Nea

r-IR

Pea

k ra

tio a

naly

sis o

f PEG

-DG

E (1

M c

once

ntra

tion)

in S

BB

(pH

9.4

, 45C̊

) ove

r 96h

.

Page 244: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

218

Figu

re 6

.32:

Flu

ores

cam

ine

inve

stig

atio

n of

Ser

ine

and

Gly

cine

ver

sus C

S co

njug

atio

n to

DG

E

Page 245: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

219

Table 6.3: Summary of (A) Rate Constants and (B) % Conjugation Plateaus for CS Primary Amine Reaction to DGE of Varying MW

A EG-DGE PEG-DGE PEG (1K)-DGE

DGE Concentration

K (1/h) 95% CI K (1/h) 95% CI K (1/h) 95% CI

1mM 0.03917 0.0 to 0.1408 0.03807 0.0 to 0.1770 0.1425 0.0 to 0.3349 10mM 0.06038 0.0467 to 0.0741 0.07561 0.0611 to 0.0901 1.298 0.3989 to 2.197 20mM 0.1089 0.09567 to 0.1221 0.1328 0.1138 to 0.1518 0.8876 0.6628 to 1.112

100mM 0.4903 0.4523 to 0.5282 0.5717 0.5257 to 0.6176 1.222 1.085 to 1.360

B EG-DGE PEG-DGE PEG (1K)-DGE

DGE Concentration

Plateau (% Conj)

95% CI Plateau (% Conj)

95% CI Plateau (% Conj)

95% CI

1mM 12.26 2.708 to 21.80 9.428 -1.014 to 19.87 19.45 12.42 to 26.47 10mM 69.46 65.22 to 73.70 76.73 72.82 to 80.65 75.41 71.97 to 78.85 20mM 86.69 83.89 to 89.49 89.07 85.66 to 92.47 87.97 85.55 to 90.38

100mM 96.82 95.52 to 98.13 97.18 95.96 to 98.41 96.84 96.05 to 97.63

Table 6.4: Table of GPC Sample Loading Parameters

Sample Name Concentration (mg/ml)

Injection Volume (μl)

PAA-CS 2 200 EG-CS 2 50

PEG-CS 2 50 Chondroitin Sulfate 2.5 50

Poly(acrylic acid) 6.7 100

Page 246: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

220

Table 6.5: Molecular Weight distributions calculated relative to polysaccharide standards

Sample Run # Mn Average Mw Average Mz Average Mw/Mn Average

CS 1 223,604

225,941 294,116

295,264 398,977

399,448 1.32

1.31 2 228,278 296,411 399,918 1.30

PAA-CS 1 18,604

18,596 150,817

149,092 388,997

383,587 8.11

8.02 2 18,588 147,367 378,177 7.93

EG-CS 1 290,972

289,574 372,124

372,358 480,044

481,201 1.28

1.29 2 288,176 372,592 482,358 1.29

PEG-CS 1 280,083

280,938 376,317

373,379 497,440

491,097 1.34

1.33 2 281,793 370,441 484,754 1.31

* Relative to Polysaccaride Standards

Table 6.6: Molecular weight distributions calculated relative to polystyrene sulfonate standards

Sample Run # Mn Average Mw Average Mz Average Mw/Mn Average

CS 1 12,535

12,359 29,260

29,113 61,919

62,227 2.33

2.36 2 12,184 28,966 62,536 2.38

EG-CS 1 18,526

18,508 39,022

38,916 73,034

73,327 2.11

2.10 2 18,490 38,810 73,620 2.10

PEG-CS 1 19,135

19,363 42,599

42,464 78,924

78,627 2.23

2.19 2 12,591 42,328 78,330 2.16

* Relative to Polystyrene Sulfonate Standards

Page 247: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

221

In this study, the step-growth grafting-through synthesis of CS-brush

macromolecules was investigated using the epoxy-amine reaction. CS with a terminal

primary amine was reacted to diepoxide monomers; ethylene glycols of varying

molecular weight terminated on both chain ends with epoxide groups: EG-DGE, PEG-

DGE, and PEG(1K)-DGE. EG-DGE and PEG-DGE were chosen due to the small

molecular spacing between epoxide groups (0.5nm and 5nm respectively). A molecular

spacing between CS chains of less than 4nm has been predicted to be required in order to

elicit electrostatic contributions the tissue mechanical properties (69). PEG(1K) was

investigated in order to better understand the effects of DGE MW on CS-macromolecule

synthesis.

6.5 Discussion

6.5.1. Reaction kinetics of the CS-DGE reaction

In step-growth polymerization of an epoxy-amine system, epoxides first react to

an available primary amine, generating a secondary amine. A second epoxide can then

react to the generated secondary amine in order to create a tertiary amine linkage. These

reactions can be applied to the CS, diglycidyl ether system as seen in Figure 6.33 to

generate linear polymers with ethylene/polyethylene backbones and pendant chondroitin

sulfate bristles. Epoxide primary amine, and secondary amine reactions are necessary in

order to build linear chains with several CS pendant groups as illustrated in Figure 6.33.

Several parameters were investigated for their influence on the CS primary amine

reaction with DGEs (Represented by reactions 1 and 2 in Figure 6.33). Temperature,

DGE concentration and DGE molecular weight were all found to modulate the reaction

Page 248: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

222

kinetics of the CS primary amine-epoxide reaction. As expected, an increase in

temperature increased the epoxy-primary amine reaction rate (174, 212). Temperature is

often used to increase conversion and reaction rates in epoxy polymer systems. Care

must be taken however at high temperatures (>147 ̊C) in order to avoid possible side

reactions such as etherification (174, 179, 181). In our system, it is especially important

to keep temperature reasonable in order to avoid the degradation of the CS chains. CS is

generally stable up to 200h at temperatures up to 60̊C, while longer incubation times

leads to degradation (92). Basic conditions also make CS more susceptible to degradation

at elevated temperatures when compared to neutral conditions (92). Therefore,

temperatures above 45̊C were not investigated for their potential impact on CS primary

amine reaction rate and reaction times were generally kept below 200h.

Interestingly, temperature did not play a significant role in epoxide-primary amine

reaction rate when DGE monomer concentrations were high ( > 100mM DGE). At high

DGE concentrations, it appears that the monomer concentration outweighs any

thermodynamic benefits gained by increased temperature reactions. Although increasing

temperature was found to be effective in increasing primay amine conjugation at

moderate DGE concentrations (< 100mM) (Figure 6.11, Figure 6.12), even at elevated

temperatures (45̊C), an excess of epoxide monomer (>10mM where CS concentration is

1.13mM) was required in order to achieve high levels of primary amine conversion

(>80% Conjugation) (Figure 6.13).

PEG-DGE molecular weight also had a significant effect on primary amine-

epoxide reaction rate with reaction rate increasing with DGE monomer molecular weight

(Figure 6.13 and Table 6.3). This dependence on molecular weight may be associated with

Page 249: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

223

the unique properties of the PEG spacing unit. In addition to the chemically stable, inert

and biocompatible natures of PEG (213-215) which make it desirable for biomedical

applications, PEG has many unique solution properties and is available in a range of

molecular weights (from oligomers to molecular weights of 7 million) where properties

of PEG tend to vary considerably with molecular weight (216). In interactions of PEG

with water and other solutes in aqueous conditions, several considerations must be made.

Attractive interactions due to the hydrogen-bond formations between ether oxygen of the

PEG chain and water, water-water interactions, and repulsive interaction between water

and the alkyl groups of PEG (hydrophobic hydration) may occur in addition to hydrogen

bonding and hydrophobic attractive interactions between PEG and other solutes (i.e.

amino acids, and in our case chondroitin sulfate) (217-219). Attractive interactions

caused by hydrogen bonding between the ether oxygen of the PEG chains and the

electrophilic groups of CS as well as hydrophobic interactions between alkyl groups of

CS and PEG can increase the entanglements between CS and PEG bring the amine and

epoxide groups in closer proximity, thereby affecting reaction rate. Increasing the PEG

MW may increase these entanglements. It is known that with increased molecular

weight, PEG itself shows increased entanglement (aggregation) (216) and increased

water association (water molecules bound per ethylene glycol segment) (213, 220) . PEG

also demonstrates drag reduction (221) and steric stabilization (222) properties which

increase with increasing PEG MW, and may also contributing to the enhanced

conjugation with increasing PEG segment MW. Reactivity of the epoxide groups and

chemistry of the backbone themselves were not expected to play a role in the MW

Page 250: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

224

dependence of reaction kinetics since the chemistry of the DGEs were similar, only

varying in the number of ethylene glycol units.

Reaction kinetics of the secondary amine (formed in the reaction of the CS

primary amine with epoxide (Figure 6.33)) with epoxide could not be monitored in our

given conditions. Epoxy-amine primary and secondary reaction kinetics are usually

investigated with a combination of differential scanning calorimetry(DSC) and Near-IR

(197, 208, 223). These investigations however are not applicable to our system due to the

majority composition of CS within our samples. The CS signal dominates the near-IR

spectra as well as the thermal properties obtained in DSC.

Direct analysis of amines is most routinely used to quantify primary amines (145,

224). Some methods have been developed to monitor secondary and tertiary amines

however they are generally more sensitive to external factors such as water, other amine

and functional group populations and steric hindrances of the amines themselves (225-

230). In addition, available methods often require a good deal of sample processing as

well as complex titrations in order to determine the quantity of the desired amine. These

methods are also sensitive to the type of amine (most assay development is conducted on

aromatic amines) as well as the concentration of amines (methods are not highly sensitive

to trace concentrations of amine). These limitations in secondary amine and tertiary

amine direct monitoring has often limited these techniques to the analysis of amines only

in the specific settings in which the assay was developed (assays have not become

commercially available, and papers on secondary and tertiary amine detection method

development have not been highly cited since their publication in the 1940’s, 1950’s,

1960’s 1970’s and 1980’s) (225-231). The analysis of secondary amine reactions in our

Page 251: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

225

system of chondroitin sulfate and epoxide reaction in aqueous conditions was therefore

not feasible and our analysis of reaction kinetics was focused on the initial primary amine

reaction with epoxide. The determination of reaction products size and chemical

composition however give insight to the reactions taking place within our system and

serve as a secondary means of investigating amine-epoxide reactions.

6.5.2. Purification of synthesized CS-Macromolecules

Purification of the epoxide-CS reaction mixture was accomplished by extensive

membrane dialysis against DI water. In membrane dialysis, the sample is contained in

dialysis tubing of a particular molecular weight cutoff. The MWCO is defined as the

molecular weight of the solute that is 90% retained by the membrane during a 17 hour

period. Molecules that are larger than the MWCO are theoretically retained in the

dialysis membrane while the driving force of diffusion forces smaller solutes out of the

membrane tubing. As a rule of thumb, in order to separate molecules via membrane

dialysis, there must be at least a 5x difference in molecular weight of the species being

separated and MWCO is generally chosen such that it is ½ the MWCO of the molecules

that needs to be retained. In the current case, CS has an approximate MW of 22K while

the largest DGE has a MW of approximately 1K making dialysis an appropriate choice

for purification. In order to purify CS-macromolecule samples, dialysis was conducted

within a 6K-8K MWCO regenerated cellulose membrane against DI water. DI was

chosen in order to fully purify CS-macromolecules from excess salts introduced in the

sysnthesis process. By dialyzing against DI, resulting products will have only the salts

associated with the CS molecule and resulting yields are not skewed by added salts in the

system. Dialysis was monitored daily in order to determine minimal time required to

Page 252: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

226

purify un-reacted DGE monomers. After 48h of dialysis both PEG-CS (Figure 6.14) and

EG-CS (Figure 6.16) compositions showed no DGE contaminant peaks as well as no

epoxide peaks (Figure 6.15 and Figure 6.17). PEG(1K)-CS samples showed trace amounts

of epoxide and likely require 72h of dialysis (Figure 6.18).

The dialysis process generates a pressure within the dialysis membrane due to the

large salt concentration associated to the CS molecules and the lack of salts in the DI

water. This pressure swells the dialysis tubing and can sometimes cause it to burst,

therefore, it is important to only partially fill dialysis tubes, to allow room for DI water o

enter the tube and dilute the sample. An additional consequence of the introduced

swelling pressure is opening of the pores of the dialysis membrane. The dialysis

membrane used, is only rated for pressures up to 1.5psi before MWCO is compromised

(232). Osmotic pressure of CS solutions at 25mg/ml (the concentration used in synthesis

and purification) is approximately 11psi in 0.15M NaCl and expected to be higher in zero

salt conditions (233). Therefore it is not unlikely that the actual MWCO of the dialysis

membranes was greater than 6K-8K MWCO. Monitoring the yield of samples after

different times of dialysis, it was noticed that significant material loss occurs over the

dialysis period (Figure 6.34). It is therefore important to minimize amount of the dialysis

used to maximize product yield. Smaller MWCO membranes may also be used, but

longer dialysis times may be required to complete dialysis. Dialysis against osmotically

conditioned solutions of high molecular weight polyethylene glycol or dextrans may also

be utilized in order to balance the internal and external osmotic pressures of the dialysis

bags.

Page 253: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

227

Dialysis against DI water also has the positive consequence of hydrolysis of

remaining epoxides due to the slightly acidic nature of DI water (pH ~5) (Figure 6.35).

This is important for biocompatibility issues, where epoxides are highly reactive

functional groups and may interfere with normal cellular function (234). As a result of

epoxide hydrolysis however, care must be taken to monitor not only the epoxide 1H-

NMR peaks since these peaks may disappear before full purification is reached. A

drawback of this purification technique however is that any epoxides that may have been

remaining at the terminal end of CS as a result of the synthesis procedure would not be

detected because of hydrolysis. If dialysis is conducted against SBB (0.1M, pH 9.4),

epoxides may be preserved and detected (Figure 6.30), but the low concentrations of

epoxides expected in CS-macromolecule samples may make them difficult to detect over

signal noise. Another methods that may be used to determine the end functionality of

synthesized CS-macromolecule chains is matrix-assisted laser desorption/ionization time-

of-flight mass spectrometry (MALDI-TOF MS) (235).

6.5.3. Chemical structure of the CS-Macromolecules

The chemical structure of CS was monitored over the synthesis process in order to

investigate the reaction of CS to DGE. EG-CS, PEG-CS and PEG(1K)-CS 1H-NMR

spectra matched well with that of un-reacted CS (CS was processed similarly to reacted

CS but not exposed to DGE). For EG-CS and PEG-CS monitored over the course of

synthesis, all peaks corresponding to CS maintained their spectral shifts (Figure 6.19 and

Figure 6.20) and had only small variations in integrated area (Figure 6.21). The only trend

in change in integrated are was noticed for PEG-DGE samples at the PEG spectral shift

location (3.66ppm) which overlaps with a broad CS band between 3.58 and 3.86ppm.

Page 254: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

228

Due to this overlap, the actual amount of PEG incorporated into the CS molecule cannot

be quantified. It is interesting to note however, the progressive increase in this peak over

the reaction time of PEG-DGE to CS. Fluorescamine analysis of the reaction rate of

PEG-DGE with CS at 20mM DGE concentration indicates a time constant of primary

amine reaction of between 6.58h and 8.78h. This indicates that the system is expected to

reach approximately 95% of its maximum conjugation within 24h. 1H-NMR results

suggest however that addition of PEG to the CS molecule (20mM DGE reaction)

proceeds after the initial 24h period. This may be due to reactions at the secondary

amine. Secondary amines generally react with epoxides at a slower rate than the original

primary amine due to a negative substitution effect (180-181). The secondary amine

hydrogen is more sterically hindered than its predecessor primary amino hydrogen

reducing the rate of the reaction. It is therefore necessary to allow epoxy-amine systems

to react over a relatively long period of time to first allow the primary amines to react and

allow more time to allow the secondary amines to react. This phenomenon was seen in

our current PEG-DGE-CS reaction system where primary amine reactions proceeded

within the first 24h but additional incorporation of monomer was seen even after 96h of

reaction.

It is possible that side reactions could have occurred in this system with other

epoxide reactive groups. The most likely side reaction would occur at the GalNAc

primary hydroxyl (primary hydroxyls are more available for reaction than secondary

hydroxyls), however this signal is combined in the 3.58 and 3.86ppm band and would be

obscured by the PEG signal at 3.66ppm. PEG-CS, EG-CS, and PEG(1K)-CS samples,

synthesized for 96h with 20mM DGE, were investigated with ATR-FTIR to further

Page 255: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

229

investigate any side possible side reactions, especially located at the hydroxyl groups. In

particular, in a study conducted to fabricate a hydrogel from CS and PEG-DGE, CS was

reacted with PEG-DGE at high pH (1M NaOH, pH~14) to induce epoxide reaction with

hydroxyls to crosslink the CS network. Changes in FTIR spectra were seen with a shift

of the C-O stretching (236) from 1035cm-1 to 1100cm-1 (237). It has also been

demonstrated that in the C-O stretching mode, shifting occurs with changes to the

hydroxyl group orientation (156). Therefore, this peak was monitored for shifts in the

ATR-FTIR spectra of EG-DGE, PEG-DGE and PEG(1K)-DGE (Figure 6.24). No major

peak shifts were detected at the 1032cm-1 peak s of CS-macromolecules indicating that

there was no significant bonding of the CS hydroxyls by the DGEs investigated. 1H-

NMR analysis (Figure 6.22 and Figure 6.23) and ATR-FTIR analysis (Figure 6.24) of final

20mM DGE formulations of EG-CS, PEG-CS and PEG(1K)-CS indicate no significant

reactions of DGE with the CS molecule other than the incorporation of PEG groups onto

the CS chain. The incorporation is likely due to primary amine reactions which were

monitored with the fluorescamine assay. Secondary amine reactions may also be

occurring as evidenced by the continued incorporation of PEG into PEG-DGE samples,

even after primary amine reactions have reached a plateau.

6.5.4. CS-Macrmolecule molecular weight distribution

EG-CS and PEG-CS molecular weight distributions were estimated with GPC.

Method development for GPC was contracted from Jordi Labs due to the complex nature

of SEC of hybrid molecules. Stationary phase column material, mobile phase solvents

and injection procedures were optimized for the elution of CS and CS-macomolecules.

Initial runs with polysaccharide standards resulted in highly over estimated molecular

Page 256: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

230

weight for both CS and CS-macromolecules (Table 6.5) however additional trials with

polystyrene sulfonated standards resulted in more accurate MWD determinations (Table

6.6). GPC measurement of molecular weight values was validated with viscosity average

molecular weight measurements of CS. CS Mw was near the expected value of 22K for

CS-4 and Mz values for CS were consistent with contamination with larger MW CS-6

(65, 139). Molecular weight increased with incorporation of PEG based synthetic

backbones of increasing MW (i.e. PEG-CS > EG-CS > CS). Based on these MWD for

CS and the CS-macromolecules, it is likely that the EG-CS and PEG-CS samples

contained a mix of dimmers (two linked CS chains) and monomers (individual CS chains

or CS chains linked to one or two DGE monomers) as illustrated in products A, B, C, and

D of Figure 6.33.

In addition to GPC quantification of molecular weight, gel electrophoresis and

dynamic light scattering techniques may be used to obtain molecular size information.

Both gel electrophoresis (96, 238-239) and dynamic light scattering have been used to

establish polysaccharide and proteoglycan molecular size (240-241) however their use in

characterizing heterogeneous synthetic polymer brushes is limited (106). Conventional

techniques for the determination of molecular weight and size distribution (i.e. GPC, light

scattering, viscometry and electrophoresis) are often limited when applied to complex

macromolecules that involve branching and heterogeneous composition (106).

Therefore, visualization of individual macromolecules is often utilized to characterize

brush polymer molecular weight distribution and is recommended as a secondary method

for the determination of CS- macromolecule size and structure (106).

Page 257: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

231

In healthy discs, aggrecan macromolecules can contain approximately 100 CS

chains and 30 KS chains resulting in aggrecan molecular masses of around 2,000kDa

(242). In degenerating NP tissues, enzymatic degradation of aggrecan by MMPs and

Aggrecanases results in the fragmentation of the aggrecan macromolecule into smaller

subunits. Aggrecan fragments in the range of 100kDa to as low as 55kDa have been

found in NP tissues at both early (Grade 2) and advanced (Grade 4) stages of

degeneration (78, 243). This indicates that aggrecan molecules of these sizes can be

retained in the NP matrix. Therefore, in the synthesis of biomimetic aggrecan

macromolecules for injection and retention in the NP, macromolecules sizes of greater

than 55kDa may be applied. The synthesis strategy utilized in the current study resulted

in CS-macromolecules of MW ranging from 18,500Da (Mn of EG-CS) to 78,600Da (Mz

of PEG-CS). The larger of these molecules may have extended residence time in NP

tissues; however, they are only a small portion of the synthesized population of

macromolecules. The Average Mw of the EG-CS and PEG-CS macromolecules were

approximately 38,900Da and 42,500Da respectively, indicating that the majority of the

synthesized macromolecules would be too small to remain in NP tissue over time. In a

separate study, injections of CS and injections of a water control were made to a human

cadaveric intervertebral disc. After cycling through a diurnal cycle, there was no

difference in the CS disc from the water control or from the intact condition (data not

shown).

In order to investigate possible causes for the low molecular weight polymers

obtained, a series of studies were conducted. Both the primary amine group of CS and

the epoxide groups of the DGE monomers were investigated for their stability in the

Page 258: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

232

synthesis reaction conditions. Both groups were found to be stable over a 96h period as

determined by the fluorescamine assay for amines (Figure 6.29) and near-IR for epoxides

(Figure 6.30 and Figure 6.31). Finally, the contribution of steric hindrances to the

conjugation of long CS chains to the PEG backbones was investigated by using L-serine

(the amino acid residue of the CS terminal primary amine) for conjugation as a

comparison. L-serine molecules are small and do not have the steric hinderence

associated with the charges and long chains of CS. Conjugation was seen to be slightly

increased for L-serine samples indicating that steric hinderences may be playing in a

small part in reduced attachment of CS to DGEs. Conjugation however did not proceed

to 100% when equimolar and twice equimolar concentrations of amine were combined

with epoxide (Figure 6.32). This can introduce a large problem in step-growth

polymerization, where each un-reactive group will result in dead chains ends, reducing

the final molecular weight of the polymer.

In the step-growth polymerization utilized in this synthesis procedure, primary

amines, secondary amines and epoxides all need to be highly reactive in order to promote

chain growth and the synthesis of large molecular weight species. Epoxide-amine

reactions can be influenced by several factors including, amine structure, epoxide

structure, solvent nature, and side reactions. Reaction rate generally increases with the

basicity of the reacting amine (183). In our system however, we are limited to the

inherent properties of the serine residue as our primary amine and its reaction with an

epoxide for our secondary amine. Epoxide structure does not have as large of an effect

on epoxy-amine reactions, however epoxy compounds with aromatic rings in the

backbone have been known to react more readily. The addition of aromatic rings in a

Page 259: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

233

molecule however generally makes it more hydrophobic and will therefore be difficult to

solubilize in the CS aqueous medium.

Given these considerations of the epoxy-amine reactive system and the serine

studies conducted indicating that the EG-DGE, PEG-DGE and PEG(1K)-DGE reactions

to serine amino acid residue may be somewhat inhibited, a large barrier to the use of this

step-growth strategy in the synthesis of high molecular weight CS-macromolecule

species. More reactive epoxides may be investigated, and it is possible, that intermediary

steps, which push the epoxy-amine reactions to completion, may be used to build

polymers in a more step-wise fashion. These techniques will require long synthesis times

and careful purification in order to result in high molecular weight polymers.

