280
Novel Polysaccharide Based Polymers and Nanoparticles for Controlled Drug Delivery and Biomedical Imaging by Alireza Shalviri A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Pharmaceutical Sciences University of Toronto © Copyright by Alireza Shalviri (2012)

Novel Polysaccharide Based Polymers and Nanoparticles for … · 2013. 1. 7. · polysaccharides such as biocompatibility, biodegradability, upgradability, multiple reacting groups

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

  • Novel Polysaccharide Based Polymers and Nanoparticles for Controlled Drug Delivery and

    Biomedical Imaging

    by

    Alireza Shalviri

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

    Graduate Department of Pharmaceutical Sciences University of Toronto

    © Copyright by Alireza Shalviri (2012)

  • ii

    Novel Polysaccharide Based Polymers and Nanoparticles for Controlled Delivery of Drugs and

    Imaging Agents

    Alireza Shalviri

    Doctor of Philosophy

    Graduate Department of Pharmaceutical Sciences

    University of Toronto

    2012

    Abstract

    The use of polysaccharides as building blocks in the development of drugs and contrast agents

    delivery systems is rapidly growing. This can be attributed to the outstanding virtues of

    polysaccharides such as biocompatibility, biodegradability, upgradability, multiple reacting

    groups and low cost. The focus of this thesis was to develop and characterize novel starch based

    hydrogels and nanoparticles for delivery of drugs and imaging agents. To this end, two different

    systems were developed. The first system includes polymer and nanoparticles prepared by graft

    polymerization of polymethacrylic acid and polysorbate 80 onto starch. This starch based

    platform nanotechnology was developed using the design principles based on the

    pathophysiology of breast cancer, with applications in both medical imaging and breast cancer

    chemotherapy. The nanoparticles exhibited a high degree of doxorubicin loading as well as

    sustained pH dependent release of the drug. The drug loaded nanoparticles were significantly

    more effective against multidrug resistant human breast cancer cells compared to free

    doxorubicin. Systemic administration of the starch based nanoparticles co-loaded with

    doxorubicin and a near infrared fluorescent probe allowed for non-invasive real time monitoring

    of the nanoparticles biodistribution, tumor accumulation, and clearance. Systemic administration

  • iii

    of the clinically relevant doses of the drug loaded particles to a mouse model of breast cancer

    significantly enhanced therapeutic efficacy while minimizing side effects compared to free

    doxorubicin. A novel, starch based magnetic resonance imaging (MRI) contrast agent with good

    in vitro and in vivo tolerability was formulated which exhibited superior signal enhancement in

    tumor and vasculature. The second system is a co-polymeric hydrogel of starch and xanthan gum

    with adjustable swelling and permeation properties. The hydrogels exhibited excellent film

    forming capability, and appeared to be particularly useful in controlled delivery applications of

    larger molecular size compounds. The starch based hydrogels, polymers and nanoparticles

    developed in this work have shown great potentials for controlled drug delivery and biomedical

    imaging applications.

  • iv

    Acknowledgments

    I would like to thank my supervisor Dr. Shirley Wu for believing in me as well as for her

    kindness, leadership, patience, and continuous support. She always believed in my abilities and

    pushed me to be the best I could be; for this I will always remain thankful to her. Without her

    mentorship this thesis would have not been possible.

    I am greatly appreciative to Drs. Ping Lee, Tigran Chalikian and Edgar Acosta for taking the

    time to attend my committee meetings and giving me insightful guidance and recommendations.

    I also thank Dr. Warren Foltz, Dr. Andrew Rauth, and Dr. Heiko Heerklotz for their continuous

    support and guidance throughout my thesis.

    My sincere thanks to all my colleagues who helped me, especially Ping Cai. I also extend my

    gratitude to those not mentioned here who have taken part, small or large, in making this work

    possible.

    I am grateful to the Ontario Graduate Scholarship Program, Natural Sciences and Engineering

    Research Council of Canada, BioPotato network (Co-leaders: Drs. Helen Tai and Yvan

    Pelletier), Agricultural Bioproducts Innovation Program (ABIP) of Agriculture & Agri-Food

    Canada, CIHR/CBCRA, University of Toronto and Leslie Dan Faculty of Pharmacy for

    scholarships and research funding.

    I thank my wife, Mana, for always being there for me. Without her love, care and undivided

    attention I would not have achieved any of this.

    I thank my parents and my sister for their belief in my abilities and constant moral and financial

    support. I owe all my success to them. Finally, I thank the mice. Without them meaningful

    innovation would not be possible.

  • v

    Table of Contents

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

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

    List of Tables ................................................................................................................................ xii

    List of Figures .............................................................................................................................. xiii

    List of Abbreviations ................................................................................................................... xxi

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

    1.1 Breast Cancer .......................................................................................................................2

    1.1.1 Epidemiology ...........................................................................................................2

    1.1.2 Breast Cancer Cells and Tumor Microenvironment ................................................2

    1.1.3 Doxorubicin: a Potent Drug for Breast Cancer Chemotherapy ...............................5

    1.1.4 Barriers to Cancer Chemotherapy ............................................................................7

    1.1.5 Nanoparticulate Systems in Cancer Therapy .........................................................13

    1.1.6 Nanoparticles as Theranostics in Cancer ...............................................................19

    1.2 Polysaccharides in Drug Delivery .....................................................................................22

    1.2.1 Starch .....................................................................................................................23

    1.2.2 Xanthan gum ..........................................................................................................26

    1.3 Biomedical Imaging ...........................................................................................................27

    1.3.1 In vivo Fluorescence Imaging ................................................................................27

    1.3.2 Magnetic Resonance Imaging ................................................................................30

    1.4 Drug Delivery to the Brain ................................................................................................40

    1.4.1 Brain Anatomy .......................................................................................................41

    1.4.2 Strategies to Enhance Drug Delivery to the Brain .................................................43

    1.5 Goal for this work ..............................................................................................................47

    1.6 Synopsis .............................................................................................................................48

  • vi

    Chapter 2. Design of pH-responsive Nanoparticles of Terpolymer of Poly(methacrylic acid),

    Polysorbate 80 and Starch for Delivery of Doxorubicin ...........................................................51

    2.1 Abstract ..............................................................................................................................52

    2.2 Introduction ........................................................................................................................53

    2.3 Materials and Methods .......................................................................................................55

    2.3.1 Materials ................................................................................................................55

    2.3.2 Synthesis of PMAA-PS 80-g-St Nanoparticles .....................................................55

    2.3.3 FTIR and 1H NMR Spectroscopy ..........................................................................56

    2.3.4 Examination of the Nanoparticles with TEM ........................................................57

    2.3.5 Determination of Particle Size and Surface Charge ..............................................57

    2.3.6 Titration Studies .....................................................................................................57

    2.4 Results and Discussion ......................................................................................................58

    2.4.1 PMAA-PS 80-g-St Nanoparticles were Synthesized Using a Simple One-pot Method ...................................................................................................................58

    2.4.2 Polymer Composition of the Nanoparticles ...........................................................60

    2.4.3 Nanoparticles Size and Morphology ......................................................................64

    2.4.4 PMAA-PS 80-g-St Nanoparticles Show pH-responsive Swelling in Physiological pH Range .........................................................................................67

    2.4.5 Properties of Carboxylic Acid Groups in the Nanoparticles .................................68

    2.4.6 Effect of Processing Parameters on Particle size and pH Sensitivity ....................71

    2.5 Conclusions ........................................................................................................................74

    2.6 Acknowledgements ............................................................................................................74

    Chapter 3. pH Dependent Doxorubicin Release by Nanoparticles Based on Terpolymer of

    Poly(Methacrylic Acid), Polysorbate 80, and Starch for Overcoming Multi-drug

    Resistance in Breast Cancer Cells .............................................................................................75

    3.1 Abstract ..............................................................................................................................76