Page 260: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

234

Figure 6.33: Primary and secondary amine reactions of CS with PEG-DGE

See Appendix 2 for alternate view of amino acid terminated CS

Page 261: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

235

Percent Yield Over Dialysis Timeof Synthsized CS-Macrmolecules

0h (S

ynthes

is)

24h D

ialys

is

48h D

ialys

is

72h D

ialys

is

96h D

ialys

is0

20

40

60

80

100

Perc

ent Y

ield

(%)

Figure 6.34: % Yeild of CS-Macromolecules (combined results for EG-CS and PEG-CS) over the course of dialysis (starting CS weight of 250mg)

Page 262: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

236

Figure 6.35: near-IR analysis of PEG(1K)-DGE in DI water over 96h. Epoxides peaks highlighted in Blue. PEG peak highlighted in purple. Epoxide Peaks decrease

with increased time in DI water.

Page 263: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

237

The step-growth polymerization of CS with PEG based diepoxides was

investigated in this study for the synthesis brush-like CS macromolecules. The CS DGE

reaction was demonstrated to be localized primarily to the CS terminal primary amine

and polymer products were isolated from the reaction with no residual DGE monomer.

Investigations into the molecular weight distribution of the synthesized CS-

macromolecules indicated that the reaction products consisted primarily of CS monomer

and dimmers. The lack of extended monomer addition is likely due to low reactivity of

the serine residue terminal amine to the DGE epoxide functionalities. A reduction in

functional group reactivity can greatly decrease the molecular weights of the polymer

products. Although large brush molecules were not likely synthesized, CS-

Macromolecules were synthesized which incorporate a synthetic PEG polymeric

component generating novel CS macromolecular hybrid structures. These structures will

be further investigated in order to understand the impact of the hybrid nature of these

macromolecules on functional properties such as viscosity, osmotic pressure, and

cytotoxicity. In addition, the techniques of reaction analysis, chemical characterization,

and purification developed in this study will be valuable to the development and

evaluation of future biomimetic aggrecan synthesis strategies.

6.6 Conclusions

Page 264: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

238

Chapter 7 : Characterization of CS Macromolecules synthesized via the Epoxy-Amine Step-Growth “Grafting-Through” technique of Polymer Synthesis

The human intervertebral disc is the primary compression-carrying component of

the spine. Its role is to transmit and distribute loads, and allow for the necessary

flexibility of the spine. During a diurnal cycle, the intervertebral disc experiences

approximately 16 hours of functional loading (standing, sitting, etc.), followed by 8 hours

of recovery (lying prone). Therefore, the fluid lost during the loading period must be

replenished in half the time that it is expelled. As the disc is compressed and fluid is

exuded, the density of the fixed charges within the nucleus pulposus is increased, creating

an osmotic gradient with the interstitial fluid surrounding the disc (4). This osmotic

potential aids in drawing fluid back into the disc. Additionally, the chemical and

structural arrangements of key biomolecules in the nucleus pulposus provide mechanical

resistance to compressive loads. In particular, the large bottle brush proteoglycan,

aggrecan, has two major modes of function in providing mechanical support in the NP:

1) it allows water uptake by the nucleus due to sulfated groups on the chondroitin and

keratin sulfate regions which, in part, provide intradiscal pressure and 2) it provides

electrostatic repulsion due to the 3D macromolecular structure, which contributes to

intradiscal pressure and disc height. Specifically, electrostatic forces between closely

arranged CS chains have been theoretically suggested to account for 50% (290kPa) of the

equilibrium compressive elastic modulus in disc like tissues (70, 72).

7.1 Introduction

Page 265: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

239

Alterations in the major biochemical constituents of the intervertebral disc

coincide with aging and disc degeneration, and can subsequently alter the ability of the

disc to support load (20, 244-246). While disc degeneration is poorly understood, it has

been shown that the proteoglycan content of the inner, nucleus pulposus region of the

disc, decreases linearly with age and degeneration with almost 40% of proteoglycans lost

by the age of 40 (3, 24). This loss of proteoglycans leads to a loss of hydration of the

nucleus manifested by a reduction of disc height and intradiscal pressure. Proteoglycans

work to resist mechanical forces in the nucleus and provide a hydrostatic pressure to the

outer annulus fibrosus. In a dehydrated disc, the function of the nucleus, namely load

transfer to the annulus through creation of an intradiscal pressure is hindered and the

mechanics of the degenerated disc are altered compared to those of a healthy disc (29).

Prior work has investigated the role of the nucleus pulposus in human lumbar

intervertebral disc mechanics and studies show that the nucleus is critical in providing

stability to the intervertebral disc (2, 247).

In order to mechanically stabilize the disc early in the degenerative cascade and

prevent the need for spinal fusion surgeries, we have proposed the development of a

hybrid-bio/synthetic biomimetic proteoglycan macromolecule for injection into the

nucleus pulposus of the intervertebral disc in the early stages of degeneration, after a first

or second episode of back pain. The biomimetic proteoglycan is based on the major disc

proteoglycan aggrecan which is a bottle brush macromolecule with a protein core and

radiating negatively charged chondroitin sulfate (CS) bristles. In the biomimetic

approach, the protein core is replaced with a synthetic polymeric backbone in order to

resist enzymatic and hydrolytic degradation within the nucleus environment while

Page 266: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

240

maintaining the osmotically active chondroitin sulfate bristles in their natural bottle brush

arrangement.

While administrations of natural aggrecan may be useful, the cost of the material

at this point in time is prohibitive for any type of realistic intervention (A1960: Aggrecan

from bovine articular cartilage, $225USD/mg from Sigma(248)). In addition,

commercially available aggrecan would be subject to the same limitations as the body’s

own aggrecan, enzymatic degradation of the protein core, which fragments the molecule

and allows for migration of the fragments from the intradiscal space (18, 23).

The function of biomimetic aggrecan in the restoration of NP tissue hydration and

mechanical properties is several fold (Table 4.1). The biomimetic aggrecan must be of

the appropriate size to maintain residence in the NP tissue matrix (generally greater than

70kDa (78, 243)) and the structure should be similar to that of native aggrecan, with

charged groups maintained in close proximity (2-4nm spacing between adjacent CS

molecules (69)). In addition, the biomimetic aggrecan should be maintained in a fluid

aqueous matrix in order to ensure integration with the surrounding NP tissue matrix (i.e.

not a cross-linked network). The biomimetic aggrecan also must be able to remain stable

in the harsh hydrolytic and enzymatically active NP environment. Finally, since the

biomimetic aggrecan will be integrated into existing NP tissue with a cellular population

of nucleus pulposus cells (approximately 2,500cells/mm3 (249)), it is important that

biomimetic aggrecan be biocompatible with minimal cytotoxic effects on NP cells.

In the previous chapter, biomimetic aggrecan in the form of CS-macromolecules

with poly(ethylene glycol) based synthetic backbones was synthesized via a step-growth

Page 267: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

241

“grafting-through” strategy. The PEG based backbones are inherently hydrolytically and

enzymatically stable (215, 250), thereby rendering the CS-macromolecules resistant to

degradation in the the acidic and enzymatically active NP environment (aggrecanases and

MMPs in the NP breakdown native aggrecan at specific amino acid sequences of the

aggrecan core protein (79)). PEG backbone segment length (i.e. spacing between

adjacent CS molecules) was also chosen in the range of 0.5 to 5nm (EG-CS and PEG-CS

formulations respectively) in order to mimic the close packing of CS in aggrecan.

Investigations into the molecular size of the synthesized CS-macromolecules indicated

the formation of some larger macromolecular species in the range of 70kDa (in the range

of aggrecan fragments found in native NP tissue), however the Mw of EG-CS and PEG-

CS molecules were low, approximately 40kDa (i.e. dimmers). Therefore, future work

was recommended in order to increase the CS-macromolecule molecular mass.

The current formulations of CS-macromolecules are novel in their incorporation

of a synthetic backbone between CS-segments. Therefore, in the current study these CS-

macromolecules were investigated for the influence of the synthetic backbone on

macromolecule function with regards to key functional requirements of biomimetic

aggrecan. CS-macromolecules were characterized for their physical structure using

transmission electron microscopy (TEM) and viscosity measurements. The osmotic

functions of the macromolecules were investigated by monitoring their fixed charge

density as well as their osmotic potential. Finally, CS-macromolecule formulations were

investigated for their ctyocompatibility.

Page 268: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

242

7.2 Methods

EG-CS and PEG-CS prepared at 20mM DGE and 1.13mM CS concentration

formulations (SBB Buffer, pH 9.4, 45̊C, 96h synthesis) were utilized in these studies.

PEG(1K)-CS (20mM formulation) was also investigated in some instances as a point of

comparison, however it was not extensively investigated due to its long (~12nm) and

non-physiologically relevant segment length between CS chains. All plots were

generated in GraphPad Prism v5.02 and statistical analysis done in GraphPad Prism

v5.02.

7.2.1. Transmission Electron Microscopy (TEM) Imaging of PEG-CS Macromolecules

The structure of the CS-macromolecules was investigated via TEM imaging. PEG-

CS samples were prepared as previously described and purified via gravity column

filtration (GE PD-10 Desalting Columns). Desalting removes the majority of PEG-DGE

monomers however trace monomer contamination may remain (data not shown).

Samples were lyophilized, and then reconstituted in distilled deionized water at 1mg/ml

concentrations. A 10µL volume of PEG-CS solution was pipetted onto a 20-30nm

carbon coated, 400 mesh copper grid (Electron Microscopy Science) and allowed to

incubate at room temperature (covered to avoid evaporation) for 1h. Samples were then

stained with uranyl acetate solution (10µl of 1mM in distilled deionized water, Electron

Microscopy Science) and rinsed three times with fresh distilled deionized water. Images

were collected on a JEOL JEM2100 Transmission Electron Microscope managed by the

Page 269: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

243

Centralized Research Facilities of the Drexel University College of Engineering. Images

were kindly taken by trained super user Christopher Winkler of the Dynamic

Characterization Group in the Department of Materials Science and Engineering at

Drexel University.

7.2.2. Viscosity of CS-macromolecule solutions

Viscosity ( η, pa.s) is the resistance of a fluid to flow and is dependent on the

structure and properties of polymer fluids where:

𝜂 =𝜎21�̇�

and η is the dynamic viscosity, σ21 is shear stress and γ is shear rate. The effect of a

dissolved solute on a solution is given by the relationships of relative viscosity, ηrel, and

specific viscosity ηsp, where

𝜂𝑟𝑒𝑙 = 𝜂𝜂𝑜

𝑎𝑛𝑑 𝜂𝑠𝑝 = 𝜂𝑟𝑒𝑙 − 1

and η is the viscosity of the solution, ηo is the viscosity of the solvent and. Both relative

viscosity and specific viscosity are unit-less.

Viscosity of CS-macromolecule solutions was determined on an AR2000ex rheometer

(TA Instruments) with cone-on-plate configuration (1̊ Cone angle). Samples were

brought into solution at 25mg/ml in 1X PBS buffer and 500µl pipette onto the peltier

plate surface. Temperature was controlled at 25̊C and gap distance was 32µm. Shear rate

was logarithmically ramped from 10/s to 500/s at 10 points per decade, resulting in the

Page 270: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

244

collection of a total of 18pts. Specific viscosity was calculated relative to PBS buffer

control samples, and the average steady state specific viscosity calculated (n=3).

7.2.3. Fixed Charge Density of CS-Macromolecules

Fixed charge density (FCD) in the IVD is a measure of the negative charges/unit

volume attached to the disc matrix (251). Electrostatic interactions between fixed

charges and mobile free ions result in the generation of physiochemical and electro-

kinetic effects including Donnan osmotic and swelling pressures in the tissue (252).

Fluid and solute transport as well as swelling behavior in the IVD are greatly affected by

the fixed charged density in the tissue, thus the FCD of the synthesized CS-

macromolecule is an important parameter to maintain in a biomimetic approach (84,

244).

The Fixed charge density of a CS solution can be estimated from the GAG

content of the solution using the relationship between the molecular weight of the CS

disaccharide and the quantity of charge on each disaccharide unit. Charges on the CS-4

disaccharide unit arise from the charged sulfate group in N-acetylgalactosamine

(GalNAc) as well as the charged carboxylic acid in Glucuronic Acid (GlcUA) resulting in

2 moles of charge per mole of CS disaccharide. The molecular weight of the CS

disaccharide is 502.5g/mole. The FCD (mEq/g of Macromolecule) can be estimated as:

𝑐𝐹 =2 𝑚𝑜𝑙 𝑐ℎ𝑎𝑟𝑔𝑒𝑠

502.5𝑔 𝐶𝑆 𝐷𝑖𝑠𝑎𝑐𝑐ℎ𝑎𝑟𝑖𝑑𝑒∗

𝑊𝐺𝐴𝐺

𝑔 𝑚𝑎𝑐𝑟𝑜𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒∗ 103

where WGAG is the CS weight in grams in the solution (252-253).

Page 271: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

245

CS content (WGAG) of CS-macromolecules was determined using the colorometric

1,9-dimethyl-methylene blue (DMMB) glycosaminoglycan assay (Blyscan Assay,

Biocolor, UK). DMMB sees a shift in absorbance when it associates with repeating

negative charges (sulfates) on the GAG chains that is detected spectrophotometrically at

530nm in the 96-well format (254). The detection range of this assay is 2-50µg/mL with

a detection limit at about 200ng. Fixed charge density of the CS-macromolecules was

then calculated and compared to natural CS (n=3).

7.2.4. Osmotic pressure of CS-macromolecule solutions

Osmotic pressure is the pressure that must be applied to a solution to prevent the

inward flow of fluid, and in the intervertebral disc, it is very sensitive to GAG

concentration. It depends mainly on the concentration of fixed charges on the PGs (i.e.

fixed charge density), as it arises from the Donnan distribution of ions between PGs and

the external fluid. Swelling pressure, the pressure at which there is no driving force for

fluid flow, results from the osmotic pressure exerted by the PGs and the resulting tension

in the collagen network of the IVD, which tends to restrain the swelling tendencies of the

PGs. At equilibrium, the osmotic pressure of the PGs is balanced by the tensile response

in the collagen network, opposing swelling (251).

Osmotic pressure of the CS-macromolecule solutions as well as natural CS was

measured via membrane osmometry. A custom built membrane osmometer was designed

and fabricated from 316 stainless steel in our laboratory (233) (Figure 7.1). In the

membrane osmometer system, a 36 Stainless Steel low chamber piece was fitted with an

Page 272: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

246

O-ring, stainless steel wire mesh and semi-permeable regenerated cellulose 100-500

MWCO membrane (Spectrum Labs), then screwed onto the top chamber piece. The

chamber was then filled with CS or CS-macromolecule solution (~1ml solution in 0.15M

NaCl). The chamber was sealed by fitting a miniature flush diaphragm pressure

transducer (PX61V Series, Electrical Connector Style, Sealed Gage, 0-200psi range,

Omega (Figure 7.2)) into the upper chamber piece (pressure is monitored in order to

ensure limited pressurization). The pressure transducer was connected to an amplifier

and data was transferred to acustom written LabView data acquisition software (Figure

7.3 and Figure 7.4). Prior to assembly, the pressure transducer voltage output was linearly

equated to pressure by applying a series of known pressures. The generated conversion

factor was input into the LabView program in order to convert voltage readings to PSI

and MPa pressure outputs. After assembly, the chamber was lowered into a saline bath

(0.15M NaCl solution) avoiding bubbles. Data was then collected until pressure readings

reached a steady-state value. Measurements were conducted at room temperature (25oC).

The membrane osmometry system was validated by measuring the osmotic pressure of

polyethylene glycol (20kDa in 0.15M NaCl) solutions over a range of concentrations

(233).

7.2.4. Fibroblast and Nucleus Pulposus Cytotoxicity studies

NIH 3T3 fibroblasts and Primary rat nucleus pulposus cells were used in order to

ascertain the cytotoxicity of the EG-CS and PEG-CS macromolecule formulations.

Preliminary studies on the cytotoxicity of EG-DGE and PEG-DGE monomers conducted

Page 273: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

247

in order to determine cytocompatible concentration ranges of the DGE monomers.

Similarly, cells were dosed with EG-CS and PEG-CS and investigated for cell viability

after 48 h and 72h. Fibroblasts were seeded at a density of 10,000cells/cm2 on tissue

culture plastic (12-well plate for Live/Dead Imaging assay, 96-well plate for MTT

assay)(RPMI media, 5% fetal bovine serum, L-glutamine and 1% pen/strep) and allowed

to attach for 24h before dosing with varying concentrations of EG-DGE and PEG-DGE

monomers or EG-CS and PEG-CS macromolecules. Samples were sterilized via

exposure to UV for 2h. Primary Nucleus Pulposus Cells cultures were also dosed with

EG-CS and PEG-CS and assayed for viability after 72h with the MTT assay. NP cell

culture was carried out in collaboration with Thomas Jefferson University, Jefferson

Medical College Department of Orthopaedic Surgery (PI: Makarand V. Risbud, Ph.D,

graduate student: Joanna McGough) following previously established methods (255).

Cell viability was investigated after 48h using the LIVE⁄DEAD®

Viability⁄Cytotoxicity Kit from Invitrogen which discriminates live cells from dead cells

by staining with green-fluorescent calcein-AM to indicate intracellular esterase activity

(cytoplasm) and red-fluorescent ethidium homodimer-1 to indicate loss of plasma

membrane integrity (nucleus stain). Images were collected on an inverted fluorescent

microscope (FITC and Rhodamine filters for live and dead cell imaging respectively) and

analyzed using Image J software (ImageJ 1.43u, NIH, USA). Four randomly chosen

locations were imaged for each well of the 12-well plate (live and dead cell images).

Images were then converted into binary black and white images (white regions denoting

stained cells). Average cell area for individual cells was determined then total area was

divided by the area of a single cell to determine the total number of live or dead cells in a

Page 274: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

248

single micrograph area. The % of live and dead cells was then calculated for each of the

four areas imaged and averaged resulting in % viability for each sample (n=3).

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

kit also from Invitrogen was used in order to colorometrically monitor cell viability and

proliferation after 48 and 72h dosing with DGE and CS-macromolecue solutions.

Metabolically active cells convert the water-soluble MTT to an insoluble purple

formazan. The formazan is then solubilized in DMSO and its concentration determined

by optical density at 540nm on a spectrophotometer.

Page 275: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

249

Figure 7.1: (Top) Schematic Diagram of Custom Built Membrane Osmometer and Data Acquisition System and (bottom) Image of Assembled Osmometer in Saline Bath

Page 276: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

250

Figure 7.2: Pressure Transducer Specifications

Page 277: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

251

Figu

re 7

.3:

Cus

tom

writ

ten

lab

view

pro

gram

for d

ata

acqu

isiti

on fr

om m

embr

ane

osm

omet

er.

Prog

ram

inte

rfac

es w

ith p

ress

ure

trans

duce

r, co

llect

s vol

tage

dat

a an

d co

nver

ts d

ata

to a

pre

ssur

e ou

tput

.

Page 278: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

252

Figure 7.4: Input and output pane for LabvView program written for data acquisition and processing of membrane osmometer pressure transducer output.

Page 279: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

253

7.3.1. TEM Imaging of PEG-CS Macromolecules

7.3 Results

TEM images of Uranyl Acetate stained PEG-CS macromolecules were obtained

after 24h and 72h of synthesis. TEM of natural CS and natural bovine aggrecan were

also taken as a point of comparison. CS appears as small condensed beads under TEM

(Figure 7.5). Natural bovine aggrecan takes on a beaded chain like structure and

aggregates of condensed beads (Figure 7.6). This is consistent with previous TEM

imaging studies of cartilage proteoglycans (256-258). After 24h of synthesis, PEG-CS

macromolecules have structures similar to that of CS, individual condensed beads (Figure

7.7). TEM of PEG-CS macromolecules after 72h of synthesis reveals larger chain like

aggregates of condensed beads, similar to that of natural bovine aggrecan (Figure 7.8).

Structures seen in TEM may be a result of sample aggregation with drying, however clear

differences are seen between CS and aggrecan and PEG-CS synthesized after 24h and

after 72h indicating the formation of larger structures with PEG-CS synthesis.

7.3.2. Viscosity of CS-macromolecule solutions.

The specific viscosity of CS-macromolecule solutions were determined in a cone-

on-plate rheometer at 25mg/ml concentrations in PBS (Figure 7.9). The specific viscosity

of CS (CS 25mg/ml, ηsp = 1.48) solutions was comparable to that of previous

measurements (CS 10mg/ml ηsp = 0.54, CS 50mg/ml ηsp = 4) (240, 259). The specific

viscosities of EG-CS and PEG-CS formulations were significantly higher than that of CS

Page 280: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

254

by an average of 9% and 11% respectively (Figure 7.9). PEG(1K)-CS solutions did not

have a significantly different specific viscosity from natural CS (Figure 7.9).

7.3.3. Fixed Charge Density of CS-macromolecules.

The fixed charge density (FCD) of EG-CS and PEG-CS macromolecules was

estimated from the GAG concentration of CS-macromolecule samples. FCD was

comparable to CS samples where CS samples had an estimated average FCD of

3.43mEq/g CS, EG-CS and PEG-CS samples had average FCD of 3.41mEq/g EG-CS and

3.61mEq/g PEG-CS respectively (Figure 7.10). FCD was significantly greater for PEG-

CS samples compared to both natural CS and EG-CS (p <0.05). EG-CS FCD was not

significantly different from CS samples.

7.3.4. Osmotic Pressure of CS-Macromolecule Solutions.

A membrane osmometer was custom built in order to directly measure the

osmotic pressure of CS-macromolecule solutions. The osmometer was validated with

PEG solutions (20K MW (Mw = 19,700) PEG in 0.15NaCl solution, room temperature)

of theoretically predicted osmotic pressure. PEG osmotic pressure (π, atm) is predicted

as:

𝜋𝑅𝑇

=𝑐

19,700+ 2.59𝑋10−3𝑐2 + 13.5𝑋10−3𝑐3

where c is PEG concentration in (g/ml), R is the Gas Constant (82.057cm3 atm K-1 mol-1)

and T is temperature (298.15 Kelvin) (244). PEG osmotic pressures determined

Page 281: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

255

experimentally using the membrane osmometer matched well with predicted values

(Figure 7.11). At higher PEG concentrations, (200mg/ml) the osmotic pressure was

underestimated at 0.412 MPa compared to the predicted value of 0.55MPa indicating the

upper limit of reliability of the membrane osmometer. At these higher pressures, the

integrity of the osmometer seals as well as the membrane may be compromised resulting

in lower osmotic pressure values. Between the range of 0.06MPa to 0.28MPa the

membrane osmometer is able to accurately predict osmotic pressure. CS solutions were

also measured for their osmotic pressure in the membrane osmometer system and

compared to literature findings for CS osmotic pressure also measured via direct

membrane osmometery. CS solution osmotic pressure in 0.15M NaCl at room

temperature was predicted as:

𝜋 ≈ 𝑐1𝑐𝑓 + 𝑐2(𝑐𝑓)2

where cf is fixed charge density and c1 and c2 are viral coefficients (0.500+/-0.053 and

0.645+/-0.091 respectively) determined by Chahine et al. in a custom built membrane

osmometer system for CS-6 (233). CS solution osmotic pressures were in general

agreement with predicted values (Figure 7.11). Small deviations in CS osmotic pressure

from predicted values is likely due to the primarily CS-4 population of our CS while

primarily CS-6 was used for determination of the viral coefficients. Together these

studies validate the use of the custom built membrane osmometer system for the

determination of CS-macromolecules solution osmotic pressure.

Osmotic pressure of 50mg/ml (0.15M NaCl, room temperature) CS-

macromolecule solutions was determined directly using the membrane osmometer. EG-

Page 282: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

256

CS, PEG-CS and PEG(1K)-CS average osmotic pressures were 0.145MPa, 0.129MPa,

and 0.128MPa respectively (Figure 7.12). Average CS osmotic pressure was 0.155MPa.

PEG-CS and PEG(1K)-CS solutions had significantly lower osmotic pressure than CS

solutions, while EG-CS did not have significantly different osmotic pressure from CS.

PEG-CS and PEG(1K)-CS solution osmotic pressures were also significantly lower than

that of EG-CS (n=3).

7.3.5. Cytotoxicity of DGE monomers and CS-macromolecules.

Cytotoxicity of DGE monomers and CS-macromolecules was determined by

dosing Fibroblast and NP cell cultures with varying concentrations of DGE monomer

(fibroblasts only) and CS-macromolecules. Live/Dead image analysis (48h) and MTT

metabolic activity analysis (48h and 72h) of Fibroblast cultures dosed with DGE

indicated that fibroblasts remain viable (no significant decrease in cell viability or

metabolic activity from PBS controls) when dosed with 0.001mM concentrations of DGE

(Figure 7.13). At higher concentrations of EG-DGE ( > 0.01mM EG-DGE), cell viability

is compromised as seen by a statistically significant decrease in the % of live cells and

metabolically active cells when compared to control cultures. PEG-DGE maintains

cytocompatibility at 0.01mM and has significantly higher percentage of live and

metabolically active cells than EG-DGE dosed cultures at 0.01 and 0.1mM

concentrations (Figure 7.13). Cytocompatibility of 0.001mM EG-DGE dosed and

0.001mM and 0.01mM PEG-DGE dosed cultures was maintained after 72h of culture

while 1mM DGE concentrations were acutely toxic to cells (Figure 7.13).

Page 283: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

257

Fibroblast cultures were dosed with varying concentrations of EG-CS and PEG-

CS macromolecules and allowed to culture for 48h. Cultures were stained for live and

dead cells to determine cell viability. Cultures dosed with 0.2mg/ml and 2mg/ml CS,

EG-CS and PEG-CS appeared maintain cell viability while in 20mg/ml dosed cultures, a

large number of dead cells is seen in CS cultures, and cells appear more rounded in EG-

CS and PEG-CS cultures (Figure 7.14). Quantification of Cell viability after 48h of

dosing demonstrates significantly reduced fibroblast viability in 20mg/ml CS and EG-CS

cultures, while PEG-CS cultures do not have reduced viability at any concentration tested

(Figure 7.14). Due to the overall toxicity seen in 20mg/ml dosed cultures, this

concentration was not investigated further.

Primary nucleus pulposus cell and fibroblasts were both dosed with CS

and CS-macromolecules for 72 and investigated for culture metabolic activity. Both

Fibroblast and NP cell cultures maintained viability at 0.2mg/ml and 2mg/ml for all

macromolecules investigated (no significant decrease in cell metabolic activity compared

to TCP controls) (Figure 7.15). In fibroblast cultures, 0.2mg/ml PEG-CS dosed samples

had significantly increased cell metabolic activity compared to TCP controls while this

increase in metabolic activity was seen at 2.0mg/ml PEG-CS in NP cell cultures.

Fibroblast and NP cell cultures dosed with 10mg/ml CS and EG-CS had significantly

lower metabolic activity compared to TCP controls while 10mg/ml PEG-CS samples did

not show any significant cytotoxic effects (Figure 7.15). NP and Fibroblast Cultures

dosed with the full range of PEG-CS concentrations maintained cell viability and in some

instances, increased proliferation in Fibroblast cultures.

Page 284: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

258

Figure 7.5: TEM of CS (A) and close-up of condensed bead structure of CS (B)

Figure 7.6: TEM of (A) Several aggrecan molecules clustered together. (B) Close-up of natural bovine ggrecan where condensed beads are arranged in a chain pattern.

Page 285: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

259

Figure 7.7: TEM of PEG-CS macromolecules after 24h of synthesis at lower (A) and higher (B) magnification. PEG-CS appears as small aggregates and individual condensed beads.

Figure 7.8: TEM of PEG-CS synthesized for 72h at lower magnification (A) demonstrating aggregates of PEG-CS and (B) higher magnification demonstrating beaded chain like structure of

PEG-CS.

Page 286: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

260

Specific Viscosity of CS-Macromolecules(25mg/ml in PBS, 25°C)

10 1001.3

1.4

1.5

1.6

1.7

1.8CSEG-CS (EG MW 174)PEG-CS (PEG MW 526)PEG-1000-CS (PEG1K MW 1000)

Shear Rate (1/s)

Spec

ific

Vis

cosi

ty

Change in Specific Viscosity ofCS-Macromolecules in Comparison to CS

CS

EG-CS

PEG-CS

PEG(1K)-C

S0.0

0.5

1.0

1.5**

***

*

* Statistical Difference, p < 0.05** Statistical Difference, p < 0.01

CS-Macromolecule

Cha

nge

in S

peci

fic V

isco

sity

(Nor

mal

ized

to C

S)

Figure 7.9: (Top) Specific Viscosity with shear rate of CS-macromolecules with varying PEG backbone lengths. (bottom) Specific viscosity of CS-macromolecules relative to CS

(1-way ANOVA statistical analysis, n=3).

Page 287: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

261

Fixed Charge Density of CS-Macromoleculesbased on GAG content

CS

EG-CS

PEG-CS

0

1

2

3

4*

*

* Statistically Significant Difference, p < 0.05CS-Macromolecule

Fixe

d C

harg

e D

ensi

ty(m

Eq/g

mac

rom

olec

ule)

Figure 7.10: Fixed Charge Density of CS-Macromolecules as determined by the Blyscan GAG assay. (1-way ANOVA statistical analysis, n=3)

Page 288: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

262

Validation of Membrane Osmometerwith PEG (20K MW) Solutions

0 100 200 3000.0

0.2

0.4

0.6

0.8

1.0

Predicted PEG MPaExperimental PEG MPa

PEG Concentration (mg/mL)

Osm

otic

Pre

ssur

e (M

Pa)

CS Solution Osmotic PressureRelative to Predicted Values

0 50 100 150 200 2500.0

0.2

0.4

0.6

0.8

1.0Predicted CS MPa Lower LimitPredicted CS MPa Upper LimitExperimental CS MPa

CS Concentration (mg/mL)

Osm

otic

Pre

ssur

e (M

Pa)

Figure 7.11: (Top) Validation of membrane osmometer using PEG solutions of known osmotic pressure (20K MW PEG solutions in 0.15M NaCl, room temperature) (244).

(Bottom) Validation of CS measurements on membrane osmometer with comparison to predicted literature values(233). Upper and Lower limit of CS osmotic pressure refers to the (+/-) standard

deviation of the viral coefficients.

Page 289: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

263

Osmotic Pressure Curves fromMembrane Osmometer

0 5 10 150.00

0.05

0.10

0.15

0.20

PEG-CS

CSEG-CS

PEG(1K)-CS

Time (h)

Osm

otic

Pre

ssur

e (M

Pa)

Osmotic Pressure with Varying DGE

CS

EG-CS

PEG-CS

PEG(1K)-C

S0.00

0.05

0.10

0.15

0.20

**

* Significantly different (p<0.05) from CS+ Significantly different (p<0.05) from EG-CS

+ +

Osm

otic

Pre

ssur

e (M

Pa)

Figure 7.12: (Top) Representative Osmotic Pressure curves from Membrane Osmometer of CS-macromolecule solutions (50mg/ml, 0.15M NaCl, 25C̊, 1mL Volume). (Bottom) Average Steady State Osmotic Pressure of CS-macromolecule solutions (1-way ANOVA statistical analysis n=3).

Page 290: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

264

48h Fibroblasts ViabilityDosed With Diglycidyl Ethers (DGEs)

00.0

01 0.01 0.1 1

0

20

40

60

80

100PBSEG-DGEPEG-DGE

******

******

*** ***

DGE Concentration (mM)

% L

ive

Cel

ls(L

ive/

Dea

d As

say)

48h Fibroblast Metabolic Activitywith Diglycidyl Ethers (DGEs)

00.0

01 0.01 0.1 1

0.0

0.5

1.0 *

***

******

****

* Significant Difference, p < 0.05** Significant Difference, p < 0.01

*** Significant Difference, p < 0.001(n = 3)

DGE Concentration (mM)

Rat

io o

f M

etab

olic

ally

Act

ive

Cel

ls(M

TT C

onve

rsio

n)

72h Fibroblast Metabolic Activitywith Diglycidyl Ethers (DGEs)

00.0

01 0.01 0.1 1

0.0

0.5

1.0

PBSEG-DGEPEG-DGE

***

***

***

***

***

***

***

**

DGE Concentration (mM)

Figure 7.13: Fibroblast cytotoxicity with exposure to varying concentrations of EG-DGE and PEG-DGE as determined by (top) live/dead imaging (Images taken at 10X magnification) and

(bottom) MTT metabolic activity assay. (2-way ANOVA statistical Analysis)

Page 291: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

265

48h Viability of FibroblastsDosed With Biomimetic Aggrecan

0 0.2 2 200

20

40

60

80

100PBSCSEG-CSPEG-CS

***

***

******

***

*** Significant Difference, p < 0.001(n=3)

CS-Macromolecule Concentration (mg/ml)

% L

ive

Cel

ls(L

ive/

Dea

d As

say)

Figure 7.14: Live/Dead imaging and % Viability analysis of Fibroblasts dosed with varying concentrations of EG-CS and PEG-CS macromolecules after 48h incubation.

(2-way ANOVA statistical Analysis)

Page 292: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

266

72h FibroblastMetabolic Activity

Dosed with CS-Macromolecules

0 0.2 2 100.0

0.5

1.0CSEG-CSPEG-CS

TCP*

**

*

CS-Macromolecule Concentration (mg/ml)

Rat

io o

f M

etab

olic

ally

Act

ive

Cel

ls(M

TT N

orm

aliz

ed to

TC

P C

ontr

ol)

72h Nucleus Pulposus CellMetabolic Activity

Dosed with CS-Macromolecules

0 0.2 2 100.0

0.5

1.0CSEG-CSPEG-CS

TCP

**

***

*******

* ***

* Statistical Difference, p < 0.05** Statistical Difference, p < 0.01

*** Statistical Difference, p < 0.001(n=4)

CS-Macromolecule Concentration (mg/ml)

Rat

io o

f M

etab

olic

ally

Act

ive

Cel

ls(M

TT N

orm

aliz

ed to

TC

P C

ontr

ol)

Figure 7.15: 72h Cytotoxicity of Fibroblasts and NP Cells dosed with CS-macromolecules at varying concentrations as determined by the MTT assay

(2-way ANOVA statistical analysis)

Page 293: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

267

In this study, the physical and function characteristics of CS-macromolecules

were investigated in relation to the requirements of a biomimetic aggrecan

macromolecule for the restoration of nucleus pulposus mechanical and osmotic function.

CS-macromolecules were synthesized with CS chains linked by synthetic ethylene glycol

chains of length 0.5 to 5nm (EG-CS and PEG-CS formulations respectively).

7.4 Discussion

7.4.1. Physical structure of CS-macromolecules

Previous GPC studies suggested that EG-CS and PEG-CS macromolecules

consisted of a mixed population of chains with single CS attachments (Mn ~ 19KDa),

two CS attachments (dimmers Mw ~ 42KDa) and possibly small populations of

macromolecules with more than two CS attachments (Mz~78KDa) (Table 6.6). PEG-CS

macromolecules were further investigated for their molecular size and structure by

visualization with TEM. TEM images suggest the formation of larger molecular

structures after 72h of synthesis compared to natural CS and samples synthesized for just

24h. Reaction kinetics of the CS primary amine, indicate that the majority of primary

amine reactions are complete after 24h however, possible secondary amine reactions may

continue to occur after the 24h period resulting in the step-wise synthesis on some larger

structures (also supported by 1H-NMR spectral analysis indicating continued

incorporation of PEG moieties even after the initial 24h primary amine reaction time).

Under TEM investigation PEG-CS samples take on an aggregated and chain like

conformations similar to that of natural aggrecan however these structures may be

Page 294: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

268

influenced by sample staining and dehydration (258). The processing of CS-

macromolecules and aggrecan for TEM imaging requires the dehydration of the samples

and staining with uranyl acetate. This processing causes condensation of the CS chains

into bead like structures versus their more natural rod like structures, thereby making it

difficult to discern the bottle brush arrangement of CS seen in natural aggrecan. In

addition drying may result in aggregates that are not necessarily present in solution.

Cryo-TEM techniques, where samples are placed onto the copper imaging grid in a thin

layer of water then flash frozen to avoid crystal formation and disruption of the molecule,

may be used as an alternative to tradition TEM in order to image CS-macromolecules and

aggrecan in a more natural state (260). Atomic force microscopy (AFM) has also been

used to visualize aggrecan structures and may be useful in the investigation of CS-

macromolecules structure (75).

Viscosity investigations of CS-macromolecule solutions also indicated the

formation of structures different from natural CS as evidenced by an increase in viscosity

of EG-CS and PEG-CS compared to CS. Viscosity of a macromolecule solution is

influenced by the solute molecular weight as well as its hydrodynamic radius and

structure (i.e. rigid backbone, worm-like configuration, branching, charge) (170, 261).

The increased viscosity of PEG-CS and EG-CS confirms observations of increased MW

as seen in GPC as well as structure formation as seen in TEM. Interestingly, PEG(1K)-

CS samples did not have an increased viscosity in comparison to CS. Previous primary

amine reaction studies and 1H-NMR-investigations indicated incorporation of PEG(1K)

at the CS primary amine, however GPC and TEM studies were not conducted on these

samples due to the long PEG chain incorporation which is not physiologically relevant.

Page 295: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

269

A lack of increased viscosity of these samples indicate that the structures formed in

PEG(1K)-CS do not have significantly different flow properties from natural CS. This

may be due to the relatively long PEG linking chain (~12.5nm) between CS side chains

compared to EG-CS and PEG-CS macromolecules. This chain length is compared to the

radius of gyration of CS-6 which has been estimated at 12.8nm (207). The radius of

gyration of CS-4 (used in the current investigation) is likely smaller than 12nm due to the

smaller MW of CS-4 compared to CS-6. Therefore, the PEG distance between CS chains

in PEG(1K)-CS is longer than that of the CS radius of gyration and interaction distance.

Because of this large distance between CS chains, strong interactions between individual

CS chains on the PEG(1K)-CS may not be occurring, reducing the effects of CS linking

on solution viscosity (262-264). Also, since large MW species are not likely being

synthesized in the current samples, the direct influence of MW on viscosity may be low.

Viscosity can be a powerful tool to probe polyelectrolyte and brush polymer structure

where the concentration dependence of viscosity is largely influenced by macromolecule

charge, rigidity and branched structure (205, 261, 263, 265-267). As larger MW and

highly branched CS-macromolecules are synthesized, viscosity techniques may be useful

to characterize the structural properties of the brush-like CS-macromolecules.

7.4.2. Osmotic Function and Fixed Charge Density of CS-macromolecules

In the function of CS-macromolecules as a biomimetic replacement for natural

aggrecan in NP tissues it is important the molecules provide the osmotic functions of

natural aggrecan. These properties of natural aggrecan are provided by the CS chains

Page 296: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

270

attached to the aggrecan core protein (244). Due to the fixed charges on CS immobilized

in the NP matrix, counter ions build up in the NP and the tissue is able to draw in fluid (to

balance the ionic imbalance) until a hydrostatic pressure is generated that resists stresses

arising from tensile forces of the collagen network and external loads applied by muscles

and ligaments Figure 7.16. The osmotic pressure of proteoglycans at concentrations found

in NP tissues (0.18 – 0.35 meq/gH2O, fixed charge density) has been determined to lie in

the range of approximately 0.03 to 0.3MPa (268).

The osmotic potentials of EG-CS, PEG-CS and PEG(1K)-CS were

determined in comparison to that of natural CS. EG-CS macromolecules, which have the

smallest PEG linkage length, did not have significantly different osmotic pressure than

CS indicating maintenance of the CS osmotic potential in the EG-CS macromolecule

formulation (Figure 7.12). PEG-CS and PEG(1K)-CS macromolecule solution did have

significantly decreased osmotic pressure compared to CS and EG-CS. This is likely due

to the incorporation of larger MW PEG segments in PEG-CS and PEG(1K)-CS (Figure

7.12). PEG has the ability to generate an osmotic potential when separated from an ionic

solution via a semi-permeable membrane, however, the osmotic pressure generated by

PEG solutions is much lower than that generated by CS of the same concentrations

(Figure 7.11). This is due to the high concentration of negative charges along the CS

backbone. The incorporation of the synthetic less osmotically-active PEG backbone into

the CS-macromolecule slightly reduces the osmotic potential of the CS-macromolecules.

Although there is a slight drop in osmotic potential, CS-macromolecule osmotic potential

is still in the range of natural CS and much higher than that of 20K MW PEG at the same

concentration (0.026MPa for 20K MW PEG vs 0.128MPa for PEG-CS and PEG(1K)-

Page 297: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

271

CS) indicating dominance of the CS portion of the CS-macromolecules in osmotic

function (Figure 7.12).

Osmotic pressure measurements can also be used to probe the structural

properties of macromolecules where the configurational entropy of the macromolecules

in solution have been theoretically demonstrated to contribute to the osmotic swelling

pressure of concentrated proteoglycan solutions (269). It is expected that the

organization of CS chains into a bottle brush structure will impact the osmotic pressure of

the macromolecule solutions but it is unlikely that we will see such effects in our

macromolecules due to the low degree of polymerization in our formulations. An

osmotic technique for determining macromolecule structural properties may be useful in

the future as additional CS-macromolecule species are developed.

The osmotic function of CS is largely due to the fixed charges along the CS

backbone. With aging and degeneration the fixed charge density of the NP decreases. In

one study, FCD ranged as high as 1.4mEq/g dry tissue weight in a 14y.o. human lumbar

NP to as low as 0.15mEq/g dry tissue weight in 91y.o. human lumbar NP thus reducing

the osmotic potential of the NP (244). In a recent modeling study conducted by Massey

et al. (270-271), the effect of FCD restoration on intervertebral disc stress distributions

was investigated. Several grades of disc degeneration (Grade 1 (“normal”) to Grade

5(degenerate)) were modeled (axisymmetric, osmo-poroelastic finite element model) by

altering tissue material properties, such as modulus, fixed charge density, and tissue

permeability. In order to model the effect of FCD restoration on stress distribution in the

NP and annulus fibrosus (AF), FCD profiles for healthy Grade 1 discs (Figure 7.17) were

applied to degenerated discs while keeping all other degenerated material properties

Page 298: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

272

constant. Reverting the fixed charge density profile to its original Grade 1 state

decreased the stress experienced by the annulus fibrosus and drastically changed the

stress on the nucleus pulposus, even though other material properties were still in a

degenerated condition (270-271). This implies that a treatment, which restores fixed

charge density to the NP may have far reaching effects on both nucleus and annulus

stress distributions and reactions under loading.

In order to investigate the use of CS macromolecules for the restoration of FCD to

degenerated NP, CS-macromolecule FCD was determined and compared to natural CS.

FCD of EG-CS macromolecules were not significantly different from that of CS while

PEG-CS macromolecules had slightly higher FCD. This may be due to small degrees of

non-specific interactions between the PEG backbone and Blyscan assay reagents. In both

cases, CS-macromolecules had FCD comparable to natural CS. The average FCD of the

CS-macromolecules was approximately 3.5 mEq/g macromolecule. Using this FCD

value, the amount of CS-macromolecule needed to restore FCD as modeled in the study

of Massey et al. was calculated and is presented in Table 7.1 and compared to that of

natural aggrcan. These values are based o the extrapolated FCD for the varying

degenerative grade tissue models at mid nucleus (Fractional Sagittal Section = 0.5)

(Figure 7.17). The worst case scenario of FCD restoration is seen for Grade 5 tissues,

however it is unlikely that such an intervention would take place since a biomimetic

aggrecan replacement approach would require a distinct NP region and an intact annulus

which is not the case for Grade 5 Degeneration. By Grade 4 or 5 degeneration, the most

likely intervention would be spinal fusion or total disc replacement. At the lower grades

of degeneration, Grade 3 and 2, approximately 21.43 and 10.71mg/g wet tissue of CS-

Page 299: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

273

macromolecule would be required to restore Grade 1 FCD profiles. If the NP is

approximated to be 3g wet weight, approximately 64mg and 32mg of CS-macromolecule

would be needed to restore Grade 3 and Grade 2 discs respectively. These are reasonable

masses that may be implanted via solution injection into the NP where concentrations of

CS-macromolecule up to 1g/ml are injectable through a 16 gauge needle (data not

shown). The cost of therapeutic amounts of natural bovine aggrecan and synthesized CS-

macromolecules was estimated and compared to the cost of the commonly utilized steroid

injections, for the treatment of back pain (Table 7.2). Aggrecan costs ($7,837-$31,347 to

restore Grade 2-Grade 5 respectively) far exceed the currently accepted cost for injectable

treatments to the spine (~$1,850 (272)) while CS-macromolecule costs are considerably

less for therapeutic masses ($2.35-$9.39 to restore Grade 2-Grade5 respectively) further

motivating the usefulness of a biomimetic aggrecan strategy.

It is important to note that these restorative amounts are based on estimated NP

volume as well as estimated FCD at the varying degenerative grades. FCD

measurements of IVD tissue vary greatly due to differences in tissue processing, FCD

detection and reporting (i.e. FCD/wet weight, GAG/dry weight etc) as well as natural

variations in FCD from disc to disc and between spinal level (30, 84, 244, 253, 268, 273).

In some reports, proteoglycan content in healthy NP has been estimated as high as

700ug/mg dry tissue, much higher than that utilized in our estimations (30). However,

the restorative CS-macromolecule amounts determined here have been demonstrated

through modeling studies to have significant effects on stress distribution in the NP and

AF. In addition, FCD is intimately related to osmotic pressure, where direct relationships

between fixed charge density and osmotic pressure have been established for GAG

Page 300: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

274

solutions (233, 268). These direct positive outcomes of FCD restoration demonstrate the

potential benefits of biomimetic aggrecan injections on IVD mechanical function.

Another major element of the aggrecan molecule in imparting mechanical

properties to IVD tissue is the arrangement of fixed charges in the bottle brush

arrangement. The distance between adjacent CS molecules effects their physical

resistance to force. Electrostatic repulsion forces occur when intermolecular distance

between CS chains is 2-4nm (69). These electrostatic forces between CS chains account

for 50% (290kPa) of the equilibrium compressive elastic modulus as predicted by

theoretical modeling (70, 72). The arrangement of fixed charges in proteoglycans has

also been associated with higher osmotic pressure generation than CS alone (268).

Additionally, the brush structure of aggrecan has been demonstrated to act as an ion

reservoir further providing osmotic resistance to applied loads (274). However, FCD

arrangement is not taken into account in the modeling of FCD restoration to the NP. It

is also assumed that the FCD carriers remain immobilized in the matrix once inputed. It

is therefore important to continue to work towards large MW bottle brush like CS-

macromolecules in order to capture the 50% contribution to modulus of electrostatic

repulsions as well as maintain FCD immobilization in the NP matrix. Strategies which

immobilize CS bottle brush macromolecules in the NP matrix and thereby restore NP

FCD and electrostatic forces may lessen the need for nucleus pulposus and annulus

fibrosus tissue to remodel and accommodate abnormal stresses experienced during

degeneration, thereby halting the degenerative cascade.