    3.2 Introduction ........................................................................................................................77

    3.3 Materials and Methods .......................................................................................................80

  • vii

    3.3.1 Materials ................................................................................................................80

    3.3.2 Cell Maintenance ...................................................................................................80

    3.3.3 Synthesis of PMAA-PS 80-g-St Nanoparticles .....................................................81

    3.3.4 Fourier Transform Infrared Spectroscopy .............................................................81

    3.3.5 Isothermal Titration Calorimetry (ITC) .................................................................81

    3.3.6 Dynamic Light Scattering ......................................................................................82

    3.3.7 Transmission Electron Microscopy .......................................................................82

    3.3.8 Drug Loading Studies ............................................................................................83

    3.3.9 X-ray Powder Diffraction (XRPD) ........................................................................84

    3.3.10 In vitro Drug Release .............................................................................................84

    3.3.11 Cell Uptake Studies Using Fluorescence Microscopy ...........................................84

    3.3.12 Cellular Uptake of Nanoparticles by Flow Cytometry ..........................................85

    3.3.13 In vitro Assessment of Anticancer Efficacy of Dox-loaded Nanoparticles ...........86

    3.4 Results ................................................................................................................................86

    3.4.1 Properties of PMAA-PS 80-g-St Nanoparticles and Their High Capability of Efficiently Loading Dox without Loss of Colloidal Stability ................................86

    3.4.2 FTIR, XRD and ITC Experiments Revealed Strong Ionic Interaction between the Nanoparticles and Dox .....................................................................................88

    3.4.3 The Nanoparticles Exhibited Sustained and pH Dependent Release of Dox in vitro ........................................................................................................................95

    3.4.4 Substantial Cellular Uptake of the Nanoparticles Evidenced by Fluorescence Microscopy, TEM and Flow Cytometry ................................................................96

    3.4.5 The Dox Loaded Nanoparticles Were Significantly More Effective Against

    MDR1 Cells than Free Dox .................................................................................101

    3.5 Discussion ........................................................................................................................103

    3.6 Conclusions ......................................................................................................................105

    3.7 Acknowledgements ..........................................................................................................105

  • viii

    Chapter 4. Evaluation of New Multifunctional Nanoparticles based on Terpolymer of

    Polymethacrylic acid, Polysorbate 80, and Starch as a Theranostic Nanoplatform for

    Simultaneous in vivo Imaging and Treatment of Breast Cancer .............................................106

    4.1 Abstract ............................................................................................................................107

    4.2 Introduction ......................................................................................................................108

    4.3 Materials and Methods .....................................................................................................110

    4.3.1 Materials ..............................................................................................................110

    4.3.2 Preparation of Dual Mode Nanoparticles ............................................................111

    4.3.3 Dynamic Light Scattering, Electrophoretic Mobility Measurements, and Transmission Electron Microscopy .....................................................................113

    4.3.4 Drug Release Studies ...........................................................................................113

    4.3.5 Cell Lines .............................................................................................................114

    4.3.6 Animal Model and In vivo Treatment Protocol in Tumor Bearing Mice ............114

    4.3.7 Real Time In vivo and Ex-vivo Near-infrared Fluorescent Imaging ....................115

    4.3.8 Ex-vivo Tumor Fluorescence Microscopy ...........................................................117

    4.4 Results ..............................................................................................................................118

    4.4.1 Properties of the Nanoparticles ............................................................................118

    4.4.2 Distribution and Tumor Accumulation of the PF-NPs and SA-NPs in Whole Animals In vivo ....................................................................................................122

    4.4.3 Real Time Pharmocokinetics of Nanoparticles in Tumor Tissue ........................125

    4.4.4 Organ Distribution of the Nanoparticles Determined by Ex Vivo Imaging ........126

    4.4.5 Microscopic Imaging of Tumor tissue Demonstrated Extravasation of the Nanoparticles in the Tumor .................................................................................128

    4.4.6 Anti-tumor Efficacy of the Nanoparticles in a Murine Breast Cancer Tumor Model ...................................................................................................................130

    4.4.7 Preliminarily Assessment of Toxicity of the Nanoparticles ................................132

    4.5 Discussion ........................................................................................................................134

    4.6 Conclusions ......................................................................................................................136

    4.7 Acknowledgments............................................................................................................137

  • ix

    Chapter 5. A New Starch-based Polymeric MRI Contrast Agent with Superior Signal

    Enhancement in Blood and Tumor .........................................................................................138

    5.1 Abstract ............................................................................................................................139

    5.2 Introduction ......................................................................................................................140

    5.3 Materials and Methods .....................................................................................................142

    5.3.2 Cell line and Maintenance ...................................................................................142

    5.3.3 Preparation of the Gd3+

    Loaded PMAA-PS 80-g-St-DTPA Polymer (PolyGd) ..143

    5.3.4 Confirmation of DTPA Conjugation to Starch ....................................................144

    5.3.5 Determination of DTPA Content and Binding Affinity of Gd3+

    to St-DTPA .....145

    5.3.6 Determination of Cytotoxicity of PolyGd............................................................145

    5.3.7 Comparison of the Relaxivity Properties of PolyGd and Omniscan® in vitro ....146

    5.3.8 Experimental Animals and Induction of Subcutaneous Breast Tumors ..............146

    5.3.9 Determination of In vivo MRI Contrast Enhancement of PolyGd in Mice .........147

    5.3.10 Validation of Whole-body Gd3+ Distribution by Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES) ........................................................148

    5.4 Results and Discussion ....................................................................................................148

    5.4.1 Confirmation of DTPA Conjugation to Starch ....................................................148

    5.4.2 Determination of DTPA Content and Binding Affinity of Gd3+ to St-DTPA .....154

    5.4.3 PolyGd Exhibited Lower Cytotoxicity than Free Gd3+ ........................................155

    5.4.4 PolyGd showed much higher relaxivity than Omniscan®...................................156

    5.4.5 In vivo MRI Contrast Enhancement of PolyGd in Mice ......................................159

    5.4.6 Biodistribution and Clearance of the PolyGd from the Body ..............................167

    5.5 Conclusions ......................................................................................................................168

    5.6 Acknowledgments............................................................................................................169

    Chapter 6. Evaluating the Capability of Novel Starch Based Nanoparticles for Controlled

    Delivery of Drugs and Imaging Agents to the Brain ..............................................................170

    6.1 Abstract ............................................................................................................................171

  • x

    6.2 Introduction ......................................................................................................................172

    6.3 Materials and Methods .....................................................................................................174

    6.3.1 Materials ..............................................................................................................174

    6.3.2 Synthesis and Preparation of the PMAA-Ps 80-g-St Polymer and

    Nanoparticles .......................................................................................................175

    6.3.3 Physicochemical Characterization of the PMAA-PS 80-g-St Polymer and Nanoparticles .......................................................................................................175

    6.3.4 Time-of-Flight-Secondary Ion Mass Spectrometry .............................................176

    6.3.5 Animal Studies .....................................................................................................176

    6.3.6 In vivo Magnetic Resonance Imaging (MRI).......................................................176

    6.3.7 Ex-vivo Fluorescence Imaging of the Brain .........................................................177

    6.3.8 Fluorescence Microscopy ....................................................................................178

    6.4 Results ..............................................................................................................................179

    6.4.1 Properties of PMAA-PS 80-g-St Polymer and Nanoparticles .............................179

    6.4.2 Brain Accumulation of the PMAA-PS 80-g-St Nanoparticles at Macroscopic

    and Microscopic Levels .......................................................................................184

    6.5 Conclusions ......................................................................................................................189

    Chapter 7. Novel Modified Starch-xanthan Gum Hydrogels for Controlled Drug Delivery:

    Synthesis and Characterization ...............................................................................................190

    7.1 Abstract ............................................................................................................................191

    7.2 Introduction ......................................................................................................................192

    7.3 Materials and Methods .....................................................................................................193

    7.3.1 Chemicals .............................................................................................................193

    7.3.2 Synthesis of Cross-linked Fully Gelatinized Starch and Xanthan Gum Hydrogels .............................................................................................................193

    7.3.3 Examination of Film Morphology .......................................................................194

    7.3.4 Confirmation of Cross-linking by Fourier Transformed Infrared Spectroscopy .194

    7.3.5 Solid State 31P NMR Spectroscopy......................................................................194

  • xi

    7.3.6 Study of the Swelling Behavior of Cross-linked Starch-xanthan Gum Polymer .195

    7.3.7 Determination of Gel Mesh Size..........................................................................195

    7.3.8 Permeability Studies ............................................................................................197

    7.3.9 Statistical Analysis ...............................................................................................198

    7.4 Results and Discussion ....................................................................................................198

    7.4.1 Morphology..........................................................................................................198

    7.4.2 Cross-linking of Starch-Xanthan Gum with STMP was Confirmed ...................199

    7.4.3 Swelling Kinetics and Equilibrium Swelling Ratio .............................................205

    7.4.4 Effect of STMP and Xanthan gum Concentration on the Film Swelling ............206

    7.4.5 Effect of Medium pH and Buffer salts on the Film Swelling ..............................209

    7.4.6 Mesh Size of the Modified Starch-Xanthan Gum Gels .......................................210

    7.4.7 Permeability Studies ............................................................................................212

    7.5 Conclusion .......................................................................................................................215

    7.6 Acknowledgements ..........................................................................................................216

    Chapter 8. Conclusions and Future Perspectives .........................................................................217

    8.1 Overall Conclusions and Original Contributions of This Thesis .....................................217

    8.2 Limitation of the Work and Future Directions ................................................................222

    8.2.1 Biodegradation, Body Clearance, and Biocompatibility of the PMAA-PS 80-

    g-St Nanoparticles ................................................................................................222

    8.2.2 Utility of the PMAA-PS 80-g-St Polymer and Nanoparticles as a Dual Model Imaging Probe ......................................................................................................223

    8.2.3 Active Targeting ..................................................................................................224

    8.2.4 Delivery of Dual Agents by the Nanoparticles ....................................................225

    8.2.5 In vivo assessment of PMAA-PS 80-g-St for drug delivery to the CNS .............226

    8.2.6 Optimal Polymerization Method for Making More Uniform Polymers .................227

    8.2.6 In Vivo Assessment of Efficacy in Multidrug Resistant Tumor Model ..............227

    Bibliography ................................................................................................................................228

  • xii

    List of Tables

    Table 1.1 Examples of drug-nanoparticles system for delivery to the brain ................................ 44

    Table 2.1 Nanoparticle preparation recipes and the polymer composition.. ................................ 60

    Table 2.2 Intensity-weighted hydrodynamic diameter of nanoparticles with different feed molar

    ratio of MAA/St in 0.15 M PBS of various pH.. .......................................................... 65

    Table 3.1 Characterization of the drug-loaded nanoparticles. The effect of drug loading on the

    particles size and surface charge is investigated.. ......................................................... 87

    Table 4.1 Summary of physicochemical properties of SA-NPs and PF-NPs ............................. 118

    Table 4.2 Tumor associated pharmacokinetic data derived from the tumor average fluorescence

    intensity versus time curve for SA-NPs and PF-NPs.. ............................................... 126

    Table 5.1 Gd3+

    content, molecular weight, and r1 for Omniscan® and PolyGd. The r1 were

    measured in saline at 3 and 7 T.. ................................................................................. 158

    Table 7.1 Equilibrium swelling ratio of films with 10% xanthan gum and various cross-linker

    STMP levels.. .............................................................................................................. 206

    Table 7.2 Equilibrium volume swelling and gel mesh size of starch-xanthan gum gels containing

    10% xanthan gum and varying concentrations of cross-linker (STMP).. ................... 212

    Table 7.3 Equilibrium swelling ratio of films in 0.15M phosphate buffer of pH=7.4 and

    permeability of vitamin B12 across modified starch-xanthan gum films of various

    compositions. .............................................................................................................. 213

    Table 7.4 Permeability of macromolecules and drugs of various molecular weights and charges

    across the starch-xanthan gum gel film containing 10% XG and 5% STMP in 0.15 M

    phosphate buffer of pH=7.4.. ...................................................................................... 214

    Table 7.5 Permeability of two weakly acidic drugs across starch-xanthan gum films containing

    10%XG and 5% STMP in pH 2 and 7.4 buffer solutions with ionic strength of 0.15M.

    .................................................................................................................................... 215

  • xiii

    List of Figures

    Figure 1.1 Chemical structure of Doxorubicin ............................................................................... 6

    Figure 1.2 Factors leading to MDR in cancer. One of the leading causes of treatment failure in

    cancer is the onset of resistant disease. MDR can be divided into two broad categories:

    Cellular resistance and non-cellular resistance. ............................................................ 11

    Figure 1.3 Passive targeting of the tumor by EPR. The EPR effect is the balance of enhanced

    tumor permeability with poor tumor interstitial fluid drainage, resulting in the selective

    uptake and retention of nanoparticles in the tumor tissue. ........................................... 14

    Figure 1.4 Drug loaded nanoparticles can overcome MDR cancer cells. Endocytosis of the drug

    loaded nanoparticles in membrane bound vesicles protects the drug from the action of

    the membrane efflux pumps. The nanoparticles release the drug deep inside the cell

    and the drug can gain access to its cellular target site (e.g. DNA) ............................... 14

    Figure 1.5 Chemical structure of starch ........................................................................................ 24

    Figure 1.6 Xanthan gum chemical structure ................................................................................. 27

    Figure 1.7 Schematic comparison of fluorophores in the visible spectrum versus the near infrared

    for deep in vivo fluorescent imaging. Fluorophores in the visible regions are limited by

    poor penetration of the excitation photon or the poor penetration of the emission

    photons preventing the detection of fluorescent signal in vivo. Near infrared photons

    can provide deep tissue penetration. ............................................................................. 29

    Figure 2.1 Schematic reaction of starch grafting and terpolymer formation. FTIR spectra of

    Starch, PMAA-PS 80, and PMAA-PS 80-g-St. ............................................................................ 61

    Figure 2.2 1H NMR spectra of A) PS 80, B) Starch, C) PMAA-PS 80, D) PMAA-PS 80-g-St-2 in

    0.05M NaOD ................................................................................................................ 63

    Figure 2.3 A) Intensity-weighted hydrodynamic diameter of the PMAA-PS 80-g-St-2

    nanoparticles in 0.15 M pH 7.4 PBS. The particles showed a narrow size distribution.