Page 301: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

275

7.4.2. Cytotoxicity of CS-macromolecules

In a therapeutic strategy, biomimetic aggrecan would be injected into the nucleus

without removal of the nucleus tissue (i.e. augmentation). In this scenario, cells in NP

would be directly exposed to the injected macromolecules. NP cells are not highly

active, however they are capable of producing biomolecules in order to maintain the NP

extracellular environment (76). The synthesis process of the cells is balanced by the

expression on MMPs and aggrecanses which break down the matrix for ECM turnover

(7). In the degenerated case, there is an imbalance in matrix molecule synthesis and

breakdown, resulting in a decrease in matrix material properties. If the cells of the NP

are removed or adversely affected by the interventional strategy, any matrix conservation

capacity of the NP will be severely hindered, possibly causing greater long term damage.

Therefore it is important to ensure that CS-macromolecule formulations do not have

cytotoxic effects on the cells of the NP. Initial cytotoxicity studies were conducted on a

fibroblast cell line. EG-DGE and PEG-DGE were investigated for their direct effect on

cell cytotoxicity. Both EG-DGE and PEG-DGE were cytocompatible at trace amounts

(0.001mM) while PEG-DGE was slightly more cytocompatible than EG-DGE (PEG-

DGE maintained cytocompatibility at 0.01mM concentrations) (Figure 7.13). Therefore it

is reasonable to assume, that if trace amounts of EG-DGE or PEG-DGE remain in the

CS-macromolecule formulations, no significant cytotoxic effects would be expected.

This is an important point as 1H-NMR investigates demonstrate that the bulk of excess

DGE is removed from the CS-macromolecules, but there may be trace amounts of DGE

in the final formulations that are un-detectable by the 1H-NMR spectral analysis Figure

6.14 and Figure 6.16). DGE cytotoxicity is in agreement with previous studies on the

Page 302: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

276

cytotoxicity of a range of diepoxy compounds where higher MW DGEs were more

cytocompatible than smaller MW species (234). Cytotoxicity of the DGE monomers is

likely due to irreversible chemical modification of proteins or polysaccharides present on

cells by the epoxides of the diepoxide compound (234).

EG-CS and PEG-CS macromolecule cytotoxicity was investigated for both

fibroblast and NP cell cell lines. EG-CS and PEG-CS macromolecules were not

cytotoxic to cells at 0.2mg/ml. This concentration is in the range typically experienced

by cells in the NP. If the density of cells in a relatively young NP (16-29y.o.) is

estimated at approximately 2,780,000cells/cm3 (249) and the FCD and therefore GAG

content of a similarly aged NP (27y.o.) is assumed to be 0.35mEq/g H2O (equivalent to

90mg/ml of GAG) (275) then the GAG/CS ratio is approximately 3.23x10-5mg/cell of

GAG. In the current culture system, 2,800 cells were seeded at a density of

10,000cells/cm2 to TCP then dosed with 0.2mg/ml of CS or CS-macromolecules resulting

in a concentration of approximately 7.6x10-5mg/cell. The vast majorities of studies

which investigate the effects of in vitro CS on chondrocyte cultures (cell lines similar to

NP cells) also investigate CS concentrations in the range of 10-1000µg/ml and have

found that these concentrations of exogenously applied CS actually promote the

production of proteoglycans and inhibit MMP synthesis in vitro (276-278).

Since CS-macromolecules would likely be introduced to the NP environment in a

bolus, a transient period will exist where NP cells experience CS concentrations much

higher than usually physiologically experienced. It is therefore important to investigate

the possible cytotoxic effects of the CS-macromolecules at higher concentrations than

seemingly physiologically relevant. Concentrations of CS, EG-CS and PEG-CS were

Page 303: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

277

investigated up to 20mg/ml. At 10mg/ml and 20mg/ml (fibroblast only) CS and EG-CS

macromolecules were found to by cytotoxic to fibroblasts after 48h of culture while PEG-

CS did not appear to significantly affect cell viability. All macromolecules were non-

cytotoxic at 2mg/ml in fibroblasts while EG-CS showed some cytotoxic effects at this

concentration in NP cells. Overall, cell reaction to CS and CS macromolecule dosing

was similar in fibroblasts and NP cells with NP cells being slightly more sensitive to the

exogenous macromolecules. This is not surprising as the NIH 3T3 fibroblast cell line is

fairly robust, while primary cell lines such as the NP cells studied here are more difficult

to maintain in culture. NP cells cultures studied here were cultured in normal oxygen

tension and pH while NP cells in vivo are accustomed to very specific oxygen, osmotic

and pH environments that are unique to this tissue. A surprising result was the increased

cytocompatibility of PEG-CS macromolecules at high concentrations compared to both

CS and EG-CS cultures. PEG-CS incorporates a longer PEG linker than EG-CS. The

increased incorporation of the synthetic PEG backbone may have had beneficial effects

on cell cytocompatibility. This has been seen in a study conducted by Petersen et al

(279) where linear PEG chains were grafted to branched polyethleneimine (PEI) cationic

polymers to generate cationic polyelectrolytes grafted with nonionic hydrophilic

polymers, similar to our system. In their study, a series of polymers were synthesized

with varying MW PEG chains and they found that cytotoxicity of 3T3 mouse fibroblasts

decreased with increasing PEG chain MW. This was likely due to shielding of the

cationic blocks by the PEG chains which reduced the cell membrane destruction caused

by the cationic domains (279). A similar scenario may be occurring in our PEG-CS

macromolecule system. PEG has been demonstrated to exclude proteins and other

Page 304: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

278

macromolecules and particulates from its surroundings due to its high chain mobility,

conformational flexibility and water binding ability (250). The ability of PEG to shield

itself from other macromolecules may have reduced the cytotoxicity of CS seen at these

high concentrations. These studies suggest that the initial introduction of PEG based

CS-macromolecules to the NP at higher than physiological concentrations will not have a

deleterious cytotoxic effect.

Page 305: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

279

Figure 7.16: Schematic diagram showing the fixed charges of the NP extracellular matrix and associated mobile counter ions (252).

Page 306: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

280

Figure 7.17: Extrapolated fixed charge density profiles for intervertebral disc tissue of varying degenerative grade (assuming 26yo represents Grade 1 (“healthy”) and 74yo represents Grade 5

discs) (84, 270-271)

Page 307: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

281

Table 7.1: Predicted Restorative Masses of CS-macromolecules and Natural Aggrecan based on FCD of degenerated NP tissues (84, 270-271)

Degenerative Grade

Nucleus Pulposus

FCD (mEq/g wet

tissue)

FCD Required to Restore to Grade 1 FCD (mEq/ g wet

tissue)

£Mass of Natural Aggrecan Required

to Restore FCD (mg/g wet tissue)

(FCD~3.23mEq/g)

Mass of CS-Macromolecule Required

to Restore FCD (mg/g wet tissue)

( FCD ~3.5mEq/g)

Grade 5 0.15 0.15 46.44 42.86 Grade 4 0.19 0.11 58.05 32.14 Grade 3 0.23 0.08 69.66 21.43 Grade 2 0.26 0.04 81.27 10.71 Grade 1 0.3 N/A N/A N/A

£FCD of natural aggrecan calculated from Blyscan Assay analysis of bovine cartilage aggrecan (sigma) sulfated GAG content (data not shown). GAG estimated to comprise of 77% CS chains and 23% KS Chains (7). FCD (mEq/g aggrecan) estimated based on(253):

𝑐𝐹 = ��2 𝑚𝑜𝑙 𝑐ℎ𝑎𝑟𝑔𝑒𝑠 𝐶𝑆502.5𝑔

∗ 0.77∗𝑊𝐺𝐴𝐺𝑔 𝑚𝑎𝑐𝑟𝑜𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒

�+ �1 𝑚𝑜𝑙 𝑐ℎ𝑎𝑟𝑔𝑒𝑠 𝐾𝑆454𝑔

∗ 0.23∗𝑊𝐺𝐴𝐺𝑔 𝑚𝑎𝑐𝑟𝑜𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑒

�� ∗ 103

Page 308: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

282

Table 7.2: Estimated Cost of CS-macromolecule for therapeutic application in comparison to estimated cost of restoration with natural aggrecan and current treatment option of steroid

injections.

Degenerative Grade

£Cost of Aggrecan to Restore FCD (USD)

(assuming 3g NP wet weight)

§Cost of CS-Macromolecule to Restore

FCD (USD) (assuming 3g

NP wet weight)

Mean Cost per Epidural Steroid Injection

Treatment (as estimated in 2008)

(272) Grade 5 $31,347 $9.39

$1,850 Grade 4 $23,510 $7.04 Grade 3 $15,673 $4.69 Grade 2 $7,837 $2.35 Grade 1 N/A N/A N/A

£Calculated based on natural bovine cartilage aggrecan cost of $225/mg (Sigma)

§Calculated based on the following figures:

"Grafting-Through" Synthesis Supplies

Estimated Cost to Synthesize 1g

SBB Buffer $0.07 Dialysis Tubing $3.82 Chondroitin Sulfate $54.20 PEG-DGE $0.12 EG-DGE $0.22 25% Loss due to Purification $14.55

Total Cost PEG-CS $72.76 Total Cost EG-CS $72.89

Page 309: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

283

The results presented here demonstrate that CS-macromolecules incorporating a

synthetic polyethylene glycol component have functional properties which make them

suitable for NP proteoglycan restoration. CS-macromolecules were demonstrated to have

structural features distinct from that of natural CS, indicating the formation of larger

structures. These macromolecules were further found to have high osmotic potential and

fixed charge density similar to that of natural CS. The osmotic function of CS-

macromolecules is critical in the restoration of NP hydration and hydrostatic pressure.

CS-macromolecules were also found to be non-cytotoxic at both physiological

concentrations and higher concentrations that may occur as a result of CS-macromolecule

delivery to the NP. The incorporation of a synthetic polymeric linkage between CS

chains had effects on several CS properties including specific viscosity, osmotic pressure

and cytotoxicity. The linkage of CS with relatively short PEG chains resulted in an

increase in solution viscosity compared to natural CS while incorporation of relatively

long PEG chains resulted in a decrease in solution osmotic pressure. Cytotoxicity of CS

was decreased the incorporation of PEG chains. CS-macromolecule properties are

clearly dominated by the CS component however deviations in properties from natural

CS can be directly attributed to properties of the incorporated synthetic polyethylene

glycol backbones. Therefore, the choice of synthetic backbone in the synthesis of

biomimetic aggrecan will be critical in the function of the biomimetic macromolecule.

CS-macromolecules investigated in this study have functional properties which will be

beneficial in a biomimetic aggrecan approach to the replacement of depleted

proteoglycan content in the NP. In a completely biomimetic strategy, larger structures

7.5 Conclusions

Page 310: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

284

which incorporate closely packed CS chains will need to be synthesized in order to

ensure macromolecule immobilization in the NP matrix and the generation of

electrostatic forces critical in NP mechanical function.

Page 311: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

285

Chapter 8 : Conclusions

The goal of this study was to develop an enzymatically resistant, hybrid

synthetic/bio-based biomimetic aggrecan macromolecule for use in the treatment of

intervertebral disc degeneration. Biomimetic aggrecan was designed such that it

incorporates a synthetic polymeric core to resist enzymatic degradation while maintaining

natural CS bristles which are biological activity and carry high fixed charge density. In

order to synthesize bottle-brush like macromolecules from natural CS a terminal handle

on CS was first identified then utilized in “grafting-to” and “grafting-through” strategies

of bottle brush polymer synthesis.

8.1 Summary

A terminal primary amine on commercially available CS-4 was identified and

found to be reactive to epoxides, aldehydes and carboxylic acids at varying reactivities.

Carboxylic acid-amine reactions were investigated for the “grafting-to” strategy due to

the high water solubility and biocompatibility of carboxylic acid polymers while epoxide-

amine reactions were investigated for the “grafting-through” strategy due to the relatively

high reactivity of the CS primary amine to epoxides. CS-macromonomers were

synthesized by the covalent reaction of the epoxide allyl glycidyl ether to the CS terminal

amine. These vinyl terminated CS-macromonomers may be utilized in future studies for

the free-radical polymerization of CS into bottle brush structures via the “grafting-

through” strategy. Due to complexity involved in possible side chain reactions and harsh

conditions generally associated with free-radical polymerization, this strategy was not

investigated further in this study.

Page 312: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

286

In the “grafting-to” strategy CS was covalently bound to poly(acrylic acid)

backbones. Several parameters including reaction time, reaction temperature, solution

ionic strength and polymer:CS molar ratio were investigated for their effect on CS

conjugation to the PAA backbone. CS-PAA macromolecules were synthesized with as

many as 9 CS chains on each 250K MW PAA backbone, resulting in sparsely grafted

chains. Due to the limitations of this strategy in fabricating densely grafted CS-

macromolecules with electrostatically relevant CS spacing (~2-4nm spacing), further

synthesis strategies were investigated.

The step-growth “grafting-through” strategy of polymer synthesis was

investigated as a means of synthesizing CS-macromolecules with controlled CS-spacing.

Step-growth polymerization is also desirable due to the relatively mild reaction

conditions that may be applied. In this study, the epoxide-amine reaction was utilized,

employing the reactivity of epoxides to both the hydrogens of the CS terminal amine

(primary amine then secondary amine reactions). Poly(ethylene glycol) based di-

epoxides, EG-DGE and PEG-DGE were utilized in order to respectively impart ~ 0.5 and

5nm spacing between CS chains in the final CS-macromolecule. Several parameters

including reaction time, reaction temperature, reactant ratio, and DGE MW were

investigated for their effect on CS primary amine to DGE epoxide reaction rate and CS-

macromolecule synthesis. CS macromolecules were synthesized that incorporated

synthetic PEG components, likely at the terminal end of CS, linking CS chains together.

Molecular weight and structural analysis of EG-CS and PEG-CS macromolecules suggest

that the macromolecules consist of a mix of CS monomers, dimmers and possibly some

larger molecular weight species. Osmotic properties and cytocompatibility of CS-

Page 313: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

287

macromolecules suggest that CS-macromolecules with poly(ethylene glycol) based

backbones may be useful in the restoration of proteoglycan function in degenerated

nucleus pulposus.

Several parameters were set fourth for the design of a biomimetic aggrecan

macromolecule that can restore the function of the nucleus pulposus in the early stages of

disc degeneration. Table 8.1 summarizes the properties of CS-macromolecules

synthesized via both the “grafting-to” and “grafting-through” strategies as they relate to

the criteria for a functional biomimetic aggrecan macromolecule and as they compare to

natural aggrecan and chondroitin sulfate. CS-PAA macromolecules were lacking in their

CS grafting density, resulting in widely spaced CS chains, thereby not fulfilling a key

functional requirement of a biomimetic aggrecan. A major limitation in our study of the

“grafting-to” technique was the utilization of a covalent binding chemistry that has a

short reactivity window. Higher grafting density may be achieved if a backbone is

utilized which has reactive groups that can remain reactive over a long period of time,

thereby allowing dense brush formation. In our studies, we have found that the epoxy

amine reaction remains reactive through the full 96h time frame investigated. Therefore,

a poly-epoxide polymeric backbone is recommended for further investigation of the

“grafting-to” strategy of bottle brush polymer synthesis. In order for the poly-epoxide to

be applicable in the current reaction conditions (pH 9.4 aqueous reactions), the polymer

must be water soluble and epoxides must have high reactivity to the CS primary amine.

The utilization of step-growth polymerization was investigated due to the innate

ability it affords in controlling monomer spacing. In step-growth polymerization of CS

with di-epoxides, the spacing between CS chains is directly controlled by the MW or

Page 314: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

288

length of the di-epoxide monomer. EG-CS and PEG-CS macromolecules had low

average molecular weight compared to that required for a biomimetic aggrecan

replacement. Studies conducted here demonstrated that the low degree of polymerization

seen in EG-CS and PEG-CS macromolecules was likely due to low reactivity of the EG-

DGE and PEG-DGE monomers to the CS primary amine. CS-macromolecule synthesis

through the step-growth polymerization strategy may be improved by utilizing different

reactive chemistries for the step growth reaction. In the presented studies we

demonstrated that the epoxy-primary amine reaction can go to complete reaction of the

CS-primary amine in excess of epoxide. This epoxy-primary amine reaction may be

utilized to functionalize the terminal end of CS with a bi-functional step growth moiety,

followed by step-growth polymerization of bi-functional CS with another bi-functional

monomer in order to improve the step-growth polymerization technique and generate

dense CS brush structures. Although, EG-CS and PEG-CS MW was relatively low,

structures larger than CS were observed and with improvements to the step-growth

synthesis strategy, PEG based macromolecules may be candidate materials for a

biomimetic aggrecan replacement of PGs in the NP. This is further supported by the

positive functional properties of the EG-CS and PEG-CS macromolecules including high

fixed charge density, high osmotic potential, low cytotoxicity and low cost. CS-

macromolecules had similar properties to that of natural CS.

To stabilize the disc early in the degenerative cascade, we have designed a

biomimetic aggrecan material for injection to the nucleus pulposus to enhance the

mechanical, osmotic and hydration potential of the tissue while also serving to enhance

the intradiscal pressure, thus “re-inflating the flat tire”. This approach is also intended to

Page 315: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

289

mechanically protect the annulus fibrosus from abnormally high stresses which may be

responsible for the formation of tears and fissures. In this study, we have synthesized, for

the first time, hybrid bio/synthetic CS macromolecules with properties necessary for the

restoration of proteoglycan function in the nucleus pulposus of the intervertebral disc. In

this work we have set a stable foundation for the synthesis and characterization of

biomimetic aggrecan macromolecules via both the “grafting-to” and “grafting-through”

techniques of brush polymer synthesis. From these studies, several synthesis strategies

may be approached for the synthesis of large molecular weight densely grafted CS-

macromolecules for the restoration of NP function.

This work is the first proposal of an enzymatically resistant hybrid bio/synthetic

biomimetic aggrecan macromolecule for treatment of early degenerative changes

to the intervertebral disc

8.2 Novel Contributions

Prior to this study, natural chondroitin sulfate had not been utilized in the

synthesis of biomimetic macromolecules

This is the first report of the CS terminal primary amine being utilized for the

immobilization of CS into brush structures

This is the first report of the synthesis of vinyl terminated CS

This is the first report of the synthesis of CS macromolecules via the “grafting-to”

strategy (CS-PAA macromolecules)

Page 316: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

290

This is the first report of the synthesis of CS-macromolecules via a step-growth

“grafting-through” strategy (EG-CS and PEG-CS macromolecules)

This is the first account of the effect of a synthetic polymeric linker

(poly(ethylene glycol)) on CS-macromolecule function.

Page 317: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

291

Table 8.1: Properties of “Grafting-to” and “Grafting-through” CS-macromolecules as they relate to the criteria for a functional biomimetic aggrecan macromolecule

(red: does not meet criteria, green: may meet criteria)

Page 318: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

292

Notes for Table 8.1

1 Aggrecan fragments in the range of 100kDa to as low as 55kDa have been found in NP

tissues at both early (Grade 2) and advanced (Grade 4) stages of degeneration (78, 243).

Therefore it is reasonable to assume that similarly sized biomimetic aggrecan would

remain immobilized in the NP matrix

2Reference (63)

3Reference (139)

4,7,8Reference (7)

5Reference (89)

6Reference (69)

9Aggrecan osmotic pressure was extrapolated from aggrecan osmotic pressure

measurements obtained by equilibrium dialysis against PVA gels of known osmotic

pressure (280)

10CS osmotic pressure was predicted using an exponential viral coefficients reported by

Chahine et al (233) in 𝜋 ≈ 𝑐1𝑐𝑓 + 𝑐2(𝑐𝑓)2.

11Aggrecan in solution self assemble with increasing concentration and generate loose

gel-like associations (280)

12CS and CS-macromolecule solutions were found to be injectable through a 16 gauge

needle up to 1g/ml (solutions took on a paste like consistency). Data collected by

Biomed Sr. Design Team 2010-2011 (Data not shown).

Page 319: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

293

Chapter 9 : Future Work and Recommendations

As discussed previously, this study aimed to develop a novel biomimetic aggrecan

macromolecule for the early-interventional treatment of low back pain. Several

advancements were made through these investigations in the synthesis and

characterization of an enzymatically resistant, hybrid bio/synthetic biomimetic CS

macromolecule. The following are recommendations for future work in order to address

the previously stated limitations of the investigated synthesis techniques as well as

suggestions for future investigations into the structure and function of biomimetic

aggrecans.

9.1.1. “Grafting-To” Synthesis Strategies

9.1 Recommendations for Biomimetic Aggrecan Synthesis

The main limitation of the “grafting-to” strategy investigated in this study was the

limited grafting density achieved. In the “grafting-to” synthesis strategy, a key factor in

achieving high CS grafting density is allowing a long enough reaction window in order to

transition through the various regimes of brush synthesis kinetics (i.e. mushroom,

sparsely grafted, densely grafted) (171-173, 252). The epoxide-amine chemistry which

has a relatively long reaction window, has been suggested for use in the “grafting-to”

synthesis strategy. CS chain grafting to poly-epoxide backbones via the CS terminal

primary amine may be investigated. Possible epoxy backbones include poly(glycidyl

Page 320: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

294

methacrylate), polyglycerol polyglycidyl ether and sorbitol polyglycidyl ether (234,

281).

9.1.2. Step-Growth “Grafting-Through” Synthesis Strategies

The main limitation of the step-growth “grafting-through” synthesis strategy

investigated in this study was the low degree of polymerization achieved, likely due to

low reactivity of the diglycidyl ether epoxide to the CS terminal primary amine.

Although it may be possible to increase epoxy-amine reactivity by using a different form

of diglycidyl ether and further optimizing reaction conditions (i.e. adjusting the solvent

system and solution pH), the utilization of the CS terminal amine as a direct reactive

group in polymerization, limits the potential of this strategy. Reaction kinetics will

continue to depend on CS primary amine reaction kinetics followed by the reaction

kinetics of the generated secondary amine.

Step-growth polymerization strategies however are desirable due to the high

control over CS spacing possible as well as mild aqueous reaction conditions associated

with this technique. It is therefore recommended that CS-macromolecule synthesis

through step-growth polymerization be further investigated using different reactive

chemistries for the step growth reaction. The epoxy-primary amine reaction may be

utilized to functionalize the terminal end of CS with a bi-functional step growth moiety,

followed by step-growth polymerization of bi-functional CS with another bi-functional

monomer resulting in linear brush polymers. As an example, trimethylolpropane

triglycidyl ether (TMP), a tri-epoxide molecule, may be used to introduce two epoxide

Page 321: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

295

molecules to the terminal end of CS (i.e. reaction of the CS terminal amine with one

epoxide of the TMP). Di-epoxide CS-TMP monomers can then be reacted with di-amine

monomers (i.e. Poly(ethylene glycol) bis(amine) (PEG-bis-amine)), again utilizing the

more facile epoxide-primary amine reaction. Resulting brush CS-macromolecules would

still incorporate a PEG spacer, however reaction kinetics would only depend on TMP and

PEG-bis-amine epoxy-primary amine reactions. Reaction kinetics could be modulated by

varying both tri-epoxide and di-amine chemistry and structure. Alternatively, highly

chemo-selective reactive chemistries such as the azide–alkyne click reactions may be

utilized as step-growth reactive monomers (118, 282-283).