    B) TEM images of PMAA-PS 80-g-St-2 in 0.15 M PBS of pH=7.4.. ......................... 66

  • xiv

    Figure 2.4 A) Relative diameter vs. pH for the nanoparticles with different feed molar ratio of

    MAA/St in 0.15 M PBS. DpH/D4 represents particles diameter at different pH relative

    to pH 4. B) Effect of pH on surface charge for particles of various MAA/St molar

    ratio.. ............................................................................................................................. 68

    Figure 2.5 A) Potentiometric titration curves. Empty triangles represent the uncorrected

    potentiometric titation curve for PMAA-PS 80-g-St-2 latex dispersion. Filled circles

    represent the titration curve after correction. Empty circles show the blank titration

    curve. The arrow represents the equivalence point. The equivalence points are used to

    calculate the MAA contents in various nanoparticle batches. B) Variation in the

    apparent dissociation constant (pKa) as a function of the degree of ionization (α) for

    nanoparticles of different starch and MAA contents. ................................................... 70

    Figure 2.6 Effect of A) SDS, B) PS 80, C) total monomer concentration, D) cross-linker molar

    ratio on particle size and pH sensitivity.. ...................................................................... 73

    Figure 3.1 A) Number-weighted Gaussian distribution of PMAA-PS 80-g-St nanoparticles

    loaded with doxorubicin (LC=33%) in 0.15 M phosphate buffer at pH 7.4, B)

    Transmission electron micrograph of doxorubicin loaded nanoparticles (LC=33%)... 88

    Figure 3.2 FTIR spectra of A) PMAA-PS 80-g-St nanoparticles, B) Dox, C) Dox loaded PMAA-

    PS 80-g-St nanoparticles.. ............................................................................................. 89

    Figure 3.3 XRD spectrum of A) Doxorubicin in native form, B) PMAA-PS 80-g-St

    nanoparticles, C) Doxorubicin loaded nanoparticles (LC=50%), D) doxorubicin loaded

    nanoparticles (LC=50%) after 6 months storage at room temperature.. ....................... 91

    Figure 3.4 A) The blank differential enthalpy curves of titrating 8.5 mM doxorubicin into buffers

    of various pH. B) Differential enthalpy curves of titrating 8.5 mM doxorubicin into

    0.1mg/ml PMAA-PS 80-g-St nanoparticles in buffers at different pH. The ionic

    strength was kept constant at 0.15M by addition of NaCl. C) The blank differential

    enthalpy curves of titrating 8.5 mM doxorubicin into DDIW with different NaCl

    contents, D) Differential enthalpy curves of titrating 8.5 mM doxorubicin into

    0.1mg/ml PMAA-PS 80-g-St nanoparticles in DDIW with different NaCl contents. .. 94

  • xv

    Figure 3.5 Effect of pH on kinetics of doxorubicin release from the naoparticles with drug

    loading content of 33% at 37 ºC. The release of free doxorubicin from the dialysis bag

    was used as control. For each buffer system the ionic strength was kept constant at

    0.15 M by adding NaCl. ................................................................................................ 95

    Figure 3.6 A) Fluorescence microscopy image of MDA-MB435/LCC6 cells with and without

    (control) 4 hr incubation with fluorescent nanoparticles at final concentration of 0.25

    mg/ml. Nuclei were stained with Hoechst 33342 and visualized with DAPI filter, cell

    membranes were stained with Vybrant™DiI and visualized with Cy3 filter, and NPs

    were labelled fluoresceinamine isomer I and visualized with FITC filter. Optical slices

    were taken every 2 µm from the uppermost and lowermost regions of the cell, allowing

    for selection of an image at approximately the midpoint of the nucleus. B) TEM

    micrographs of MDA-MB435/LCC6 cells treated with 0.25 mg/ml PMAA-PS 80-g-St

    NPs for 4 hrs. The nanoparticles were loaded with gadolinium (metal) and appear as

    electron dense deposits. ................................................................................................ 98

    Figure 3.7 Flow cytometry histograms for MDA-MB435/LCC6 cells showing the effect of

    incubation time and temperature on particle uptake. The cells were incubated with

    fluorescent labelled naoparticles at the final nanoparticle concentration of 0.25 mg/ml.

    A) MDA-MB435/WT (1) untreated cells at 37 ºC, (2) 1 hr incubation at 37 ºC, (3) 4

    hrs incubation at 37 ºC, (4) 24 hrs incubation at 37 ºC. B) MDA-MB435/WT (1)

    untreated cells, (2) 1 hr incubation at 4ºC, (3) 1hr incubation at 37 ºC. C) MDA-MB

    435/MDR1 (1) untreated cells, (2) 1 hr incubation at 37 ºC, (3) 4 hrs incubation at 37

    ºC, (4) 24 hrs incubation at 37 ºC. D) MDA-MB435/MDR1 (1) untreated cells, (2) 1 hr

    incubation at 4ºC, (3) 1 hr incubation at 37 ºC.. ......................................................... 100

    Figure 3.8 Determination of the response of MDA-MB435/LCC6 cell types to free doxorubicin

    and doxorubicin loaded nanoparticles by MTT assay. (A-B) Cell viability of MDA-

    MB435/LCC6/WT (n=3) cells after exposure to increasing concentrations of blank

    nanoparticles (blank NPs), free doxorubicin and doxorubicin loaded nanoparticles

    (Dox-NPs) for 24 hrs (A) and 48 hours (B). (C-D) Cell viability of

    MDA435/LCC6/MDR1 (n=3) cells after exposure to increasing concentrations of

    blank NPs, free doxorubicin and Dox-NPs for 24 hrs (C) and 48 hrs (D). Cells with no

  • xvi

    treatment and incubated with blank nanoparticles were used as control for free drug

    and drug loaded nanoparticle respectively. Cell viability is expressed as the percent of

    control for each treatment group. ................................................................................ 102

    Figure 4.1 A) Schematic diagram of PF-NPs and SA-NPs and the reaction scheme for

    conjugation of the NIR dye and loading of Dox. B) Chemical structure of the PMAA-

    PS 80-g-St polymer. .................................................................................................... 119

    Figure 4.2 Size distribution and shapes of A) SA-NPs and B) PF-NPs, as determined by dynamic

    light scattering (DLS) and transmission electron microscopy (TEM), respectively. . 120

    Figure 4.3 Drug release kinetics from the SA-NPs (LC=21.1%) and PF-NPs (LC=49.7%). Drug

    release was measured in 0.15 M Tris/NaCl buffer, pH=7.4, at 37 ºC. The release of free

    doxorubicin from the dialysis bag was used as control. ............................................. 121

    Figure 4.4 A) Whole animal real time biodistribution and tumor targeting of SA-NPs and PF-

    NPs in mice bearing an orthotopic breast tumor model. Nanoparticle-associated

    fluorescence was determined prior to intravenous injection (baseline), and then at

    various hours following nanoparticles injection up to 14 days. B) Time-dependent

    excretion profiles of SA-NPs and PF-NPs from the whole body (left) and tumor (right).

    The fluorescence intensity for the region of interest was recorded as average radiant

    efficiency.. .................................................................................................................. 124

    Figure 4.5 Quantitative results of tissue distribution and tumor accumulation for SA-NPs and PF-

    NPs. Ratio of the relative fluorescence intensity in major organs, tumor, and blood as a

    function of time after intravenous injection of nanoparticles, compared to normal

    major organs and tumors not injected with NIR dye conjugated-nanoparticles.. ....... 127

    Figure 4.6 Microscopic distribution of the nanoparticles within the tumor. Fluorescent signal

    observed in tumor tissue treated with A) vehicle only; B) FITC-labeled SA-NPs; C)

    FITC-labeled PF-NPs. Tumors arising from orthotopically implanted EMT6/WT cells

    were allowed to grow for 8 days prior to injection of nanoparticles. Four hours

    following nanoparticle introduction, animals were sacrificed and the distribution of

    particles assessed within both core and peripheral regions. ....................................... 129

  • xvii

    Figure 4.7 Anti-tumour activity of starch based nanoparticles in EMT6/WT tumor bearing mice.

    Tumor cells were implanted orthotopicly on day zero. Mice were treated with 5%

    dextrose (n=2×4), free Dox (n=8), PF-NPs (n=2×4), and SA-NPs (n=2×3) at a dose of

    2×10 mg/kg equivalent to Dox on day 8 and 15. A) Tumor volume up to day 62. Each

    curve represents one animal. B) Kaplan Meier survival curves for 5% dextrose, free

    Dox, PF-NPs, and SA-NPs. The trend in survival curves were significantly different

    (p=0.0033, Mantel Cox).............................................................................................. 131

    Figure 4.8 Time profiles of body weight of tumor-bearing mice treated with 5% dextrose

    (n=2×4), free Dox (n=2×4) , PF-NPs (n=2×4), and SA-NPs (n=2×3) at a dose of 2×10

    mg/kg equivalent to Dox. Balb/c mice were inoculated with EMT6/WT tumor in the

    mammary fat pad and received treatment on day 8 and 15 post inoculation. Each curve

    represents one animal. ................................................................................................. 132

    Figure 4.9 Histological sections of lung, kidney, liver, and heart of tumor-bearing mice treated

    with 5% dextrose, free Dox, PF-NPs, and SA-NPs at a dose of 2×10 mg/kg equivalent

    to Dox. The samples were not collected at the same time after treatment, but rather

    collected after euthanasia of the animals as necessitated by tumor size end point.

    Sections were stained with H&E and observed under a light microscope. ................ 134

    Figure 5.1 A) Reaction of starch with DTPA-bisanhydride to form St-DTPA by direct acylation

    of starch hydroxyl groups. DTPA-bisanhydride can react with one starch molecule to

    form St-DTPA or two starch molecules to form ST-DTPA-St. B) Possible chemical

    structures of PolyGd. .................................................................................................. 150

    Figure 5.2 FTIR spectra of A) starch, B) DTPA, C) St-DTPA.. ................................................ 152

    Figure 5.3 1H NMR spectra of A) DTPA, B) starch, C) St-DTPA. Major peaks for DTPA and

    starch have been assigned on the molecular schemes.. ............................................... 153

    Figure 5.4 Normalized differential heat (NDH) curves titrating 2 mM Gd3+

    into 0.1 mg/ml

    aqueous buffer solution of St or St-DTPA at 25 ºC and pH 5.6. Titration of Gd3+

    into

    starch did not produce any heat while titration into St-DTPA produced a large

    endothermic heat indicating binding of Gd3+

    to St-DTPA. Assuming one Gd3+

    ion

  • xviii

    binds to one DTPA molecule, the amount of covalently bound DTPA to starch is

    calculated at the titration end point (inflection point) indicated by an arrow. ............ 155

    Figure 5.5 The toxicity of saline, blank polymer, PolyGd, and free Gd3+

    to rat hepatocytes in

    culture exposed for 30 min, 60 min, 120 min, or 240 min. “%live” represents the

    percent of hepatocytes excluding trypan blue.. ........................................................... 156

    Figure 5.6 Coronal T1-weighted (3D-FLASH, TE/TR 3/25 msec, flip angle 20º) whole body

    images of Balb/c mice injected with Omniscan® (0.1 mmol/kg Gd3+

    ) and PolyGd

    (0.025 mmol/kg Gd3+

    ). At one fourth the dose of Omniscan®, the PolyGd produces a

    much higher contrast over an extended period of time in the cardiovascular system. 160

    Figure 5.7 Quantitative MRI of whole-body distribution: A) R1 maps of Balb/c mice injected

    with PolyGd (0.025 mmol/kg Gd+3

    ). B) Change in relaxation rates, ∆R1, of left

    ventricular blood, liver, bladder, and kidneys for Omniscan® (0.1 mmol/Kg Gd3+

    ) and

    PolyGd (0.025 mmol/Kg Gd3+

    ) overtime relative to baseline. The Gd3+

    loaded polymer

    causes a much higher increase in blood relaxation rate for an extended period of time

    compared to Omniscan®.. .......................................................................................... 162

    Figure 5.8 MR angiography: A) MIP angiogram displaying contrast enhancement of (1) whole

    body and (2) neck and head regions, obtained prior to and at 15 minutes following

    PolyGd injection at 0.025 mmol/kg Gd3+

    . B) Kinetics of vascular signal to noise (S/N)

    ratio and contrast to-noise (C/N) ratio measured from the inferior vena cava in whole-

    body angiograms.. ....................................................................................................... 164

    Figure 5.9 Tumor distribution of PolyGd (0.025 mmol/kg Gd3+

    ): A) T1-weighted images (1) and

    the corresponding R1 maps (2). B) Time course of ∆R1 in tumor periphery and tumor

    core, displaying elevated tumor R1 even 48 hours after contrast agent injection.. ..... 166

    Figure 5.10 Biodistribution, elimination and tumor accumulation of the PolyGd (0.025 mmol/kg

    Gd3+

    ) in tumor bearing Balb/c mice. The Gd3+

    content was determined using ICP-

    AES.. ........................................................................................................................... 168

    Figure 6.1 A) 1H NMR spectra of 1) PS 80, 2) PMAA-PS 80-g-St in 0.1M NaOD. B) polymer

    composition and physical properties of the PMAA-PS 80-g-St nanoparticles. For size

  • xix

    measurements, the particles were disperse in PBS pH of 7.4 and ionic strength of 150

    mM. For ξ-potential measurements PBS buffers of 7.4 and ionic strength of 10 mM

    was used ...................................................................................................................... 180

    Figure 6.2 A) Schematic diagram of self-assembly of PMAA-PS 80-g-St terpolymer into

    nanoparticles upon complexation with Gd3+

    and conjugation of the fluorescent

    moieties. B) TEM images of the self-assembled nanoparticles in water. ................... 182

    Figure 6.3 Negative TOF-SIMS spectra of PS 80, PMAA-g-St, and PMAA-PS 80-g-St, in the

    m/z range of 0 to 300 atomic mass units. ................................................................... 183

    Figure 6.4 Quantitative MRI of brain distribution: A) R1 maps of Balb/c mice (n=3) injected with

    Gd3+

    loaded PMAA-PS 80-g-St nanoparticles (0.05 mmol/kg Gd+3

    ). B) Longitudinal

    relaxation rates (R1) of sagittal sinus, ventricles, cortex, and sub-cortex for Gd3+

    loaded

    PMAA-PS 80-g-St-DTPA polymer overtime.. ........................................................... 185

    Figure 6.5 Qualitative and quantitative results of brain distribution and accumulation for PMAA-

    PS 80-g-St nanoparticles. A) Ex-vivo near infrared fluorescence images of the whole

    brain. Ratio of the relative fluorescence intensity in brain as a function of time after

    intravenous injection of nanoparticles compared to normal brain not injected with

    nanoparticles. B) Fluorescence microscopy image of perfused mice brains 45 minutes

    following iv administration of saline, PMAA-g-St, and PMAA-PS 80-g-St. The

    particles can be detected in the perivascular regions of the brain capillaries for samples

    treated with PMAA-PS 80-g-St nanoparticles. ........................................................... 188

    Figure 7.1 SEM images of (A) surface (left-bottom corner) and cross section (right-top corner),

    (B) surface, (C) cross section of cross-linked starch-xanthan gum film containing 10%

    xanthan gum and 5% STMP. ...................................................................................... 199

    Figure 7.2 FTIR spectra of (A) pure starch, (B) starch reacted with 20% STMP, (C) pure xanthan

    gum, (D) xanthan gum reacted with 20% STMP, (E) physical mixture of starch and

    xanthan gum, (F) physical mixture of starch, xanthan gum, and 20% STMP, and (G)

    mixture of starch and xanthan gum reacted with 20% STMP .................................... 201

  • xx

    Figure 7.3 Schematic representation of cross-linking reaction of starch and xanthan gum with

    sodium trimethaphosphate (STMP) ............................................................................ 202

    Figure 7.4 A) 31

    P NMR spectra of (I) pure starch, (II) starch reacted with 5% STMP, (III) pure

    xanthan gum, (IV) xanthan gum reacted with 5% STMP. B) 31

    P NMR spectra of (I)

    starch-xanthan gum without STMP, (II) STMP, (III) physical mixture of starch,

    xanthan gum and STMP, (IV) starch and 5% xanthan gum reacted with 5% STMP. C)

    31P NMR spectra of (I) starch and 10% xanthan gum reacted with 2% STMP, (II)

    starch and 10% xanthan gum reacted with 5% STMP, (III) starch and 10% xanthan

    gum reacted with 20% STMP. .................................................................................... 205

    Figure 7.5 Swelling kinetics of cross-linked starch-xanthan gum films containing 5% xanthan

    gum and 2% or 10% STMP in 0.15M phosphate buffer of pH=7.4 ........................... 207

    Figure 7.6 Equilibrium swelling behavior of modified starch xanthan gum films of various

    compositions with respect to change in (A) xanthan gum concentration in DDIW (B)

    pH with constant ionic strength of 0.15M. ................................................................. 208

    Figure 7.7 Relative diffusion coefficients of molecular probes as a function of molecular radius

    across modified starch-xanthan gum films with various STMP levels. ..................... 211

    Figure 7.8 A plot of Y/Q-1 as a function of the hydration factor (Q-1)-1

    for starch-xanthan gum

    hydrogels of various compositions. The slope of the line is equal to the scale factor, Y.