9.1.3. Alternative “Grafting-Through” Synthesis Strategies: Free-Radical

Polymerization

In this study, CS macromonomers were synthesized via the reaction of the CS

terminal primary amine with the epoxide of allyl glycidyl ether (AGE), resulting in vinyl

terminated CS. AGE- CS may be polymerized via free-radical polymerization using

either classical or controlled methods of APS/TMEDA radical initiation or cyanoxyl-

mediated initiation respectively (123, 126). Co-polymerization of AGE-CS with other

vinyl monomers such as acrylic acid would allow for the control of CS spacing in the

final brush macromolecule and result in the incorporation of a synthetic polymeric

backbone (i.e. poly(acrylic acid)) linking CS chains. In this technique must be taken to

protect from possible side reactions and CS degradation. Additionally, degree of

polymerization may be limited.

Page 322: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

296

9.1.4. Alternative Biomimetic Aggrecan Structures

In the case that high molecular weight densely grafted linear brush CS-

macromolecules are difficult to achieve, alternative macromolecule geometries may be

considered. For example, “grafting-to” of amine terminated CS to epoxy functionalized

nanoparticles or dendrimers may allow for the synthesis of densely grafted CS

nanoparticles or star-like polymers. These structures, although they do not mimic the

linear bottle brush structure of aggrecan, would be large in molecular size and still

maintain close packing of CS chains, as well as the osmotic and biological function of

CS. Mechanical consequence of the particle like configuration of CS would need to be

investigated as well as cytocompatibility.

As larger MW CS-macromolecules are synthesized and families of CS-

macromolecules with varying architecture are designed, studies may be conducted on the

effect of CS brush architecture (i.e. CS spacing, backbone rigidity, and macromolecule

length) on macromolecule properties. Previously, investigations into these effects were

only based on theoretical modeling, approximations of CS interactions based on CS

immobilization onto flat substrates or via the synthesis of model synthetic polyelectrolyte

brushes (69, 71, 73, 284). Structure-function investigations are of utmost importance

since changes in aggrecan synthesis and architecture have been observed with aging and

degeneration (93); however, the direct consequences of these changes on tissue function

9.2 Recommendations for Biomimetic Aggrecan Characterization

Page 323: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

297

are still largely unknown. In addition, biomimetic aggrecan structural properties may be

optimized through structure-function analysis to impart the greatest effect on NP

mechanics.

In order to draw the relationship between biomimetic aggrecan function and structure,

it will be critical to accurately characterize the macromolecule size and structure. To

further characterize biomimetic aggrecan size, dynamic light scattering (DLS) techniques

are suggested (241). Solvent systems developed in this study for GPC analysis will

provide a good starting point for CS-macromolecule solubilization for DLS analysis. In

DLS analysis it will be important to maintain high enough macromolecule concentration

for detection while avoiding aggregation. For further characterization of biomimetic

aggrecan structure, visualization with atomic force microscopy is recommended due to

the high resolution of this imaging technique and its prior utilization for aggrecan

imaging (75). In addition, characterization of biomimetic aggrecan hydrodynamic

properties via viscosity and osmotic measurements in varying conditions (i.e. varying

concentrations and no-salt/varying salt conditions) are recommended for the

determination of biomimetic aggrecan conformation and brush properties (205, 261, 263,

265-267, 269).

In this study, key functional characteristics of the synthesized CS-macromolecules

were investigated including fixed charge density, osmotic potential and cytotoxicity.

These characteristics are critical for the potential function of the CS-macromolecules in a

9.3 Recommendations for Biomimetic Aggrecan Functional Characterization

Page 324: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

298

biomimetic strategy for the restoration of NP function. In addition to the direct properties

of CS-macromolecules, it will be important to their function with respect to the cells and

tissue structures of the intervertebral disc and spinal unit. CS is often used as a

component in pharmaceutical osteoarthritis treatments, and CS has been demonstrated to

promote aggrecan synthesis in chondrocyte cultures while decreasing MMP expression

(89, 278). These properties of CS may allow CS-macromolecules to not only have a

mechanical benefit to the NP but also a regenerative one. It is therefore recommended

that the biosynthetic activity of biomimetic aggrecan macromolecules be investigated.

Investigations into aggrecan synthesis and MMP gene expression profiles may be

conducted on 3-D nucleus pulposus cell cultures or alternatively on whole intervertebral

disc organ cultures where appropriate oxygen tensions and culture pH is maintained (255,

285).

In addition to the biosynthetic activity of CS, the distribution of CS in NP tissue as

well as the resulting mechanical effect on spinal column function should be investigated.

Finite element modeling has demonstrated the benefit of the restoration of FCD on stress

distributions in the IVD; however, these results are based on the assumption that the

replaced FCD is physiologically distributed throughout the IVD matrix and permanently

integrated into the matrix (270). The FCD restoration model also does not take into

account the impact of electrostatic repulsive forces (generated by closely arranged fixed

charges) and their contribution to NP tissue mechanical properties. A mechanically

loaded cadaveric tissue model, consisting of an intact intervertebral disc isolated between

two vertebrae (anterior column unit) may be used for monitoring biomimetic aggrecan

distribution in the disc as well as the contribution of a biomimetic aggrecan injection to

Page 325: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

299

disc height, stiffness, and pressurization (286). Changes in these properties of the

intervertebral disc have been demonstrated previously to be associated with disc

degeneration and may be associated with back pain (2). In order to properly assess

biomimetic aggrecan distribution, and distinguish it from natural aggrecan, the

biomimetic aggrecan may be tagged for in situ monitoring. For example, for distribution

analysis via microcomputed tomography (micro-CT) a radiological contrast agent such as

meglumine diatrizoate may be covalently bound to oxidized chondroitin sulfate in the

CS-macromolecule in order to enable post-injection visualization of the macromolecule

in the IVD. Techniques such this will allow for the monitoring of biomimetic aggrecan

distribution in IVD tissue with concurrent analysis of disc mechanical properties, thereby

describing the functional impact of the biomimetic aggrecan on IVD function. In a

tagging technique, the influence of the tag on biomimetic aggrecan structure and function

must be monitored and minimized.

The recommendations for future work presented here should aid in the synthesis of a

biomimetic aggrecan macromolecule that meets all of the criteria set forth in this study.

Improvements to the synthesis technique have been suggested along with methods for

structural and functional characterization of the novel biomimetic biomacromolecules.

Functional studies which monitor the impact of biomimetic aggrecan injection into the

IVD will aid in the translation of this technology to animal models and the clinical

setting.

A vast number of additional applications are also foreseen for the CS-

macromolecules designed in this study. Aggrecan itself has been implicated in other

degenerative processes including osteoarthritis and biomimetic aggrecan may be useful in

Page 326: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

300

treatment strategies for this and other degenerative conditions. In addition to mimicing

aggrecan, the techniques developed here will allow for the synthesis of other

proteoglycan mimetics with varied GAG grafting density and core length (i.e. versican).

Any GAG demonstrated to have a reactive terminal primary amine may be utilized in the

synthesis of biomimetic proteoglycans. For example, studies in our lab have already

demonstrated that dermatan sulfate, an isomer of CS also has a terminal primary amine

and can be covalently bound to synthetic polymeric backbones. These new classes of

proteoglycan mimetics may be useful as versican replacements in skin tissue in order to

restore tissue resilience and act as scaffolding in skin tissue engineering (8). Versican

has also been found to decrease in the intervertebral disc, making this proteoglycan

another target for disc treatment (91). With the large diversity of proteoglycans and their

implication in biological function, the techniques and materials developed in this study

can have far reaching effects in restorative and regenerative medicine.

Page 327: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

301

Appendix 1: Select Literature Review of Key Relevant Technologies

A Brief Review of Key Advancements in GAG/Charge Density restorative/utilization

strategies

Growth factor injections into the IVD. The growth factor, osteogenic protein-1 (OP-1,

also known as BMP-7) has been investigated as a possible intradiscal injection treatment

for the restoration of GAG content and hydration in degenerated discs. OP-1 has been

injected into rabbit stab models for degenerative disc disease and disc height, hydration

as well as proteoglycan content was found to be restored after 6 weeks and maintained

through the investigation period of 24 weeks (287). In addition, discs isolated after OP-1

injection (100µg/10µl, 8 weeks), were found to have higher elastic modulus as well as

viscous modulus than those injected with control solutions (lactose, 10µl) (288). These

results were concurrent with an increase in proteoglycan content in the NP and AF as

well as collagen content in the NP (288). Application of OP-1 to degenerated NP discs

was demonstrated to have an anabolic effect on cells which also manifested as

biomechanical restoration. These results however have not been demonstrated in human

degenerated discs. Cellular response in rabbit discs may be very different from that in

humans as rabbit discs contain a large number of notochordal cells that have metabolic

activities different from that of the human adult NP cells (289).

Key Advancement: Induced increase in NP PG content with concurrent increase in disc

height, hydration, and mechanical properties.

Page 328: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

302

Limitation in function as restorative treatment for PG loss in NP: Reliant on NP cell

activity and nutrient supply which is compromised in an adult human disc.

Negatively charged Hydrogels(290). Negatively charged polyelectrolyte gels are

charged polymer networks with ions fixed on the polymer chains and counter-ions

localized withing the network frame. Generally polyelectrolyte gels are able to absorb

significant amounts of water (up to ~2000 times the polymer weight) without dissolving

in water itself (290). In addition, these gels can respond to external stimuli such as

changes to pH, ionic strength and electrical stimuli. Generally, hydrogels made from

either natural or synthetic sources, especially polyelectrolyte gels, have low mechanical

strength, however, gels with double network structures (i.e. composed of two

independently cross-linked networks) have been demonstrated have fracture strengths as

high as several tens of megapascals, comparable to that of articular cartilage (290).

Double network gels (DN gels) composed of anionic poly 2-acrylamido-2-methly-1-

propanesulfonic acid (PAMPS) and neutral poly-acrylamide) PAAm sustain higher

compression while PAMPS gels alone break down easily (290). PAMPS gels were

broken at stress of 0.4 MPa, while DN gel sustained stresses of 17.2 MPa. These

materials, due to their high strength, low surface friction and high water content have

been suggested for articular cartilage and other bio-tissue replacements (290).

Key Advancement: Mechanically strong hydrogel with high water content and

immobilized fixed charges (anionic groups).

Page 329: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

303

Limitation in function as restorative treatment for PG loss in NP: Synthetic system which

does not incorporate biologically active components and also does not maintain critical

arrangement of charges

Chondroitin Sulfate Containing Hydrogels. Chondroitin sulfate has been applied to

many biomaterial and biological applications as a bioactive material, a natural polymer

scaffold for tissue engineering and as a carrier polymer network for cell therapies. CS

scaffolds have also been applied in cartilage tissue regeneration by combining CS with

biodegradable and natural components (291). In addition, hydrogels composed of

collagen/HA and CS-6 have been synthesized as scaffolds for NP tissue engineering

(292). These systems were primarily investigated for their biological function.

The CS-PEO(polyethylene oxide) cross-liked hydrogel system has been

investigated for mechanical properties as well as physiochemical properties (293).

Hydrogels were synthesized by the modification of the CS carboxylic acid with adipic

dihydrazide followed by modification with acrylate groups (Figure 18)(293). Acrylated-

CS was then crosslinked with PEO-hexa thiols solution where CS-PEO hydrogels formed

spontaneously via Michael type addition reactions without further treatment. CS-PEO

hydrogel swelling was measure in water and at different pHs at 37̊C for up to 24h.

Compression strength of the hydrogels was investigated with an apparatus (MT-LQ

material tester), 5kg load cell) which applied a circular probe (~12mm in diameter) onto

the surface of the hydrogel moving at a rate of 0.2mm/sec (293). CS-PEO hydrogels had

high swelling capacity. 5% CS-PEO hydrogel swelled to 170% in water at pH 2, 224%

Page 330: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

304

in pH 4, 324% in pH 7.4, and 525% in pH 9, while the 10% hydrogel swelled to 166% in

pH 2, 218% in pH 4, 319% in pH 7.4, and 456% in pH 9 (293). Hydrogel fabricated with

lower concentration imbibed higher water content and swelling increased when the water

environment was changed from pH 2 to 9. The higher swelling under basic condition was

attributed to the repulsive force between the negatively charged carboxylic acid groups of

the chondroitin sulfate network at pH 9. At 10% CS-PEO gel concentrations, the fully

hydrated hydrogel had an approximately 3Pa yield strength (293). This is a low value

compared to the 50kPa compressive modulus generally estimated to be required for the

restoration of normal compressive disc mechanics (as determined by cadaveric testing

and mathematical modeling (43).

In a separate study, CS hydrogel was applied as an intradiscal injection to the

rabbit degenerated disc stab model (294). Cross-linked CS hydrogels were formed by

first conjugating cinnamic acid to the carboxyls groups of CS then exposing the CS

derivative to UV light for crosslinking (5.0 wt%) (294). Intradiscal injection of sodium

hyaluronate, cross-linked hyaluronate hydrogels and cross linked CS-hydrogels (300µl)

were conducted two weeks after puncture injury. Before therapeutic injections, a

reduction in signal intensity on T2-weighted MR images was detected, indicating loss of

water content in the disc. One month after injection, all IVDs maintained normal

intensity levels. Three months after injection only IVDs injected with cross-linked

hyaluronate hydrogels or cross-linked CS hydrogels showed normal MR imaging signal

intensities. After six months, all IVDs lost signal intensity (294).

Key Advancement: Fabrication of biocompatible and regenerative chondroitin sulfate

hydrogels and application of cross-linked CS hydrogels to IVD restoration.

Page 331: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

305

Limitation in function as restorative treatment for PG loss in NP: Low mechanical

properties of CS hydrogels or otherwise uncharacterized. CS arrangement in the bottle

brush structure not maintained in the hydrogel configuration.

Aggrecan Containing Hydrogels (295). Agarose hydrogels (2% low melting

temperature) have been synthesized with variable ratios of aggrecan (generally 2mg/ml),

hyaluronan and link protein embedded in the agarose matrix (295). Aggrecan aggregates

were assembled in situ and resulting hydrogels were investigated for aggrecan retention

as well as mechanical properties and biological function. Compressive properties of the

constructs increased with the content of aggrecan retained (constructs soaked and agitated

in PBS, ph 7.0 for 3 days) in the matrix with equilibrium stress increasing from ~18kPa

in control agarose gels to ~20kPa in aggrecan containing gels and ~30 in highly

aggregating aggrecan hydrogels under 90% compressive strain (constructs placed

between two porous platens filled with PBS and tested at 30-90% compressive strains of

initial thickness to equilibrium relaxation times) (295). Transport of the aggrecan

components from the 2% agarose gel was consistent with diffusion processes and on the

basis of the hydrodynamic radius estimations for the aggrecan and aggregates (295).

Aggregate assembly was compatible with chondrocytes and chondrocytes maintained

epression of the chondrogenic phenotype (i.e. secreting type II collagen) (295). These

studies demonstrated the biomechanically relevant retention of aggrecan in hydrogel

constructs and their potential as scaffolds for cartilage tissue engineering.

Page 332: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

306

Key Advancement: Fabrication of biocompatible hydrogel with embedded and retained

aggrecan.

Limitation in function as restorative treatment for PG loss in NP: Incorporation of

natural aggrecan which is susceptible to proteolytic degradation. Function in IVD has

not been demonstrated.

Summary. CS hydrogels have been formed via the cross-linking of CS into networks.

CS cross-linking and immobilization is primarily obtained by the modification of the CS

backbone and CS is immobilized in random arrangements. CS hydrogels have shown

highly promising biological properties in both cartilage and NP tissue engineering

applications. The mechanical properties and physical properties of CS hydrogels have

not however been thoroughly demonstrated. Synthetic polyelectrolyte hydrogels may

provide an alternative more mechanically stable hydrogel for replacement of fixed

charges in the IVD however their application as an injectable integrative material for NP

stabilization has not been demonstrated and charges remain unorganized.

A Brief Review of Key Advancements in the Synthesis of Biomimetic Aggrecan

Macromolecule: Literature Landscape

A polymer bottle/cylindrical brush is defined as an assembly of polymer chains which are

densely tethered by one end to a polymeric backbone (106). As a result of high steric

crowding, the chains are forced to stretch away from the surface to avoid overlap,

elongating the polymer backbone and stiffening the molecule (106, 296). Synthetic

routes for the preparation of polymer brushes include: (1) “grafting-to”, (2) “grafting-

Page 333: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

307

through” and (3) “grafting-from”. Glycopolymers can be defined as synthetic polymers

possessing a non-carbohydrate backbone that carry carbohydrate (sugar) moieties as

pendant groups (106, 296).

Synthetic polysulfonated comb aggrecan mimetic polymers. Synthetic polysulfated

cylindrical bottle brush polymers have been synthesized using atom transfer radical

polymeriziation (ATRP) techniques (297). Styrene sulfonated ethyl ester (SSE) and

styrene sulfonated dodecyl ester (SSD) polymer brushes were synthesized using the

“grafting-from” technique after first synthesizing the styrene sulfonate ethyl ester

macroinitiator (Figure 19) (117, 297). This method was advantageous as incomplete

sulfation, crosslinking, and degradation are avoided in comparison to polymer-analogous

sulfonation (297). Brushes with backbone length of approximately 600 and 1,600 repeat

units and grafting densities from 10 to 40% and 10 to 100% for SSD and SSE polymers

respectively were synthesized (297). Molar mass determinations were made by analytical

ultracentrifugation, GPC-MALLS and 1H-NMR spectroscopy (SSD polymers could not

be analyzed via GPC due to lack of elution of this polymer). Brush like structure was

confirmed via AFM and TEM imaging and AFM measurements revealed that polymer

length increased with increasing grafting density (Figure 19) (297). Both polymers were

presented as intermediates to obtaining corresponding styrene sulfonated polyelectrolyte

brushes for model compounds mimicking the polyelectrolyte brushes of human cartilage

(297). No studies have been presented however on the application of these polymer

brushes to biological systems. Similar sulfated comb polymers however have been

compared to natural aggrecan regarding solution properties and were found to exhibit

similar viscosities and shear moduli to that of natural aggrecan (240, 263-264, 284).

Page 334: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

308

Key Advancement: Synthesis of synthetic densly grafted high fixed charge density

cylindrical bottle brushes with solution properties similar to natural aggrecan.

Limitation in function as restorative treatment for PG loss in NP: Biocompatibility has

not been demonstrated and does not incorporate biologically active chondroitin sulfate

chains.

“Grafting-To” click chemistry polymerization of glycopolymer brushes (298-299). A

library of glycopolymers was synthesized through the “grafting-to” method of polymer

synthesis utilizing facile “click” chemistry reactions(299). Click chemistry describes

several classes of chemical reactions which have very high efficiency in conversion and

selectivity (i.e. stereoselective reactions, excellent functional group compatibility) under

mild reaction conditions (298, 300). These features have allowed for the tailor-made

synthesis of complex materials including dendrimers, bioconjugates, therapeutics,

functionalized polymers and sugar derivatives(299). First a clickable alkyne polymeric

backbone was synthesized via ATRP (atom transfer radical polymerization) (Figure 20).

Mannose- and galactose based azidosugar derivatives were then synthesized (Figure 21)

in order to develop glycopolymers with concavalin A lectin binding ability (299).

Concavalin A is an α-mannose-binding lectin involved in a number of biological

processes. Click reaction of the sugar azides to the alkyne polymer was achieved using

[(PPh3)3CuBr] as the catalyst in the presence of DIPEA (Figure 21)(299). After

conjugation, an increase in the polymer hydrodynamic volume (size exclusion

chromatography) was observed with little change in the molecular weight distribution.

Page 335: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

309

Complete (~100%) conversion of the alkyne groups into triazoles was demonstrated with

1H-NMR and FT-IR analysis and final mannose to galactose ratios in to polymer were

similar to the intial feed ratios (299). Glycopolymers synthesized in this study were

demonstrated to successfully precipitate Concavalin A present in solution (i.e. high

binding) (299).

Key Advancement: Synthesis of biologically active high grafting density glycopolymer

brushes via facile and selective reaction chemistries without protection of sugar

functional groups.

Limitation in function as restorative treatment for PG loss in NP: Requires complex

synthesis techniques to generate the polymeric backbone as well as clickable sugars.

Has not been demonstrated with high MW sugars such as chondroitin sulfate.

“Grafting-From” synthesis of cylindrical glycopolymers (“Molecular Sugar Sticks”)

(116). Glycocylindrical brushes have been synthesized with poly-(3-O-methacryloyl-α,β-

D-glucopyranose (PMAGlc) side chains using the “grafting from” approach via atom

transfer radical polymerization (ATRP) of the protected monomer, 3-O-methacryloyl-

1,2:5,6-di-O-isopropylidiene-D-glucofuranose (MAIGlc) (Figure 22). After deprotection,

water soluble glycopolymer brushes were obtained. For synthesis, CuBr and MAIGlc

were combined in ethyl acetate and added to polyinitiator PBIEM-II (poly(2-(2-

bromoisobutyryloxy)ethyl methacrylate). HMTETA was added to initiate

polymerization. Monomer conversion was detected by H-NMR to be 10%. Solution was

dissolved in THF, passed through a silica column, and precipitated from THF into

Page 336: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

310

methanol, then precipitated into petroleum ether to remove unreacted monomer (116).

Products had a Mn of 58.6E4 and a Mw/Mn = 1.07 with conventional GPC using PS

calibration and Mn = 1120E4 and Mw/Mn = 1.09 using GPC with multiangle light

scattering detector measurements (116). Products were then deprotected to transform

poly(MAIGlc) to poly(MPAGlc) using mildly acidic conditions. Polymer was then

purified by dailysis against 1000MWCO membrane. Solution properties of the polymer

were investigated using dynamic light scattering and polymers were visualized using

scanning force microscopy and cryo-TEM. Polymers were confirmed to have stretched

wormlike structure and grafting efficiency was 0.20 < f < 0.40 (initiation efficiencies, i.e.

not as densely spaced as those for some PS and polyacrylate brushes) (116). Behavior

was found to be typical for cylindrical brushes rather than comb-shaped polymers.

Investigations were reported to be in progress for biological and medicinal application

but no such reports have been made yet (116).

Key Advancement: “grafting-from” synthesis of densly grafted glycopolymer brushes

with bottle brush properties.

Limitation in function as restorative treatment for PG loss in NP: Utilizes synthetic

sugars and requires complex synthesis technique including protection and de-protection

of sugar functional groups. Length of sugar chains limited.

“Grafting-Through” controlled free radical polymerization of glycopolymers. Grande

and Chaikof et al. polymerized model vinyl derivatized sulfated glycomonomers using

both the cyanoxyl mediated and APS/TMEDA mediated polymerizations (123). The

Grande group has also synthesized several other mono- and di-saccharide GAG-mimetic

Page 337: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

311

polymers using these methods (123-126, 301). Vinyl terminated glycomonomers were

synthesized from N-acetyl-D-glucosamine by treatment with 4-penten-1-ol or 10-

undecene-1-ol in the presence of 10-camphorsulfonic acid (CSA) as catalyst (123)

(Figure 23). These two vinyl monomers provided varying spacing length between the

sugar and vinyl moieties of the glycomonomers. For controlled free-radical

polymerization of the glycomonomers (Figure 23), Cyanoxyl radicals were generated in

situ by an electron-transfer reaction between cyanate anions and p-

chlorobenzenediazonium cations (prepared in water through the diazotization reaction of

p-chloroanaline) (123). The purified, degassed, vinyl-CS glycomonomer were then

added to the reaction medium along with acrylamide as a spacer, and sodium cyanate.