    .................................................................................................................................... 212

  • xxi

    List of Abbreviations

    1H-NMR hydrogen nuclear magnetic resonance spectroscopy

    31P-NMR phosphorous nuclear magnetic resonance spectroscopy

    AcAn acetic anhydride

    ALT alanine aminotransferase

    α-MEM alpha-modified minimal essential medium

    ANOVA analysis of variance

    ATP adenosine-5’-triphosphate

    AUC area under the curve

    BBB blood brain barrier

    CCAC Canada council on animal care

    CK creatine kinase

    CO2 carbon dioxide

    D2O deuterated water

    DAPI 4’,6-diamino-2-phenylindole

    DDIW distilled deionized water

    DLS dynamic light scattering

    DMSO dimethyl sulfoxide

    DNA deoxyribonucleic acid

    Dox doxorubicin

    DTPA diethylenetriaminepenta acetic acid

    DTPA-bis-An diethylenetriaminepenta acetic acid bisanhydride

    EDC N-(3-Dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride

    EE encapsulation efficiency

  • xxii

    EPR enhanced permeability and retention effect

    EtOH ethanol

    FA fluoresceinamine isomer I

    FBS fetal bovine serum

    FDA food and drug administration

    FITC fluorescein isothiocyanate

    FTIR Fourier transform infrared spectroscopy

    Gd gadolinium

    GD growth tumor delay

    GSH glutathione

    H&E hematoxylin and eosin

    HCl hydrochloric acid

    HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid

    HF 750 HiLyte Fluor TM

    750 hydrazide

    IC50 inhibitory concentration for 50% effect

    ITC isothermal titration calorimetry

    Kel elimination rate constant

    KPS potassium persulfate

    LDH lactate dehydrogenase

    λem emission wavelength

    λex excitation wavelength

    LC loading content

    LRP lung resistance protein

    LV left ventricle

    MAA methacrylic acid

    MBA N,N′-Methylenebisacrylamide

  • xxiii

    MDR multidrug resistance

    MPS mononuclear phagocytic system

    MRI magnetic resonance imaging

    MRP1 multidrug resistance protein 1

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

    NADPH nicotinamide adenine dinucleotide phosphate

    NaOH sodium hydroxide

    NHS N-Hydroxysuccinimide

    NIR near-infrared

    NPs nanoparticles

    PBS phosphate buffered saline

    PEG polyethylene glycol

    PET positron emission tomography

    PF-NPs preformed nanoparticles

    Pgp P-glycoprotein

    PMAA polymethacrylic acid

    PMAA-PS 80-g-St polymethacylic acid grafted starch

    PS 80 polysorbate 80

    Py pyridine

    Ri relaxation rate

    ri relaxivity

    ROS reactive oxygen species

    S/V surface area to volume ratio

    SA-NPs self-assembled nanoparticles

    SD standard deviation

    SDS sodium dodecyl sulphate

  • xxiv

    SEM standard error of the mean

    St starch

    STMP sodium trimethaphosphate

    STS sodium thiosulfate

    t1/2 half life

    TEM transmission electron microscopy

    v/v volume by volume

    w/v weight by volume

    XG xanthan gum

    XRPD x-ray powder diffraction

    ξ zeta potential

  • 1

    Chapter 1. Introduction

  • 2

    1.1 Breast Cancer

    1.1.1 Epidemiology

    Cancer is a serious and prevalent health problem that affects us all. There were 12,661,000 new

    cases of cancer in 2008, and 7,564,000 people had died from previous cancer diseases [1]. In

    2011, 177,800 Canadians were newly diagnosed with cancer, and cancer related deaths

    accounted for 29.3% of total deaths in Canada [2].

    Breast cancer is the most prevalent form of cancer among women. There were 1,384,000 new

    cases in 2008, making up 23% of all women’s cancer [1]. With 458,000 deaths worldwide, breast

    cancer is considered the most frequent cause of cancer related death in women. According to

    Canadian Cancer Society, there were 23,600 new cases of breast cancer in 2011, making it the

    most common cancer in Canadian women. [2]. Although the mortality rates are on a gradual

    decline in most developed countries since 1990, mainly due to better screening techniques and

    more effective treatment strategies, the disease claimed the lives of 5,100 Canadian women,

    comprising 14.8% of total cancer deaths, second only to lung cancer [2].

    1.1.2 Breast Cancer Cells and Tumor Microenvironment

    Breast cancer, as in other solid cancers, is fundamentally characterized by continuous and

    uncontrolled growth. Many researchers share the view that tumorigenesis proceeds via a process

    analogous to Darwinian evolution, in which a succession of genetic changes, each conferring one

    or another type of growth advantage, leads to the progressive conversion of normal human cells

    into cancer cells. Hanahan and Weinberg have proposed that six essential alterations in cell

    physiology enabled by the genetic instability result in malignant growth [3, 4]. These include

    continuous proliferation independent of cells’ microenvironment, acquired resistance to anti-

    proliferative signals, evasion of apoptosis, ability to replicate without limits, ability to induce

  • 3

    angiogenesis, and the ability of some cancer cells to undergo metastasis. These acquired

    physiological alterations generally apply to solid tumors, including breast cancer.

    Breast tumors are not merely an accumulation of neoplastic cells, but rather a complex tissue

    with blood vessels, stromal cells, infiltrating immune-competent cells, and a differentiated

    extracellular matrix [3-5]. All these cell types interact with each other to build a unique tumor

    microenvironment. This heterogeneous mass grows until it reaches an approximate volume of 2

    mm3, beyond which the diffusion of nutrients and oxygen cannot take place and areas of hypoxia

    and acidosis develop [6, 7]. As a result, the tumor contains interspersed regions of well

    oxygenated (pO2 > 2.5 mmHg) and poorly oxygenated (pO2 ≤ 2.5 mmHg), or hypoxic, tissue

    heterogeneously distributed throughout the tumor mass. Cancer cells of a breast tumor can adapt

    to thrive in areas of low oxygen concentrations that would otherwise induce normal cell death [8,

    9]. Normal cells fulfill 90% of their energy requirements through the Krebs cycle which uses

    pyruvate formed from glycolysis in a series of reactions that donate electrons via NADH and

    FADH2 to the respiratory chain complexes in mitochondria [8]. This high efficiency glucose

    metabolism requires oxygen. With limited oxygen, such as with muscles that have undergone

    prolonged exercise, pyruvate is not used in the Kreb’s cycle and is converted into lactic acid by

    lactate dehydrogenase (LDH) in a process termed anaerobic glycolysis. This process can also

    produce cellular energy but with poor efficiency. Warburg was the first one to report that even in

    the presence of oxygen, 50% of tumor ATP is produced through glycolytic catabolism [8]. This

    switch from high efficiency aerobic oxidative phosphorylation to low efficiency anaerobic

    glycolysis for cellular chemical energy production in tumors is known as the Warburg effect [8,

    9]. Through reduction of the oxygen consumption, the Warburg effect enhances cancer cell

    survival in hypoxic tumor microenvironment by decreasing the oxygen-starved fraction of the

    tumor distal from blood vessels, and by reducing the production of the reactive oxygen species

  • 4

    (ROS) that are the by-product of electron transport in the mitochondria during the Kreb cycle

    [10]. Due to cytotoxic nature of the ROS, reduction in their production by the Warburg effect

    provides cancer cells with a survival advantage in hypoxic tumor tissues [11].