Polymerization was carried out at 50°C and reactions were isolated by precipitation in

10-fold excess cold methanol. Weight proportions of sugar residues reached as high as

50%, however, carbohydrate content in the resulting polymer was also associated with an

increase in polydispersity index, indicating loss of control over the copolymerization

process in the presence of increasing glycomonomer (123). Additionally, it was noted

that shorter spacer arm length actually resulted in increased incorporation of carbohydrate

(123). Biological application of these polymers was suggested by the authors to lie in

areas of wound repair and tissue regeneration (123). Specifically, glycopolymers were

prepared for anti-coagulant, heparin mimetic applications (124).

Key Advancement: Synthesis of sulfated glycopolymer brushes in mild conditions without

protection of functional groups.

Page 338: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

312

Limitation in function as restorative treatment for PG loss in NP: Complex synthesis of

glycomonomers. Some limitations in glycosylation and high MW polymer generation.

Has not been demonstrated with chondroitin sulfate or other similarly sized GAGs.

Step-growth polymerization of glycoproteins. Glycoproteins have been synthesized via

step-growth polymerization for use as antifreeze glycoprotein analogues (302-303).

Step-growth polymerization reactions can be used to produce glycopolymers which have

structural homogeneity while chain length is not as well controlled (compared to

controlled free radical polymerization strategies) (118). In the synthesis of antifreeze

glycoproteins, Ala-Thr-Ala peptide sequences are first synthesized coupled to mono and

disaccharide at the Thr residue (O-linked). The glycopeptides monomers are then linked

in a polycondensation reaction (step-growth polymerization) using conventional DPPA

(diphenylphosphoryl azide) or other peptide coupling methods (303) (Figure 24). These

methods have also been used to produce sulfated glycopolypeptides (303). Functional

properties of these glycoproteins have been investigated with regard to function as an

antifreeze agent (i.e. thermal hysteresis and the effect of glycopolypeptide on ice crystal

formation). Glycopolypeptides with as little as two repeat units (i.e.dimers)were found to

have effective antifreeze properties (304) and the observed effects were found to be

highly dependent on the primary structure of the polypeptide backbone (118, 304).

Key Advancement: Facile step-growth synthesis of amino acid terminated sugars to

generate synthetic glycoproteins demonstrated to have similar properties to their natural

counterparts.

Page 339: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

313

Limitation in function as restorative treatment for PG loss in NP: Synthetic

glycoproteins will have protein backbones that may be susceptible to proteolytic

degradation. Synthesis with large MW sugar side chains (i.e. Chondroitin sulfates) has

not been demonstrated.

Summary. In recent reviews of glycopolymers and cylindrical bottle brush synthesis and

applications (106, 118, 298, 300, 305-306), advancements have been made in the

synthetic strategies for brush and glycopolymer synthesis, resulting in more controlled

synthesis techniques as well as higher degrees of glycosylation. In addition advances in

polymer end-functionalization, synthesis of complex architectures, polymer

characterization, and glycomonomer synthesis has been noted. Advancements have also

been made in the application of glycopolymers in the area of therapeutics, particularly as

HIV and Alzheimer’s therapeutics as well as polymers for drug delivery (307). No

reports have been made however on the incorporation of chondroitin sulfate in

glycopolymer synthesis and the direct utilization of aggrecan and glycosaminoglycan

mimetics in the restoration of tissue mechanical function has not yet been investigated.

Page 340: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

314

Figure 18: Schematics of CS-acrylate synthesis by sequential grafting of adipic dihydrazide and then acrylic acid to CS (293).

Page 341: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

315

Figure 19: Synthesis schematic of Styrene sulfonated ethyl ester (SSE) and styrene sulfonated dodecyl ester (SSD) polymer brushes and resulting polymer brush structure as seen with atomic

force microscopy imaging. (297)

Page 342: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

316

Figure 20: Synthesis of poly-alkyne click-able polymer from protected a protected alkyne functional monomer (trimethylsilyl mechacrylate). (299)

Page 343: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

317

Figure 21: Synthesis of click-able sugars (sugar containing azide groups) and their “Grafting-to” click conjugation to an alkyne polymeric backbone. (298-299)

Page 344: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

318

Figure 22: Synthesis schematic for fabrication of “molecular sugar sticks” using the “grafting from” approach via atom transfer radical polymerization (ATRP) of the protected monomer

(bottom) scanning force microscopy height image of molecular sugar sticks. (116)

Page 345: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

319

Figure 23: Synthesis of sulfated and non-sulfated vinyl glycomonomers and subsequent cyanoxyl mediated free-radical polymerization for the synthesis of biomimetic polymers with chemically stable hydrocarbon backbones and biologically active pendant hydrophilic saccharides. (123)

Page 346: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

320

Figure 24: Polycondensation step-growth polymerization of glycopeptides for the synthesis of sulfated glycopolypeptides as antifreeze glycoprotein analogs (303).

Page 347: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

321

Appendix 2: Alternate views of Chondroitin Sulfate Disaccharide with Linkage Region and Terminal Amino Acid Residue (demonstrating glycosidic linkages)

10-50

OO

OH

NH2

O

O

O

OH

OH OH

OO

O

SO

OO

-

OH

NHO

CH3

OH

Tetrasaccharide

Linkage Region

GalN GlcN

40O

N

H

HOH

O

Page 348: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

322

List of References

1. AAOS. The Burden of Musculoskeletal Disease in the United States. Rosemont, IL: American Academy of Orthopaedic Surgeons; 2008. 2. Adams MA. Biomechanics of back pain. Acupuncture in Medicine. 2004;22(4):178. 3. Olczyk K. Age-related changes in proteoglycans of human intervertebral discs. Zeitschrift für Rheumatologie(Print). 1994;53(1):19-25. 4. Urban JPG, McMullin JF. Swelling Pressure of the Lumbar Intervertebral Discs: Influence of Age, Spinal level, Composition, and Degeneration. Spine. 1988;13(2):179-87. 5. Boswell MV, Trescot A, Datta S, Schultz D, Hansen H, Abdi S, et al. Interventional techniques: evidence-based practice guidelines in the management of chronic spinal pain. 2007. 6. Deyo RA, Gray DT, Kreuter W, Mirza S, Martin BI. United States trends in lumbar fusion surgery for degenerative conditions. Spine (Phila Pa 1976). 2005 Jun 15;30(12):1441-5; discussion 6-7. 7. Kiani C, Liwen C, Yao Jiong WU, Yee AJ, Burton BY. Structure and function of aggrecan. Cell Research. 2002;12(1):19-32. 8. Ferdous Z, Grande-Allen K. Utility and control of proteoglycans in tissue engineering. Tissue Engineering. 2007;13(8):1893-904. 9. Joshi A, Massey CJ, Karduna A, Versilovic E, Marcolongo M. The Effect of Nucleus Implant Parameters on the Compressive Mechanics of the Lumbar Intervertebral Disc: A Finite Element Study. J Biomed Mater Res (Appl Biomater). Submitted. 10. Joshi A, Fussell G, Thomas J, Hsuan A, Lowman A, Karduna A, et al. Functional compressive mechanics of a PVA/PVP nucleus pulposus replacement. Biomaterials. 2006;27:176-84. 11. Ho E, Lowman A, Marcolongo M. Synthesis and characterization of an injectable hydrogel with tunable mechanical properties for soft tissue repair. Biomacromolecules. 2006;7:3223-8. 12. Marcolongo MS, Cannella M, Massey CJ. Nucleus Replacement of the Intervertebral Disc. In: Kurtz SM, Edidin AA, editors. Spine Technology Handbook. New York, NY: Academic Press; 2006. p. 281-300.

Page 349: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

323

13. Joshi A, Mehta S, Vresilovic E, Karduna A, Marcolongo M. Nucleus implant parameters significantly change the compressive stiffness of the human lumbar intervertebral disc. Journal of Biomechanical Engineering. 2005 June;127(3):536-40. 14. Thomas J, Lowman A, Marcolongo M. Novel associated hydrogels for nucleus pulposus replacement. Journal of Biomedical Materials Research. 2003;67A:1329-37. 15. Thomas J, Gomes K, Lowman A, Marcolongo M. The effect of dehydration history on PVA/PVP hydrogels for nucleus pulposus replacement. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2004;69B:135-40. 16. White AA, Panjabi MM. Clinical Biomechanics of the Spine. 2nd ed: Lippincott Co.; 1990. 17. Errington RJ, K. Puustjarvi, I.R.F. White, S. Roberts, J.P.G. Urban. Characterization of Cytoplasm-Filled Processes in Cells of the Intervertebral Disc. J Anat. 1998;192:369-78. 18. Raj PP, Fipp A. Intervertebral Disc: Anatomy-Physiology-Pathophysiology-Treatment. Pain Practice. 2008;8(1):18. 19. Bron J, Helder M, Meisel H-J, Van Royen B, Smit T. Repair, regenerative and supportive therapies of the annulus fibrosus: achievements and challenges. European Spine Journal. 2009;18(3):301-13. 20. Adams M, Roughley P. What is intervertebral disc degeneration, and what causes it? Spine. 2006;31(18):2151. 21. Roughley PJ. Biology of intervertebral disc aging and degeneration: involvement of the extracellular matrix. Spine (Phila Pa 1976). 2004 Dec 1;29(23):2691-9. 22. Le Maitre C, Pockert A, Buttle D, Freemont A, Hoyland J. Matrix synthesis and degradation in human intervertebral disc degeneration. Biochemical Society Transactions. 2007;35:652-5. 23. Urban JPG, Smith S, Fairbank JCT. Nutrition of the intervertebral disc. Spine. 2004;29(23):2700. 24. Olczyk K. Age-related changes in glycosaminoglycans of human intervertebral discs. Folia histochemica et cytobiologica/Polish Academy of Sciences, Polish Histochemical and Cytochemical Society. 1993;31(4):215. 25. Olczyk K. Age-related changes in collagen of human intervertebral disks. Gerontology. 1992;38:196-204.

Page 350: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

324

26. Olczyk K. Age-related changes of elastin content in human intervertebral discs. Folia histochemica et cytobiologica/Polish Academy of Sciences, Polish Histochemical and Cytochemical Society. 1994;32(1):41. 27. Carragee EJ. Persistent Low Back Pain. N Engl J Med. 2005 May 5, 2005;352(18):1891-8. 28. Masuda K, Lotz JC. New challenges for intervertebral disc treatment using regenerative medicine. Tissue Eng Part B Rev. 2010 Feb;16(1):147-58. 29. Adams MA, Dolan P, McNally DS. The internal mechanical functioning of intervertebral discs and articular cartilage, and its relevance to matrix biology. Matrix Biol. 2009 Sep;28(7):384-9. 30. Antoniou J, Steffen T, Nelson F, Winterbottom N, Hollander AP, Poole RA, et al. The human lumbar intervertebral disc: evidence for changes in the biosynthesis and denaturation of the extracellular matrix with growth, maturation, ageing, and degeneration. Journal of Clinical Investigation. 1996;98(4):996. 31. Leckie S, Kang J. Recent advances in nucleus pulposus replacement technology. Current Orthopaedic Practice. 2009;20(3):222. 32. Henschke N, Kuijpers T, Rubinstein SM, van Middelkoop M, Ostelo R, Verhagen A, et al. Injection therapy and denervation procedures for chronic low-back pain: a systematic review. European Spine Journal. 2010;19(9):1425-49. 33. Relief P, Covington L, Conn A. Systematic review of caudal epidural injections in the management of chronic low back pain. Pain Physician. 2009;12(1):109-35. 34. Bao Q, Mccullen G, Higham P, Dumbleton J, Yuan H. The artificial disc: theory, design and materials. Biomaterials. 1996;17(12):1157-67. 35. White A, Lawrence J, Grauer J. 24 Intervertebral Disc Arthroplasty as an Alternative to Spinal Fusion: Rationale and Biomechanical and Design Considerations. Spinal Reconstruction: Clinical Examples of Applied Basic Science, Biomechanics and Engineering. 2007:263. 36. Martz EO, Goel VK, Pope MH, Park JB. Materials and design of spinal implants - A review. Journal of Biomedical Materials Research. 1997;38(3):267-88. 37. Rutherford E, Tarplett L, Davies E, Harley J, King L. Lumbar Spine Fusion and Stabilization: Hardware, Techniques, and Imaging Appearances1. Radiographics. 2007;27(6):1737. 38. Kuslich SD, Bagby G, editors. The BAK interbody fusion system: early clinical results of treatment for chronic low back pain1993.

Page 351: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

325

39. Cho D-Y, Lee W-Y, Sheu P-C. Treatment of multilevel cervical fusion with cages. Surgical Neurology. 2004;62(5):378-85. 40. Büttner-Janz K, Hochschuler S, McAfee P. The artificial disc. Berlin: Springer; 2003. 41. Kurtz SM, Edidin AA. Spine technology handbook: Academic Press; 2006. 42. Joshi A, Massey CJ, Karduna A, Vresilovic E, Marcolongo M. The effect of nucleus implant parameters on the compressive mechanics of the lumbar intervertebral disc: A finite element study. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2009. 43. Thomas JD, Fussell G, Sarkar S, Lowman AM, Marcolongo M. Synthesis and recovery characteristics of branched and grafted PNIPAAm-PEG hydrogels for the development of an injectable load-bearing nucleus pulposus replacement. Acta Biomaterialia. 2010;6:1319-28. 44. Di Martino A, Vaccaro AR, Lee JY, Denaro V, Lim MR. Nucleus pulposus replacement: basic science and indications for clinical use. Spine. 2005;30(16S):S16. 45. Bao QB, Yuan HA. New technologies in spine: nucleus replacement. Spine. 2002;27(11):1245. 46. Coric D, Mummaneni PV. Nucleus replacement technologies. J Neurosurg Spine. 2008 Feb;8(2):115-20. 47. Coric D, Mummaneni P. Nucleus replacement technologies. Journal of Neurosurgery: Spine. 2008;8(2):115-20. 48. Bao QB, Songer M, Pimenta L, Werner D, Reyes-Sanchez A, Balsano M, et al. Nubac disc arthroplasty: preclinical studies and preliminary safety and efficacy evaluations. SAS Journal. 2007;1(1):36-45. 49. Shim CS, Lee SH, Park CW, Choi WC, Choi G, Choi WG, et al. Partial disc replacement with the PDN prosthetic disc nucleus device: early clinical results. Journal of spinal disorders & techniques. 2003;16(4):324. 50. D'Souza W, Neen D, Birch N. "THE LAW OF UNINTENDED CONSEQUENCES" IN NEW SPINAL TECHNOLOGY: DOES THE FAILURE OF THE PDN SOLO SUGGEST THE NEED FOR A EUROPEAN EQUIVALENT OF THE FDA? J Bone Joint Surg Br. 2009 September 1, 2009;91-B(SUPP_III):479-. 51. Ratner BD, Hoffman AS, Schoen FJ, Lemons JE, editors. Biomaterials Science - An Introduction to Materials in Medicine. San Diego: Academic Press; 1996.

Page 352: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

326

52. Yuksel KU, Walsh SP, Black KS, inventors; U.S. Pat. App. #20,050,102,030, assignee. IN SITU BIOPROSTHETIC FILLER AND METHOD, PARTICULARLY FOR THE IN SITU FORMATION OF VERTEBRAL DISC BIOPROSTHETICS2005. 53. Singhal V, MacEachern CF, Craig NJA, Wardlaw D. Early clinical results of an in situ polymerizing protein hydrogel nuclear repair system. Biospine; 2006. 54. O'Halloran DM, Pandit AS. Tissue-engineering approach to regenerating the intervertebral disc. Tissue Engineering. 2007;13(8):1927-54. 55. Yang X, Li X. Nucleus pulposus tissue engineering: a brief review. European Spine Journal. 2009;18:1564-157. 56. Woods BI, Sowa G, Vo N, Kang JD. A Change in Strategy: The Use of Regenerative Medicine and Tissue Engineering to Augment the Course of Intervertebral Disc Degeneration. Operative Techniques in Orthopaedics. [doi: 10.1053/j.oto.2009.10.009]. 2010;20(2):144-53. 57. Hsieh AH, Yoon ST. Update on the pathophysiology of degenerative disc disease and new developments in treatment strategies. Open Access Journal of Sports Medicine. 2010;1:191-9. 58. Boyd LM, Carter AJ. Injectable biomaterials and vertebral endplate treatment for repair and regeneration of the intervertebral disc. European Spine Journal. 2006;15:414-21. 59. Hardingham T, Fosang A. Proteoglycans: many forms and many functions. The FASEB journal. 1992;6(3):861. 60. Iozzo RV. Matrix proteoglycans: from molecular design to cellular function. Annual Review of Biochemistry. 1998;67(1):609-52. 61. Johnstone B, Bayliss MT. The large proteoglycans of the human intervertebral disc: changes in their biosynthesis and structure with age, topography, and pathology. Spine. 1995;20(6):674. 62. Nap RJ, Szleifer I. Structure and Interactions of Aggrecans: Statistical Thermodynamic Approach. Biophysical Journal. 2008;95(10):4570-83. 63. Watanabe H, Yamada Y, Kimata K. Roles of aggrecan, a large chondroitin sulfate proteoglycan, in cartilage structure and function. Journal of biochemistry. 1998;124(4):687. 64. Dudhia J. Aggrecan, aging and assembly in articular cartilage. Cellular and molecular life sciences. 2005;62(19):2241-56.

Page 353: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

327

65. Muir H. Molecular approach to the understanding of osteoarthrosis. Ann Rheum Dis. 1977;36:199-208. 66. Yanagishita M. Function of proteoglycans in the extracellular matrix. Acta pathologica japonica. 1993;43(6):283. 67. Vertel BM. The ins and outs of aggrecan. Trends in Cell Biology. 1995;5(12):458-64. 68. Roughley P. Biology of intervertebral disc aging and degeneration: involvement of the extracellular matrix. Spine. 2004;29(23):2691. 69. Seog J, Dean D, Plaas AHK, Wong-Palms S, Grodzinsky AJ, Ortiz C. Direct measurement of glycosaminoglycan intermolecular interactions via high-resolution force spectroscopy. Macromolecules. 2002;35(14):5601-15. 70. Buschmann MD, Grodzinsky AJ. A molecular model of proteoglycan-associated electrostatic forces in cartilage mechanics. J Biomech Eng. 1995;117(2):179-92. 71. Dean D, Seog J, Ortiz C, Grodzinsky AJ. Molecular-level theoretical model for electrostatic interactions within polyelectrolyte brushes: applications to charged glycosaminoglycans. Langmuir. 2003;19(13):5526-39. 72. Eisenberg S, Grodzinsky A. Swelling of articular cartilage and other connective tissues: electromechanochemical forces. Journal of Orthopaedic Research. 2005;3(2):148-59. 73. Seog J, Dean D, Rolauffs B, Wu T, Genzer J, Plaas AHK, et al. Nanomechanics of opposing glycosaminoglycan macromolecules. Journal of biomechanics. 2005;38(9):1789-97. 74. Dean D, Han L, Grodzinsky AJ, Ortiz C. Compressive nanomechanics of opposing aggrecan macromolecules. Journal of biomechanics. 2006;39(14):2555-65. 75. Ng L, Grodzinsky AJ, Patwari P, Sandy J, Plaas A, Ortiz C. Individual cartilage aggrecan macromolecules and their constituent glycosaminoglycans visualized via atomic force microscopy. Journal of Structural Biology. 2003;143(3):242-57. 76. Roberts S, Evans EH, Kletsas D, Jaffray DC, Eisenstein SM. Senescence in human intervertebral discs. European Spine Journal. 2006;15:312-6. 77. Zhao CQ, Wang LM, Jiang LS, Dai LY. The cell biology of intervertebral disc aging and degeneration. Ageing Research Reviews. 2007;6(3):247-61.

Page 354: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

328

78. Chujo T, An HS, Masuda K, Patel KP, Sandy JD, Akeda K, et al. Aggrecanases and aggrecanase-generated fragments in the human intervertebral disc at early and advanced stages of disc degeneration. Spine. 2007;32(23):2596. 79. Roberts S, Caterson B, Menage J, Evans EH, Jaffray DC, Eisenstein SM. Matrix metalloproteinases and aggrecanase: their role in disorders of the human intervertebral disc. Spine. 2000;25(23):3005. 80. Goupille P, Jayson Iv M, Valat JP, Freemont AJ. Matrix metalloproteinases: the clue to intervertebral disc degeneration? Spine. 1998;23(14):1612. 81. Ireland D. Molecular mechanisms involved in intervertebral disc degeneration and potential new treatment strategies. Bioscience Horizons. 2009 March 2009;2(1):83-9. 82. Verzijl N, DeGroot J, Bank RA, Bayliss MT, Bijlsma JWJ, Lafeber FPJG, et al. Age-related accumulation of the advanced glycation endproduct pentosidine in human articular cartilage aggrecan: the use of pentosidine levels as a quantitative measure of protein turnover. Matrix Biology. 2001;20(7):409-17. 83. Urban J, Smith S, Fairbank J. Nutrition of the intervertebral disc. Spine. 2004;29(23):2700. 84. Urban JPG, Holm SH. Intervertebral Disc Nutrition as Related to Spinal Movements and Fusion. In: AR H, editor. Tissue Nutrition and Viability. New York: Springer-Verlag; 1986. p. 101-19. 85. Perea W, Cannella M, Yang J, Vega A, Polenova T, Marcolongo M. 2H double quantum filtered (DQF) NMR spectroscopy of the nucleus pulposus tissues of the intervertebral disc. Magnetic Resonance in Medicine. 2007;57(6):990-9. 86. Roberts SP, Urban JPGP, Evans HB, Eisenstein SMPF. Transport Properties of the Human Cartilage Endplate in Relation to Its Composition and Calcification. Spine. 1996 February;21(4):415-20. 87. Legendre F, Bauge C, Roche R, Saurel A, Pujol J. Chondroitin sulfate modulation of matrix and inflammatory gene expression in IL-1 [beta]-stimulated chondrocytes-study in hypoxic alginate bead cultures1. Osteoarthritis and Cartilage. 2008;16(1):105-14. 88. Nishimoto S, Takagi M, Wakitani S, Nihira T, Yoshida T. Effect of chondroitin sulfate and hyaluronic acid on gene expression in a three-dimensional culture of chondrocytes. Journal of bioscience and bioengineering. 2005;100(1):123-6. 89. Volpi N. Chondroitin sulfate: structure, role and pharmacological activity: Academic Press; 2006.