    Tumor growth beyond 2 mm3 is angiogenesis dependent. Angiogenesis is defined as the process

    of formation of the new blood vessels. The hypoxic nature of the tumor environment activates a

    series of hypoxia-sensitive transcription factors in cancer cells, stromal fibroblasts, and tumor

    associated macrophages. These cells all work in harmony to generate new tumor blood vessels

    [12-14]. Formation of new vascular network by angiogenesis supply the tumor with oxygen and

    nutrients required to support its continued growth. However, tumor vasculature is significantly

    different in terms of its structure and physiology from healthy blood vessels. Many tumors reveal

    chaotic networks of tortuous and distended veins, venules and venous capillaries along with

    intertwining capillaries branching from arterioles and veins [15, 16]. Irregularities of vascular

    wall structures in tumors have also been described with walls composed of a mosaic of cancer

    and endothelial cells which leads to widened interendothelial junctions and numerous endothelial

    fenestrations with the net effect of leaky blood vessels unique to tumor neovasculature [17].

    These blood vessels are highly permeable to the extravasation of therapeutic macromolecules

    and small colloidal particles [18, 19].

    The resulting intratumoral circulation is characterized by tortuous microvessels lacking the

    normal hierarchical arrangement of arterioles, capillaries and venules. Within this altered

    microenvironment, blood flow is sluggish with unstable rheology, anomalous and generally

    stagnant [15, 16]. As a result, these characteristics lead to heterogeneous perfusion with hypoxia

    and acidity in low-flow regions. In fact, hypoxia is a pathophysiological property of breast

    tumors, with up to 60% of the tumor existing in a hypoxic state [7].

  • 5

    Accumulation of lactic acid in the tumor microenvironment coupled with insufficient blood

    supply and poor lymphatic drainage results in acidic pH states in the solid tumor

    microenvironment. Although there is a distribution, in vivo pH measurements, made by needle

    type microelectrodes on human patients having various solid tumors (adenocarcinoma, squamous

    cell carcinoma, soft tissue sarcoma, and malignant melanoma) in readily accessible areas (limbs,

    neck, or chest wall), show the mean pH value to be 6.9 with values as low as 5.7 being reported

    for some tumors [20, 21]. Increasingly, it has been proposed by many researchers that the mildly

    acidic tumor environment can be exploited to achieve high local drug concentrations and to

    minimize overall systemic exposure [22, 23] . The use of pH responsive carrier systems in the

    delivery of chemotherapeutics will be discussed in section 1.1.5.3.

    1.1.3 Doxorubicin: a Potent Drug for Breast Cancer Chemotherapy

    Doxorubicin (Dox), Figure 1.1, is a chemotherapeutic agent which belongs to the anthracycline

    antibiotics family. The drug has been widely adopted as a first line chemotherapy agent, most

    often in combination with other agents, for the treatment of various types of cancer including

    hematological malignancies, carcinomas, and sarcomas [24, 25]. Its widespread use in the

    treatment of breast cancer has been recognized as one of the reasons for the decreased mortality

    rates of the disease [26]. A number of mechanisms have been proposed to account for the broad-

    spectrum anticancer activity of Dox. These include: (1) DNA intercalation by its planar three-

    ring structure, (2) poisoning topoisomerase II enhancing DNA strand breaks, (3) generation of

    free radicals, (4) possible disruption of the cell membrane functionality [27-29]. The cell nucleus

    is the main cellular target of Dox even though involvement of mitochondria in Dox cytotoxicity

    such as mitochondria-mediated apoptosis has also been observed [30]. It is likely that multiple

    pathways are used by Dox to inflict damage to cancer cells.

  • 6

    Dox is commonly administered intravenously in the form of commercially available injections

    Adriamycin® and Rubex® for maintaining the therapeutic levels in blood. In addition, two

    PEGylated liposomal formulations of Dox, Doxil® and Caelyx®, are also available.

    Dox administration in breast cancer chemotherapy is often limited by serious systemic side

    effects such as myelosuppression and congestive heart failure [31-33]. The Dox induced

    cardiomyopathy appears to be cumulative limiting the maximum lifetime dose of the drug in

    humans to 450 mg/m2 [33]. The detailed mechanism of cardiotoxicity is not fully understood;

    however, it is believed that Ca2+

    activated ATPase, cAMP, and lipid peroxidation are involved,

    as well as the ability of Dox to produce free radicals. Myocardium cells seem to have an

    enhanced sensitivity to these effects leading to cardiomyopathy, and decreased cardiac

    ventricular ejection fraction [29, 31, 32, 34].

    Although Dox is highly effective in a variety of non-resistant breast cancer cell lines, its efficacy

    is significantly reduced in multi-drug resistant (MDR) sub-cell lines such as murine

    EMT6/AR1.0 and human MDA435/LCC6/MDR1. Hence Dox is an excellent compound for

    evaluation of novel MDR-reversal approaches.

    Figure 1.1 Chemical structure of Doxorubicin

  • 7

    1.1.4 Barriers to Cancer Chemotherapy

    1.1.4.1 Drug Resistance

    Drug resistance is one of the major obstacles in successful and effective treatment of breast

    cancer. The underlying causes of drug resistance are complex and multi-factorial providing the

    cancer cells with many ways to survive cancer chemotherapy. In general, the mechanisms of

    drug resistance can be classified into non-cellular resistance and cellular resistance. These two

    mechanisms are described in more detail below.

    Non-cellular resistance: Poor efficacy in classical chemotherapy often is associated with the

    unique anatomical or physiological features of the solid tumors. It is a well-known fact that

    multi-cellular spheroids of tumor cells are more resistant to anticancer agents than the

    corresponding monolayer cultures [35]. Solid tumors are also found to be less sensitive to

    chemotherapy than malignancies consisting of individual cells, e.g. leukemia. As discussed in the

    previous section, the tumor environment and blood vessel architecture is unique and

    characterized by hypoxic regions, acidic microenvironment, sluggish, and inhomogeneous blood

    circulation. Most blood vessels inside tumor are highly disorganized as they take tortuous turns

    and many of these twisted blood vessels near the center of tumor, are crushed due to the irregular

    growth of tumor in a confined space. Some sites in the tumor are thus far away from the blood

    supply making it difficult for drug molecules, especially those with larger molecular weight and

    lipophilicity, to reach these sites [36]. In addition, due to the elevated vascular permeability of

    tumor vessels along with the absence of a functional lymphatic network, in this confined space,

    interstitial fluid is collected more abundantly than in normal tissue, creating another impediment

    to drug distribution inside tumor mass. The accumulation of this fluid leads to elevated

    interstitial fluid pressure that decreases from the core toward the periphery of the tumor [37]. The

  • 8

    altered composition of extracellular fluid, associated with increased interstitial pressure and

    sluggish blood flow hinders drug distribution and therefore the efficacy of chemotherapy.

    As discussed previously poor blood supply and excessive growth leads to development of

    hypoxic regions in tumors. Hypoxic cells may become quiescent and resistant to cytotoxic agents

    that target proliferating cells [38]. The reduction in oxygen levels also compromises the anti-

    cancer efficacy of some drugs such as Dox which is known to be much more effective in

    oxygenated regions of the tumor by generating ROS species through the redox cycling of its

    quinine nucleus [28]. Hypoxia, especially chronic hypoxia, can promote metabolic changes and

    genomic instability of cancer cells, further enabling the acquisition of random mutations and

    driving the malignancy of the tumor [38, 39]. Furthermore, the acidic environment of the tumor

    due to production of lactic acid may potentially deactivate weak base drugs such as doxorubicin

    [40].