Page 355: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

329

90. Yoon J, Halper J. Tendon proteoglycans: biochemistry and function. J Musculoskelet Neuronal Interact. 2005;5(1):22-34. 91. Cs-Szabo G, Ragasa-San Juan D, Turumella V, Masuda K, Thonar EJM, An HS. Changes in mRNA and protein levels of proteoglycans of the anulus fibrosus and nucleus pulposus during intervertebral disc degeneration. Spine. 2002;27(20):2212. 92. Volpi N, Mucci A, Schenetti L. Stability studies of chondroitin sulfate. Carbohydrate Research. 1999;315(3-4):345-9. 93. Scott J, Bosworth T, Cribb A, Taylor J. The chemical morphology of age-related changes in human intervertebral disc glycosaminoglycans from cervical, thoracic and lumbar nucleus pulposus and annulus fibrosus. Journal of Anatomy. 1994;184(Pt 1):73. 94. Lamberg SI, Stoolmiller AC. GLYCOSAMINOGLYCANS. A BIOCHEMICAL AND CLINICAL REVIEW. J Investig Dermatol. 1974;63(6):433-49. 95. Lauder RM. Chondroitin sulphate: A complex molecule with potential impacts on a wide range of biological systems. Complementary Therapies in Medicine. 2009;17(1):56-62. 96. Volpi N, Maccari F. Quantitative and Qualitative Evaluation of Chondroitin Sulfate in Dietary Supplements. Food Analytical Methods. 2008;1(3):195-204. 97. van Blitterswijk W, van de Nes J, Wuisman P. Glucosamine and chondroitin sulfate supplementation to treat symptomatic disc degeneration: Biochemical rationale and case report. BMC Complementary and Alternative Medicine. 2003;3(1):2. 98. Bryant SJ, Davis-Arehart KA, Luo N, Shoemaker RK, Arthur JA, Anseth KS. Synthesis and Characterization of Photopolymerized Multifunctional Hydrogels: Water-Soluble Poly(Vinyl Alcohol) and Chondroitin Sulfate Macromers for Chondrocyte Encapsulation. Macromolecules. 2004;37(18):6726-33. 99. Strehin I, Nahas Z, Arora K, Nguyen T, Elisseeff J. A versatile pH sensitive chondroitin sulfate-PEG tissue adhesive and hydrogel. Biomaterials. 2010;31(10):2788-97. 100. Piai J, Rubira A, Muniz E. Self-assembly of a swollen chitosan/chondroitin sulfate hydrogel by outward diffusion of the chondroitin sulfate chains. Acta Biomaterialia. 2009;5(7):2601-9. 101. Dawlee S, Sugandhi A, Balakrishnan B, Labarre D, Jayakrishnan A. Oxidized Chondroitin Sulfate-Cross-Linked Gelatin Matrixes: A New Class of Hydrogels. Biomacromolecules. 2005;6(4):2040-8.

Page 356: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

330

102. Li Q, Williams CG, Sun DDN, Wang J, Leong K, Elisseeff JH. Photocrosslinkable polysaccharides based on chondroitin sulfate. Journal of Biomedical Materials Research. 2004;68(1):28-33. 103. Stuart K, Panitch A. Characterization of gels composed of blends of collagen I, collagen III, and chondroitin sulfate. Biomacromolecules. 2009;10(1):25-31. 104. Bryant S, Davis-Arehart K, Luo N, Shoemaker R, Arthur J, Anseth K. Synthesis and characterization of photopolymerized multifunctional hydrogels: water-soluble poly (vinyl alcohol) and chondroitin sulfate macromers for chondrocyte encapsulation. Macromolecules. 2004;37(18):6726-33. 105. Kuijpers AJ, van Wachem PB, van Luyn MJA, Brouwer LA, Engbers GHM, Krijgsveld J, et al. In vitro and in vivo evaluation of gelatin-chondroitin sulphate hydrogels for controlled release of antibacterial proteins. Biomaterials. 2000;21(17):1763-72. 106. Sheiko SS, Sumerlin BS, Matyjaszewski K. Cylindrical molecular brushes: Synthesis, characterization, and properties. Progress in Polymer Science. 2008;33(7):759-85. 107. Zhang M, Muller AHE. Cylindrical polymer brushes. JOURNAL OF POLYMER SCIENCE PART A POLYMER CHEMISTRY. 2005;43(16):3461. 108. Gao H, Matyjaszewski K. Synthesis of molecular brushes by" grafting onto" method: Combination of ATRP and click reactions. Journal of the American Chemical Society. 2007;129(20):6633-9. 109. Ito K. Polymeric design by macromonomer technique. Progress in Polymer Science. 1998;23(4):581-620. 110. Zhang D, Ortiz C. Single Macromolecule Nanomechanical Design: Poly (2-hydroxyethyl methacrylate-g-ethylene glycol) Graft Copolymers of Varying Architecture. Macromolecules. 2005;38(6):2535-9. 111. Zhang D, Ortiz C. Synthesis and Single Molecule Force Spectroscopy of Graft Copolymers of Poly (2-hydroxyethyl methacrylate-g-ethylene glycol). Macromolecules. 2004;37(11):4271-82. 112. Zhang D, Macias C, Ortiz C. Synthesis and Solubility of (Mono-) End-Functionalized Poly (2-hydroxyethyl methacrylate-g-ethylene glycol) Graft Copolymers with Varying Macromolecular Architectures. Macromolecules. 2005;38(6):2530-4. 113. Ladmiral V, Melia E, Haddleton DM. Synthetic glycopolymers: an overview. European Polymer Journal. 2004;40(3):431-49.

Page 357: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

331

114. Okada M. Molecular design and syntheses of glycopolymers. Progress in Polymer Science. 2001;26(1):67-104. 115. Lutz JF, Börner HG. Modern trends in polymer bioconjugates design. Progress in Polymer Science. 2008;33(1):1-39. 116. Muthukrishnan S, Zhang M, Burkhardt M, Drechsler M, Mori H, Muller AHE. Molecular sugar sticks: Cylindrical glycopolymer brushes. Macromolecules. 2005;38(19):7926-34. 117. Lienkamp K, Schnell I, Groehn F, Wegner G. Polymerization of Styrene Sulfonate Ethyl Ester by ATRP: Synthesis and Characterization of Macromonomers for Suzuki Polycondensation. Macromolecular Chemistry and Physics. 2006;207(22):2066-73. 118. Spain SG, Matthew IG, Neil RC. Recent advances in the synthesis of well-defined glycopolymers. Journal of Polymer Science Part A: Polymer Chemistry. 2007;45(11):2059-72. 119. Cunliffe D, Pennadam S, Alexander C. Synthetic and biological polymers--merging the interface. European Polymer Journal. 2004;40(1):5-25. 120. Fraser C, Grubbs RH. Synthesis of Glycopolymers of Controlled Molecular Weight by Ring-Opening Metathesis Polymerization Using Well-Defined Functional Group Tolerant Ruthenium Carbene Catalysts. Macromolecules. 1995;28(21):7248-55. 121. Albertin L, Stenzel MH, Barner-Kowollik C, Davis TP. Effect of an added base on (4-cyanopentanoic acid)-4-dithiobenzoate mediated RAFT polymerization in water. Polymer. 2006;47(4):1011-9. 122. Kohji O, Yoshinobu T, Takeshi F. Synthesis of a well-defined glycopolymer by atom transfer radical polymerization. Journal of Polymer Science Part A: Polymer Chemistry. 1998;36(14):2473-81. 123. Grande D, Baskaran S, Baskaran C, Gnanou Y, Chaikof EL. Glycosaminoglycan-mimetic biomaterials. 1. Nonsulfated and sulfated glycopolymers by cyanoxyl-mediated free-radical polymerization. Macromolecules. 2000;33(4):1123-5. 124. Sun XL, Grande D, Baskaran S, Hanson SR, Chaikof EL. Glycosaminoglycan mimetic biomaterials. 4. Synthesis of sulfated lactose-based glycopolymers that exhibit anticoagulant activity. Biomacromolecules. 2002;3(5):1065-70. 125. Baskaran S, Grande D, Sun XL, Yayon A, Chaikof EL. Glycosaminoglycan-Mimetic Biomaterials. 3. Glycopolymers Prepared from Alkene-Derivatized Mono-and Disaccharide-Based Glycomonomers. Bioconjugate Chem. 2002;13(6):1309-13.

Page 358: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

332

126. Grande D, Baskaran S, Chaikof EL. Glycosaminoglycan mimetic biomaterials. 2. Alkene-and acrylate-derivatized glycopolymers via cyanoxyl-mediated free-radical polymerization. Macromolecules. 2001;34(6):1640-6. 127. Flitsch SL. Chemical and enzymatic synthesis of glycopolymers. Current Opinion in Chemical Biology. [doi: DOI: 10.1016/S1367-5931(00)00152-6]. 2000;4(6):619-25. 128. Varma AJ, Kennedy JF, Galgali P. Synthetic polymers functionalized by carbohydrates: a review. Carbohydrate Polymers. [doi: DOI: 10.1016/j.carbpol.2004.03.007]. 2004;56(4):429-45. 129. Anderson B, Hoffman P, Meyer K. The O-Serine Linkage in Peptides of Chondroitin 4- or 6-Sulfate. Journal of Biological Chemistry. 1965;240(1):156-67. 130. Mattern KJ, Deen WM. Binding of glycosaminoglycans to cyano-activated agarose membranes: kinetic and diffusional effects on yield and homogeneity. Carbohydrate Research. 2007;342(15):2192-201. 131. Atlas SJ, Deyo RA. Evaluating and managing acute low back pain in the primary care setting. J Gen Intern Med. 2001 Feb;16(2):120-31. 132. Medtech-Insights. Report # A606 - U.S. Surgical Procedure Volumes: Windhover Information2007 February 2007. 133. Weiler C, Nerlich A, Zipperer J, Bachmeier B, Boos N. 2002 SSE Award Competition in Basic Science: expression of major matrix metalloproteinases is associated with intervertebral disc degradation and resorption. European Spine Journal. 2002;11(4):308-20. 134. Olczyk K. Age-related changes in proteoglycans of human intervertebral discs. Z Rheumatol. 1994 Jan-Feb;53(1):19-25. 135. Hutton W, Elmer W, Boden S, Horton W, Carr K. Analysis of chondroitin sulfate in lumbar intervertebral discs at two different stages of degeneration as assessed by discogram. Journal of Spinal Disorders & Techniques. 1997;10(1):47. 136. Urban JP, McMullin JF. Swelling pressure of the inervertebral disc: influence of proteoglycan and collagen contents. Biorheology. 1985;22(2):145-57. 137. Roughley PJ, Lee ER. Cartilage proteoglycans: structure and potential functions. Microscopy research and technique. 1994;28(5):385-97. 138. Roughley PJ, Melching LI, Heathfield TF, Pearce RH, Mort JS. The structure and degradation of aggrecan in human intervertebral disc. European Spine Journal. 2006;15:326-32.

Page 359: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

333

139. Volpi N. Analytical aspects of pharmaceutical grade chondroitin sulfates. Journal of Pharmaceutical Sciences. 2007;96(12):3168-80. 140. 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 1):6-13. 141. McAlindon TE, LaValley MP, Gulin JP, Felson DT. Glucosamine and chondroitin for treatment of osteoarthritis: a systematic quality assessment and meta-analysis. Jama. 2000;283(11):1469. 142. Nishimura M, Yan W, Mukudai Y, Nakamura S, Nakamasu K, Kawata M, et al. Role of chondroitin sulfate-hyaluronan interactions in the viscoelastic properties of extracellular matrices and fluids. Biochim Biophys Acta. 1998 Mar 12;1380(1):1-9. 143. Hui JH, Chan SW, Li J, Goh JCH, Li L, Ren XF, et al. Intra-articular delivery of chondroitin sulfate for the treatment of joint defects in rabbit model. Journal of molecular histology. 2007;38(5):483-9. 144. Kjellen L, Lindahl U. Proteoglycans: Structures and Interactions. Annual Review of Biochemistry. 1991;60(1):443-75. 145. Udenfriend S, Stein S, Bohlen P, Dairman W, Leimgruber W, Weigele M. Fluorescamine: a reagent for assay of amino acids, peptides, proteins, and primary amines in the picomole range. Science. 1972;178:871-2. 146. Held PG. Fluorometric Quantitation of Protein using the Reactive Compound Fluorescamine. Application Note. BioTek Instruments Company 2001. 147. De Bernardo S, Weigele M, Toome V, Manhart K, Leimgruber W, Böhlen P, et al. Studies on the reaction of fluorescamine with primary amines. Archives of Biochemistry and Biophysics. [doi: 10.1016/0003-9861(74)90490-1]. 1974;163(1):390-9. 148. Wasteson Å. A method for the determination of the molecular weight and molecular-weight distribution of chondroitin sulphate. Journal of Chromatography A. 1971;59(1):87-97. 149. Hermanson GT. Bioconjugate Techniques. Second ed. Pierce Biotechnology TFS, editor. Rockford, IL, USA: Academic Press; 2008. 150. ThermoScientific. AminoLink Reductant. Instructions. Rockford, IL: Pierce Biotechnology: Thermo Fisher Scientific; 2007. 151. ThermoScientific. EDC Water-soluble carbodiimide crosslinker for zero-length, carboxyl-to-amine conjugation. Rockford, IL: Thermo Fisher Scientific Inc; [cited 2011]; Available from: http://www.piercenet.com/browse.cfm?fldID=02030312.

Page 360: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

334

152. Toida T, Kakinuma N, Toyoda H, Imanari T. 1 H-NMR Profile of Glycosaminoglycans in Human Urine. Analytical Sciences. 1994;10(4):537-41. 153. Toida T, Toyoda H, Imanari T. High-resolution proton nuclear magnetic resonance studies on chondroitin sulfates. Analytical Sciences. 1993;9(1):53-8. 154. Cabassi F, Casu B, Perlin AS. Infrared absorption and raman scattering of sulfate groups of heparin and related glycosaminoglycans in aqueous solution. Carbohydrate Research. [doi: 10.1016/S0008-6215(00)80924-6]. 1978;63:1-11. 155. Mathews MB. Isomeric Chondroitin Sulphates. Nature. [10.1038/181421a0]. 1958;181(4606):421-2. 156. Orr SFD. Infra-red spectroscopic studies of some polysaccharides. Biochimica et Biophysica Acta. [doi: 10.1016/0006-3002(54)90156-0]. 1954;14:173-81. 157. Foot M, Mulholland M. Classification of chondroitin sulfate A, chondroitin sulfate C, glucosamine hydrochloride and glucosamine 6 sulfate using chemometric techniques. Journal of Pharmaceutical and Biomedical Analysis. [doi: 10.1016/j.jpba.2005.01.026]. 2005;38(3):397-407. 158. Huang JM, Huang HJ, Wang YX, Chen WY, Chang FC. Preparation and characterization of epoxy/polyhedral oligomeric silsesquioxane hybrid nanocomposites. Journal of Polymer Science Part B: Polymer Physics. 2009;47(19):1927-34. 159. Parker R, Isaacs N. Mechanisms of epoxide reactions. Chemical reviews. 1959;59(4):737-99. 160. Ganesh N. Biomimetic Aggrecan Based on a Polyacrylic Acid (PAA) Core and Chondroitin Sulfate (CS) Bristles [Master's Thesis]. Philadelphia, PA: Drexel University; 2010. 161. Marques MFV, Pinto PR, de Andrade CT, Michel RC. Synthesis and characterization of low molecular weight polyacrolein. Journal of Applied Polymer Science. 2009;112(3):1771-9. 162. Nakajima N, Ikada Y. Mechanism of amide formation by carbodiimide for bioconjugation in aqueous media. Bioconjugate chemistry. 1995;6(1):123-30. 163. Kalia J, Raines RT. Advances in bioconjugation. Current organic chemistry. 2010;14(2):138. 164. O'Shaughnessy B. Effect of Concentration on Reaction Kinetics in Polymer Solutions. Macromolecules. [doi: 10.1021/ma00092a030]. 1994;27(14):3875-84.

Page 361: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

335

165. Rühe J, Ballauff M, Biesalski M, Dziezok P, Gröhn F, Johannsmann D, et al. Polyelectrolyte brushes. Polyelectrolytes with Defined Molecular Architecture I. 2004:189-98. 166. Penn L, Huang H, Sindkhedkar M, Rankin S, Chittenden K, Quirk R, et al. Formation of tethered nanolayers: Three regimes of kinetics. Macromolecules. 2002;35(18):7054-66. 167. Advincula RC. Polymer brushes: synthesis, characterization, applications: Vch Verlagsgesellschaft Mbh; 2004. 168. Odian G. Principles of polymerization: Wiley-Interscience; 2004. 169. Shin H, Jo S, Mikos AG. Biomimetic materials for tissue engineering. Biomaterials. [doi: DOI: 10.1016/S0142-9612(03)00339-9]. 2003;24(24):4353-64. 170. Allcock HR, Lampe FW, Mark JE, Allcock H. Contemporary polymer chemistry: Prentice Hall; 1990. 171. Kato K, Uchida E, Kang ET, Uyama Y, Ikada Y. Polymer surface with graft chains. Progress in Polymer Science. 2003;28(2):209-59. 172. Huang H, Rankin S, Penn L, Quirk R, Cheong T. Transition from mushroom to brush during formation of a tethered layer. Langmuir: the ACS journal of surfaces and colloids. 2004;20(14):5770. 173. Penn LS, Heqing Huang, Roderic P. Quirk, and Tae H. Cheong. Kinetics of Polymer Brush Formation With and Without Segmental Adsorption. In: RC Advincula WB, KC Caster, J Rühe, editor. Polymer brushes: synthesis, characterization, applications: Wiley-VCH; 2004. p. 317-28. 174. Shechter L, Wynstra J, Kurkjy RP. Glycidyl ether reactions with amines. Industrial & Engineering Chemistry. 1956;48(1):94-7. 175. Pascault JP, Williams RJJ. General Concepts about Epoxy Polymers. 2010. 176. Swier S, Van Assche G, Van Mele B. Reaction kinetics modeling and thermal properties of epoxy–amines as measured by modulated temperature DSC. I. Linear step growth polymerization of DGEBA+ aniline. Journal of Applied Polymer Science. 2004;91(5):2798-813. 177. Pritchard JG, Long FA. Kinetics and Mechanism of the Acid-catalyzed Hydrolysis of Substituted Ethylene Oxides1. Journal of the American Chemical Society. [doi: 10.1021/ja01593a002]. 1956;78(12):2667-70.

Page 362: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

336

178. Shechter L, Wynstra J. Glycidyl ether reactions with alcohols, phenols, carboxylic acids, and acid anhydrides. Industrial & Engineering Chemistry. 1956;48(1):86-93. 179. Smith IT. The mechanism of the crosslinking of epoxide resins by amines. Polymer. [doi: 10.1016/0032-3861(61)90010-6]. 1961;2:95-108. 180. John NAS, George GA. Diglycidyl amine -- epoxy resin networks: Kinetics and mechanisms of cure. Progress in Polymer Science. [doi: 10.1016/0079-6700(94)90032-9]. 1994;19(5):755-95. 181. Rozenberg B. Kinetics, thermodynamics and mechanism of reactions of epoxy oligomers with amines. In: Dušek K, editor. Epoxy resins and composites II: Springer Berlin / Heidelberg; 1986. p. 113-65. 182. Mijovic J, Fishbain A, Wijaya J. Mechanistic modeling of epoxy-amine kinetics. 1. Model compound study. Macromolecules. 1992;25(2):979-85. 183. Mijovic J, Wijaya J. Reaction kinetics of epoxy/amine model systems. The effect of electrophilicity of amine molecule. Macromolecules. 1994;27(26):7589-600. 184. Rogers ME, Long TE, Turner SR. Introduction to Synthetic Methods in Step-Growth Polymers. Synthetic Methods in Step-Growth Polymers: John Wiley & Sons, Inc.; 2003. p. 1-16. 185. Swier S, Van Assche G, Van Mele B. Reaction kinetics modeling and thermal properties of epoxy-amines as measured by modulated-temperature DSC. I. Linear step-growth polymerization of DGEBA+ aniline. Journal of applied polymer science. 2004;91(5):2798-813. 186. Miller JS, Shen CJ, Legant WR, Baranski JD, Blakely BL, Chen CS. Bioactive hydrogels made from step-growth derived PEG-peptide macromers. Biomaterials. 2010;31(13):3736-43. 187. Gray D. Organic and Biomaterials Chemistry - Lecture 10. Cambridge, MA: Knol; 2010 [cited 2011 July, 8]. 188. Bonora GM, Drioli S. Reactive PEGs for protein conjugation. In: Veronese FM, editor. PEGylated Protein Drugs: Basic Science and Clinical Applications: Birkhäuser Basel; 2009. p. 33-45. 189. Patiny L. Database of NMR spectra: NMR Predictor. Lausanne Switzerland [cited 2011]; Available from: http://www.nmrdb.org. 190. Bergström K, Holmberg K, Safranj A, Hoffman AS, Edgell MJ, Kozlowski A, et al. Reduction of fibrinogen adsorption on PEG coated polystyrene surfaces. Journal of Biomedical Materials Research. 1992;26(6):779-90.

Page 363: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

337

191. Mucci A, Schenetti L, Volpi N. 1H and 13C nuclear magnetic resonance identification and characterization of components of chondroitin sulfates of various origin. Carbohydrate Polymers. 2000;41(1):37-45. 192. Perlin AS, Casu B, Sanderson GR. 220MHz spectra of heparin, chondroitins, and other mucopolysaccharides. Canadian Journal of Chemistry. 1970;48(14):2260-8. 193. Welti D, Rees DA, Welsh EJ. Solution Conformation of Glycosaminoglycans: Assignment of the 300-MHz 1H-Magnetic Resonance Spectra of Chondroitin 4-Sulphate, Chondroitin 6-Sulphate and Hyaluronate, and Investigation of an Alkali-Induced Conformation Change. European Journal of Biochemistry. 1979;94(2):505-14. 194. Scott JE, Heatley F, Wood B. Comparison of secondary structures in water of chondroitin-4-sulfate and dermatan sulfate: implications in the formation of tertiary structures. Biochemistry. 1995;34(47):15467-74. 195. Gatti G, Casu B, Torri G, Vercellotti JR. Resolution-enhanced 1H-n.m.r. spectra of dermatan sulfate and chondroitin sulfates: conformation of the uronic acid residues. Carbohydrate Research. [doi: 10.1016/S0008-6215(00)84070-7]. 1979;68(1):C3-C7. 196. Mansur HS, Oréfice RL, Mansur AAP. Characterization of poly(vinyl alcohol)/poly(ethylene glycol) hydrogels and PVA-derived hybrids by small-angle X-ray scattering and FTIR spectroscopy. Polymer. [doi: 10.1016/j.polymer.2004.08.036]. 2004;45(21):7193-202. 197. Poisson N, Lachenal G, Sautereau H. Near-and mid-infrared spectroscopy studies of an epoxy reactive system. Vibrational spectroscopy. 1996;12(2):237-47. 198. CUBoulder. IR Spectroscopy Tutorial: Amines. University of Colorado, Boulder, Chemistry and Biochemistry Department; 2011; Available from: http://orgchem.colorado.edu/hndbksupport/specttutor/amines.html. 199. Settle FA. Handbook of instrumental techniques for analytical chemistry: Prentice Hall PTR; 1997. 200. Wu C. Handbook of size exclusion chromatography and related techniques: Marcel Dekker, New York; 2004. 201. Jordi. White Paper: Gel permeation Chromatography or Size Exclusion Chromatography. wwwjordilabscom. 202. Volpi N, Bolognani L. Glycosaminoglycans and proteins: different behaviours in high-performance size-exclusion chromatography. Journal of Chromatography A. 1993;630(1-2):390-6.

Page 364: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

338

203. Johnson PY, Blake DA. Rapid size-exclusion chromatography of proteoglycans. Journal of Chromatography B: Biomedical Sciences and Applications. 1997;688(1):27-33. 204. Jordi. GPC Columns. Catalouge. Bellingham, MA: Jordi labs LLC. 205. Harding SE. The intrinsic viscosity of biological macromolecules. Progress in measurement, interpretation and application to structure in dilute solution. Progress in biophysics and molecular biology. 1997;68(2):207-62. 206. Wasteson Å. Properties of fractionated chondroitin sulphate from ox nasal septa. Biochemical Journal. 1971;122(4):477. 207. Sabaratnam S, Coleman PJ, Badrick E, Mason RM, Levick JR. Interactive effect of chondroitin sulphate C and hyaluronan on fluid movement across rabbit synovium. The Journal of Physiology. 2002;540(1):271. 208. Mijovic J, Andjelic S. A Study of Reaction Kinetics by Near-Infrared Spectroscopy. 1. Comprehensive Analysis of a Model Epoxy/Amine System. Macromolecules. [doi: 10.1021/ma00112a026]. 1995;28(8):2787-96. 209. Snavely DL, Dubsky J. Near-IR spectra of polyethylene, polyethylene glycol, and polyvinylethyl ether. Journal of Polymer Science Part A: Polymer Chemistry. 1996;34(13):2575-9. 210. Landau S, Friedman S, Devash L, Mabjeesh SJ. Polyethylene Glycol, Determined by Near-Infrared Reflectance Spectroscopy, as a Marker of Fecal Output in Goats. Journal of Agricultural and Food Chemistry. [doi: 10.1021/jf011346b]. 2002;50(6):1374-8. 211. Brashear RL, Flanagan DR, Luner PE, Seyer JJ, Kemper MS. Diffuse reflectance near-infrared spectroscopy as a nondestructive analytical technique for polymer implants. Journal of Pharmaceutical Sciences. 1999;88(12):1348-53. 212. Cowie JMKG. Polymers: chemistry and physics of modern materials: CRC Press; 1991. 213. Ninni L, Burd H, Fung WH, Meirelles AJA. Kinematic viscosities of poly (ethylene glycol) aqueous solutions. Journal of Chemical & Engineering Data. 2003;48(2):324-9. 214. Morpurgo M, Veronese F. Conjugates of peptides and proteins to polyethylene glycols. Methods in molecular biology (Clifton, NJ). 2004;283:45. 215. Bailon P, Berthold W. Polyethylene glycol-conjugated pharmaceutical proteins.

Page 365: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

339

Pharmaceutical Science & Technology Today. [doi: 10.1016/S1461-5347(98)00086-8]. 1998;1(8):352-6. 216. Polverari M, van de Ven TGM. Dilute Aqueous Poly(ethylene oxide) Solutions:  Clusters and Single Molecules in Thermodynamic Equilibrium. The Journal of Physical Chemistry. [doi: 10.1021/jp960215o]. 1996;100(32):13687-95. 217. Sasahara K, Uedaira H. Volume and compressibility changes on mixing aqueous solutions of the amino acid and poly (ethylene glycol). Colloid & Polymer Science. 1994;272(4):385-92. 218. Sasahara K, Uedaira H. Solubility of amino acids in aqueous poly (ethylene glycol) solutions. Colloid &amp; Polymer Science. 1993;271(11):1035-41. 219. Sasahara K, Sakurai M, Nitta K. Volume and compressibility changes for short poly (ethylene glycol)–water system at various temperatures. Colloid & Polymer Science. 1998;276(7):643-7. 220. Eliassi A, Modarress H, Mansoori GA. Measurement of Activity of Water in Aqueous Poly (ethylene glycol) Solutions (Effect of Excess Volume on the Flory-Huggins -Parameter). Journal of Chemical & Engineering Data. 1999;44(1):52-5. 221. Berman NS. Drag Reduction by Polymers. Annual Review of Fluid Mechanics. 1978;10(1):47-64. 222. Napper DH. Steric stabilization. Journal of Colloid and Interface Science. [doi: 10.1016/0021-9797(77)90150-3]. 1977;58(2):390-407. 223. Flammersheim HJ, Hörhold H, Bellstedt K, Klee J. DSC investigations on the kinetics of the linear step growth polymerization of dian-bisglycidylether with amines. Thermochimica Acta. [doi: 10.1016/0040-6031(85)85848-2]. 1985;92:185-8. 224. Gaggini F, Porcheddu A, Reginato G, Rodriquez M, Taddei M. Colorimetric tools for solid-phase organic synthesis. Journal of combinatorial chemistry. 2004;6(5):805-10. 225. Burguera J, Townshend R. Determination of amines by using a chemiluminescent reaction in solution. Talanta. 1979;26(9):795-8. 226. Nakamura H, Tamura Z. Fluorometric determination of secondary amines based on their reaction with fluorescamine. Analytical Chemistry. [doi: 10.1021/ac50063a023]. 1980;52(13):2087-92. 227. Siggia S, Hanna J, Kervenski I. Quantitative analysis of mixtures of primary, secondary, and tertiary aromatic amines. Analytical Chemistry. 1950;22(10):1295-7.

Page 366: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

340

228. Ruch JE, Critchfield FE. Determination of Small Amounts of Tertiary Amines in the Presence of Primary and Secondary Amines. Analytical Chemistry. 1961;33(11):1569-72. 229. Umbreit G. Spectrophotometric determination of secondary amines. Analytical Chemistry. 1961;33(11):1572-3. 230. Karweik DH, Meyers CH. Spectrophotometric determination of secondary amines. Analytical Chemistry. 1979;51(2):319-20. 231. Wagner CD, Brown RH, Peters ED. The Analysis of Aliphatic Amine Mixtures; Determination of Tertiary Aliphatic Amines in the Presence of Primary and Secondary Amines and Ammonia. Journal of the American Chemical Society. 1947;69(11):2609-10. 232. SpectrumLabs. Laboratory Dialysis Frequently Asked Questions. [cited 2011]; Available from: http://www.spectrumlabs.com/dialysis/FAQ.html. 233. Chahine NO, Chen FH, Hung CT, Ateshian GA. Direct measurement of osmotic pressure of glycosaminoglycan solutions by membrane osmometry at room temperature. Biophysical journal. 2005;89(3):1543-50. 234. Nishi C, Nakajima N, Ikada Y. In vitro evaluation of cytotoxicity of diepoxy compounds used for biomaterial modification. Journal of Biomedical Materials Research. 1995;29(7):829-34. 235. Li L. MALDI mass spectrometry for synthetic polymer analysis: John Wiley & Sons Inc; 2010. 236. Cael JJ, Isaac DH, Blackwell J, Koenig JL, Atkins EDT, Sheehan JK. Polarized infrared spectra of crystalline glycosaminoglycans. Carbohydrate Research. [doi: 10.1016/S0008-6215(00)83848-3]. 1976;50(2):169-79. 237. Wang S-C, Chen B-H, Wang L-F, Chen J-S. Characterization of chondroitin sulfate and its interpenetrating polymer network hydrogels for sustained-drug release. International Journal of Pharmaceutics. [doi: 10.1016/j.ijpharm.2006.08.041]. 2007;329(1-2):103-9. 238. Moller H, Heinegard D, Poulsen J. Combined alcian blue and silver staining of subnanogram quantities of proteoglycans and glycosaminoglycans in sodium dodecyl sulfate-polyacrylamide gels. Analytical Biochemistry. 1993;209(1):169-75. 239. McDevitt C, Muir H. Gel electrophoresis of proteoglycans and glycosaminoglycans on large-pore composite polyacrylamide-agarose gels. Analytical Biochemistry. 1971;44(2):612-22.

Page 367: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

341

240. Papagiannopoulos A, Waigh TA, Hardingham T, Heinrich M. Solution Structure and Dynamics of Cartilage Aggrecan. Biomacromolecules. 2006;7(7):2162-72. 241. Dumitriu S. Polysaccharides: structural diversity and functional versatility: CRC; 2005. 242. Hardingham T, Bayliss M. Proteoglycans of articular cartilage: Changes in aging and in joint disease. Seminars in Arthritis and Rheumatism. [doi: 10.1016/0049-0172(90)90044-G]. 1990;20(3, Supplement 1):12-33. 243. Sztrolovics R, Alini M, Roughley P, Mort J. Aggrecan degradation in human intervertebral disc and articular cartilage. Biochemical Journal. 1997;326(Pt 1):235. 244. Urban JP, McMullin JF. Swelling pressure of the lumbar intervertebral discs: influence of age, spinal level, composition, and degeneration. Spine (Phila Pa 1976). 1988 Feb;13(2):179-87. 245. Buckwalter JA. Aging and Degeneration of the Human Intervertebral Disc. Spine. 1995;20:1307-14. 246. Adams MA, Dolan P, McNally DS. The internal mechanical functioning of intervertebral discs and articular cartilage, and its relevance to matrix biology. Matrix Biology. 2009;28(7):384-9. 247. Joshi A VE, Cannella M, Arthur A, Karduna A, Marcolongo M. The Effect of Partial Denucleation on the Mechanical Behavior of the Intervertebral Disc. J Biomech. 2008;41(10):2014-111. 248. Sigma. A1906 Aggrecan from bovine articular cartilage lyophilized powder, sterile-filtered. St Louis, MO: Sigma-Aldrich Co; 2010 [cited 2010]; Available from: http://www.sigmaaldrich.com/united-states.html. 249. Liebscher T, Haefeli M, Wuertz K, Nerlich AG, Boos N. Age-Related Variation in Cell Density of Human Lumbar Intervertebral Disc. Spine. 2011;36(2):153-9 10.1097/BRS.0b013e3181cd588c. 250. Zalipsky S. Chemistry of polyethylene glycol conjugates with biologically active molecules. Advanced Drug Delivery Reviews. [doi: 10.1016/0169-409X(95)00023-Z]. 1995;16(2-3):157-82. 251. Urban JP, Maroudas A. Swelling of the intervertebral disc in vitro. Connect Tissue Res. 1981;9(1):1-10. 252. Jackson AR, Yuan TY, Huang CY, Gu WY. A Conductivity Approach to Measuring Fixed Charge Density in Intervertebral Disc Tissue. Ann Biomed Eng. 2009 Sep 11.

Page 368: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

342

253. Iatridis J, MacLean J, O'Brien M, Stokes I. Measurements of Proteoglycan and Water Content Distribution in Human Lumbar Intervertebral Discs. Spine. 2007;32(14):1493-7. 254. Mort J, Roughley P. Measurement of glycosaminoglycan release from cartilage explants. Arthritis Research: Methods in Molecular Medicine. 2007;135:1940-6037. 255. Risbud MV, Guttapalli A, Stokes DG, Hawkins D, Danielson KG, Schaer TP, et al. Nucleus pulposus cells express HIF-1α under normoxic culture conditions: A metabolic adaptation to the intervertebral disc microenvironment. Journal of cellular biochemistry. 2006;98(1):152-9. 256. Buckwalter JA, Rosenberg LC. Electron microscopic studies of cartilage proteoglycans. Electron microscopy reviews. 1988;1(1):87. 257. Thornton DJ, Nieduszynski IA, Oates K, Sheehan JK. Electron-microscopic and electrophoretic studies of bovine femoral-head cartilage proteoglycan fractions. Biochemical Journal. 1986;240(1):41. 258. Buckwalter J, Rosenberg L. Electron microscopic studies of cartilage proteoglycans. Direct evidence for the variable length of the chondroitin sulfate-rich region of proteoglycan subunit core protein. The Journal of biological chemistry. 1982;257(16):9830. 259. Katta J, Jin Z, Ingham E, Fisher J. Chondroitin sulphate: an effective joint lubricant? Osteoarthritis and Cartilage. 2009;17(8):1001-8. 260. Frank J. Single-particle imaging of macromolecules by cryo-electron microscopy. Annual review of biophysics and biomolecular structure. 2002;31(1):303-19. 261. Waigh T, Papagiannopoulos A. Biological and Biomimetic Comb Polyelectrolytes. Polymers. 2010;2(2):57-70. 262. Feuz L, Leermakers FAM, Textor M, Borisov O. Bending rigidity and induced persistence length of molecular bottle brushes: a self-consistent-field theory. Macromolecules. 2005;38(21):8891-901. 263. Papagiannopoulos A, Waigh T, Fernyhough C, Hardingham T, Heinrich M. The Structure and Dynamics of Flexible Polyelectrolyte Combs. Arxiv preprint cond-mat/0508655. 2005. 264. Papagiannopoulos A, Fernyhough C, Waigh T. The microrheology of polystyrene sulfonate combs in aqueous solution. The Journal of Chemical Physics. 2005;123:214904.

Page 369: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

343

265. Dobrynin AV, Colby RH, Rubinstein M. Scaling theory of polyelectrolyte solutions. Macromolecules. 1995;28(6):1859-71. 266. Rubinstein M, Colby RH, Dobrynin AV. Dynamics of semidilute polyelectrolyte solutions. Physical review letters. 1994;73(20):2776-9. 267. Konop AJ, Colby RH. Polyelectrolyte charge effects on solution viscosity of poly (acrylic acid). Macromolecules. 1999;32(8):2803-5. 268. Urban J, Maroudas A, Bayliss M, Dillon J. Swelling Pressures of Proteoglycans at the Concentrations Found in Cartilaginous Tissues. Biorheology. 1979;16:447-64. 269. Kovach IS. The importance of polysaccharide configurational entropy in determining the osmotic swelling pressure of concentrated proteoglycan solution and the bulk compressive modulus of articular cartilage. Biophysical Chemistry. 1995;53(3):181-7. 270. Massey CJ. Finite Element Analysis and Materials Characterization of Changes Due to Aging and Degeneration of the Human Intervertebral Disc. Philadelphia: Drexel University; 2009. 271. Massey CJ, Sarkar S., van Donkelaar C., Marcolongo, M., editor. Restoration of Proteoglycan to the Nucleus Pulposus of the Intervertebral Disc. 56th Annual Meeting of the Orthopaedic Research Society; 2010; New Orleans, Louisiana. 272. Mohammed S. Cost-effectiveness of epidural steroid injections to treat lumbosacral radiculopathy in chronic pain patients managed under workers' compensation [Master of Science in Public Health]: University of South Florida; 2008. 273. Singh K, Masuda K, Thonar EJMA, An HS, Cs-Szabo G. Age-related changes in the extracellular matrix of nucleus pulposus and anulus fibrosus of human intervertebral disc. Spine. 2009;34(1):10. 274. Horkay F, Basser PJ, Hecht AM, Geissler E. Insensitivity to Salt of Assembly of a Rigid Biopolymer Aggrecan. Physical review letters. 2008;101(6):68301. 275. Urban J, Maroudas A. The measurement of fixed charged density in the intervertebral disc. Biochimica et Biophysica Acta (BBA)-General Subjects. 1979;586(1):166-78. 276. Nerucci F, Fioravanti A, Cicero MR, Collodel G, Marcolongo R. Effects of chondroitin sulfate and interleukin-1[beta] on human chondrocyte cultures exposed to pressurization: a biochemical and morphological study. Osteoarthritis and Cartilage. [doi: 10.1053/joca.1999.0302]. 2000;8(4):279-87.

Page 370: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

344

277. Sharma G, Saxena R, 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(11):1387-94. 278. Fioravanti A, Collodel G. In Vitro Effects of Chondroitin Sulfate. In: Nicola V, editor. Advances in Pharmacology: Academic Press; 2006. p. 449-65. 279. Petersen H, Fechner PM, Fischer D, Kissel T. Synthesis, Characterization, and Biocompatibility of Polyethylenimine-graft-poly(ethylene glycol) Block Copolymers. Macromolecules. [doi: 10.1021/ma012060a]. 2002;35(18):6867-74. 280. Horkay F, Basser PJ, Hecht A-M, Geissler E. Gel-like behavior in aggrecan assemblies. The Journal of Chemical Physics. 2008;128(13):135103. 281. Heath R, Di Y, Clara S, Hudson A, Manock H. The optimization of epoxide-based tannage systems: an initial study. 2005. 282. Fournier D, Hoogenboom R, Schubert US. Clicking polymers: a straightforward approach to novel macromolecular architectures. Chem Soc Rev. 2007;36(8):1369-80. 283. Hoyle CE, Lowe AB, Bowman CN. Thiol-click chemistry: a multifaceted toolbox for small molecule and polymer synthesis. Chemical society reviews. 2010;39(4):1355-87. 284. Papagiannopoulos A, Fernyhough C, Waigh T, Radulescu A. Scattering Study of the Structure of Polystyrene Sulfonate Comb Polyelectrolytes in Solution. Macromolecular Chemistry and Physics. 2008;209(24):2475-86. 285. Risbud M, Albert T, Guttapalli A, Vresilovic E, Hillibrand A, Vaccaro A, et al. Differentiation of mesenchymal stem cells towards a nucleus pulposus-like phenotype in vitro: implications for cell-based transplantation therapy. Spine. 2004;29(23):2627. 286. Cannella M AO, E Vresilovic, M Marcolongo. Nucleus Pulposus Augmentation Creates a Linear Relationship among Implant Volume, Disc Height and Intradiscal Pressure. Biomechanics. In Preparation. 287. Masuda K, Imai Y, Okuma M, Muehleman C, Nakagawa K, Akeda K, et al. Osteogenic protein-1 injection into a degenerated disc induces the restoration of disc height and structural changes in the rabbit anular puncture model. Spine. 2006;31(7):742. 288. Miyamoto K, Masuda K, Kim JG, Inoue N, Akeda K, Andersson GBJ, et al. Intradiscal injections of osteogenic protein-1 restore the viscoelastic properties of degenerated intervertebral discs. The Spine Journal. 2006;6(6):692-703. 289. Masuda K, Lotz JC. New challenges for intervertebral disc treatment using regenerative medicine. Tissue Eng Part B Rev. Feb;16(1):147-58.

Page 371: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

345

290. Kwon HJ, Gong JP. Negatively charged polyelectrolyte gels as bio-tissue model system and for biomedical application. Current Opinion in Colloid & Interface Science. [doi: 10.1016/j.cocis.2006.09.006]. 2006;11(6):345-50. 291. Chang KY, Cheng LW, Ho GH, Huang YP, Lee YD. Fabrication and characterization of poly ([gamma]-glutamic acid)-graft-chondroitin sulfate/polycaprolactone porous scaffolds for cartilage tissue engineering. Acta Biomaterialia. 2009;5(6):1937-47. 292. Huang B, Li CQ, Zhou Y, Luo G, Zhang CZ. Collagen II/hyaluronan/chondroitin 6 sulfate tri copolymer scaffold for nucleus pulposus tissue engineering. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2010;92(2):322-31. 293. Jo S, Kim D, Woo J, Yoon G, Park YD, Tae G, et al. Development and physicochemical evaluation of chondroitin sulfate-poly (ethylene oxide) hydrogel. Macromolecular Research. 2011;19(2):147-55. 294. Nakashima S, Matsuyama Y, Takahashi K, Satoh T, Koie H, Kanayama K, et al. Regeneration of intervertebral disc by the intradiscal application of cross-linked hyaluronate hydrogel and cross-linked chondroitin sulfate hydrogel in a rabbit model of intervertebral disc injury. Bio-medical materials and engineering. 2009;19(6):421. 295. Han EH, Wilensky LM, Schumacher BL, Chen AC, Masuda K, Sah RL. Tissue Engineering by Molecular Disassembly and Reassembly: Biomimetic Retention of Mechanically Functional Aggrecan in Hydrogel. Tissue Engineering Part C: Methods. 2010;16(6):1471-9. 296. Fleet R, van den Dungen ETA, Klumperman B. Synthesis of novel glycopolymer brushes via a combination of RAFT-mediated polymerisation and ATRP. South African Journal of Science. 2011;107(3/4). 297. Lienkamp K, Ruthard C, Lieser G, Berger R, Groehn F, Wegner G. Polymerization of styrene sulfonate ethyl ester and styrene sulfonate dodecyl ester by ATRP: synthesis and characterization of polymer brushes. Macromolecular Chemistry and Physics. 2006;207(22):2050-65. 298. Akeroyd N, Klumperman B. The combination of living radical polymerization and click chemistry for the synthesis of advanced macromolecular architectures. European Polymer Journal. [doi: 10.1016/j.eurpolymj.2011.02.003]. 2011;47(6):1207-31. 299. Ladmiral V, Mantovani G, Clarkson GJ, Cauet S, Irwin JL, Haddleton DM. Synthesis of neoglycopolymers by a combination of “click chemistry” and living radical polymerization. Journal of the American Chemical Society. 2006;128(14):4823-30.

Page 372: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

346

300. Slavin S, Burns J, Haddleton DM, Becer CR. Synthesis of glycopolymers via click reactions. European Polymer Journal. 2010. 301. Grande D, Guerrero R, Gnanou Y. Cyanoxyl-mediated free-radical polymerization of acrylic acid: Its scope and limitations. Journal of Polymer Science Part A: Polymer Chemistry. 2004;43(3):519-33. 302. Tsuda T, Nishimura SI. Synthesis of an antifreeze glycoprotein analogue: efficient preparation of sequential glycopeptide polymers. Chem Commun. 1996(24):2779-80. 303. Tachibana Y, Monde K, Nishimura SI. Sequential glycoproteins: Practical method for the synthesis of antifreeze glycoprotein models containing base labile groups. Macromolecules. 2004;37(18):6771-9. 304. Tachibana Y, Fletcher GL, Fujitani N, Tsuda S, Monde K, Nishimura SI. Antifreeze glycoproteins: elucidation of the structural motifs that are essential for antifreeze activity. Angewandte Chemie. 2004;116(7):874-80. 305. Voit B, Appelhans D. Glycopolymers of Various Architectures—More than Mimicking Nature. Macromolecular Chemistry and Physics. 2010;211(7):727-35. 306. Lee H, Pietrasik J, Sheiko SS, Matyjaszewski K. Stimuli-responsive molecular brushes. Progress in Polymer Science. 2010;35(1-2):24-44. 307. Spain SG, Cameron NR. A spoonful of sugar: the application of glycopolymers in therapeutics. Polym Chem. 2010.

Page 373: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule

347

Vita

Sumona Sarkar was born on January 30th, 1981 in Poughkeepsie, NY. She

graduated from Springbrook High School in Silver Spring, MD in 1999. She received

her B.S. in Biomedical Engineering from Boston University in 2003 and her M.S. in

Biomedical Engineering in 2006 also from Boston University. During her tenure at

Drexel University, Sumona received the International Travel Award from Drexel (March

2011, Strasbourg, France) as well as the Best Student Paper, Silver Award at the 2nd

International Symposium on Surface and Interface of Biomaterials (January 2010, Hong

Kong). She also received the Graduate Student Award for Inspired Teaching in 2008

2008 for her contribution as an Engineering Design Education Fellow (2007 – 2008).

Also during her time at Drexel, Sumona co-authored, Marcolongo M, Sarkar S,

Ganesh N. Comprehensive Biomaterials: Trends in Materials for Spine Surgery.

Ducheyne P, editor (2011). She was also a collaborating author on the U.S. Patent, Appl.

No. PCT/US10/56064, "Compositions and Methods for Treating a Disorder or Defect in

Soft Tissue" Marcolongo, M., Versilovic, E., Jackson, B., Sarkar, S., Schauer, C., Filed

Nov. 9, 2010. In preparation are the following publications:

Sarkar, S., Schauer, C.L., Vresilovic, E., Marcolongo, M. (2011), “Immobilization

of Chondroitin Sulfate for Biomimetic Polymer Brush Synthesis Strategies.” In

Preparation.

Massey, C.J., Sarkar, S., Marcolongo, M. (2011), “Fourier Transform Infrared

Spectroscopy Characterization of the Human Intervertebral Disc.” In Preparation.

Page 374: Synthesis and Characterization of a Chondroitin Sulfate Based Hybrid BioSynthetic Biomimetic Aggrecan Macromolecule