    In summary, non-cellular resistance is one of the major causes of poor efficacy of conventional

    chemotherapeutic agents in solid tumors as only a limited fraction of the systemically

    administered anticancer drug can reach and penetrate the target tumor site and remain active.

    Cellular resistance: The cancer cells may still survive despite the significant levels of drug in

    the tumor. Multidrug resistance (MDR) is a major barrier to effective treatment of cancer. This is

    due to the ability of cancer cells to effectively neutralize the cytotoxicity of classical agents such

    as Dox. MDR can be acquired following failed rounds of drug therapy, or it can be innate, pre-

    programmed into the genetic code of the cancer cells [40, 41, 42]. In addition, the drug resistance

    in MDR cells is rarely specific to one drug and generally develops as cross-resistance to a range

    of structurally and functionally unrelated compounds [43]. Hydrophobic, amphiphathic drugs

    such as vinca alkaloids (vincristine, vinblastine), taxanes (paclitaxel, docetaxel),

  • 9

    epipodophyllotoxins (etoposide, teniposide), anthracyclines (doxorubicin, daunorubicin) are

    more frequently associated with MDR phenomena. MDR is often attributed to cell membrane

    drug transport proteins, metabolic pathways, and intracellular targets common to cancer

    chemotherapeutics [41, 43, 44].

    Over-expression of ATP Binding Cassette (ABC) efflux pumps is one of the major mechanisms

    of the MDR. P-glycoprotein (P-gp) and Multidrug Resistance Protein 1 (MRP1) are two main

    ABC efflux pumps that are over-expressed in drug resistant breast cancers [45, 46]. Like all

    ABC family members, both P-gp and MRP1 utilize the energy released from ATP hydrolysis to

    transport drug molecules in the cell membrane to the outside. P-gp and MRP1 works in concert

    to effectively decrease the intracellular concentration of the drugs before they become active and

    reach their cellular target [41]. P-gp preferentially transports neutral or mildly cationic

    compounds while MRP1 is more effective against lipophilic anions (e.g. glucuronate or

    glutathione conjugates) [45]. The expression of both pumps is controlled by hypoxia-inducible

    transcription factors, and is responsive to changes in intracellular redox status and ROS

    production [47, 48]. Hence, the hypoxic nature of the tumor microenvironment as well as the

    redox cycling associated with certain cytotoxic drugs such as Dox can upregulate the expression

    of these plasma membrane efflux pumps induce acquired MDR. It has been suggested that

    certain non-ionic surfactants (e.g. polysorbate 80), and block co-polymers (e.g. PluoronicTM P85)

    may inhibit the P-gp efflux pumps [49]. In addition to P-gp and MRP1, breast cancer resistance

    (BCRP) has also been reported in some resistant breast cancer cells [45]. More than one of the

    mentioned membrane-associated drug transports may be present in the same cancer cell and

    render the cell even more resistant to chemotherapy.

  • 10

    Alternatively, cancer cells may become resistant by sequestration of drugs in cytoplasmic

    vesicles. Using fluorescence microscopy and taking advantage of the fluorosence properties of

    Dox, Beyer et al. have demonstrated that the subcellular distribution of the drug following cell

    uptake was significantly different between chemosensitive and MDR cell lines with higher

    accumulation of the drug in the cytosolic vesicles being observed in MDR cells [50]. By

    preventing the drug from interacting with its cellular targets, MDR cancer cells can effectively

    neutralize the chemotherapeutic agents.

    MDR can also be the result of cell ability to evade apoptosis. Apoptosis is the process of

    programmed cell death or cell suicide that can be elicited by a number of stimuli such as DNA

    damage caused by certain anticancer agents (e.g doxorubicin and cisplatin) [51]. The apoptotic

    pathway involves highly organized and specific signal transduction that is negatively or

    positively regulated by anti-apoptotic factors and pro-apoptotic factors. Pakunlu et al.

    demonstrated that Bc12, an anti-apoptotic factor, is upregulated in Dox resistant tumor cells [52].

    By avoiding the occurrence of apoptosis, cancer cells become less sensitive to certain

    chemotherapeutic agents.

    Some drug-resistance cancer cells were also found to have more active drug detoxification

    systems. Glutathione (GSH) and glutathione-S-transferase act to sequester a number of

    anticancer agents and are overexpressed in MDR cells [42]. An increase in intracellular GSH

    facilitates the removal of the drug-GSH conjugates from the cell via MRP1 which is also

    upregulated in MDR cells [48].

    Cancer cells may also protect themselves from the action of cytotoxic drugs by alteration in drug

    targets themselves. By altering the conformation of the drug target or masking the drug docking

    site through post-translational modification, the efficiency with which the target is modified or

  • 11

    its function disrupted by the drug is decreased, rendering the cell resistant to the action of the

    drug [41]. For example, altered tubulin structure may compromise the effectiveness of anti-

    microtubule agents [53].

    Many anticancer drugs exert their action by inflecting damage to DNA. The resistant cells often

    modify their DNA repair mechanisms through upregulation of the DNA repair proteins [44].

    Increased repair of Dox-DNA adducts would decrease the rate of induced DNA lesions and

    promote the restitution of DNA structure, overcoming Dox genotoxicity. In addition, DNA

    damage signalling pathways responsible for relaying the information regarding genotoxicity to

    effectors of apoptosis are shut down, rendering the cells resistant to the cytotoxic drugs [41, 44].

    As discussed, clinical drug resistance is a multi-factorial phenomenon, and the development of

    MDR phenotype in cancer patients may involve any combination of the several physiological

    alterations outlined above (Figure 1.2). While MDR may be acquired through different

    mechanisms in different patients, the development of multidrug resistance is almost universal.

    Figure 1.2 Factors leading to MDR in cancer. One of the leading causes of treatment failure in

    cancer is the onset of resistant disease. MDR can be divided into two broad categories: Cellular

    resistance and non-cellular resistance.

    Factors leading to multi-drug resistance

    Cellular resistance

    Drug efflux pumps

    up-regulation

    Intracellular Sequestration

    Drug detoxification

    enzymes

    up-regulation

    Increased DNA repair

    capacity

    Anti-apoptotic proteins

    up-regulation

    Drug targets down-

    regulation

    Non-cellular resistance

    Tumor Hypoxia

    and acidity

    Sluggish blood flow

    Low intratumoral

    drug concentration

    Elevated interstitial

    fluid pressure

  • 12

    1.1.4.2 Systemic Side Effects

    Chemotherapeutics are often non-specific to cancer cells. A systemically administered cytotoxic

    drug causes toxicity in both normal and cancer cells. As a result, a major limitation of cancer

    chemotherapy is the systemic toxicity of the agents which ultimately influences their therapeutic

    efficacy. While most of the classical chemotherapeutics such as Dox are more toxic towards

    highly proliferating cells, these drugs will still induce toxicity in other cell populations due to

    their ability to produce ROS through redox cycling reactions [28]. In addition to cancer cells,

    other cells in the body such as the cells of gastrointestinal tract, hair follicles, and bone marrow

    are continually proliferating in an adult, reducing the selectivity of classical chemotherapy.

    Hence, some side effects such as nausea, vomiting, alopecia, and myelosuppression are common

    for all cytotoxic drugs. While other side effects are more drug specific. For example, Dox

    administration is associated with irreversible cardiotoxicity that is attributed to its major phase I

    metabolite doxorubicinol, limiting the maximum allowable lifetime dose of the drug [54, 55]. It

    has been suggested that Dox induces cardiomyopathy and congestive heart failure by interfering

    with the sarcoplasmic reticulum. Co-administration of the cardio-protectants or the use of

    anthracycline derivatives with lower demonstrated cardiotoxicity have been attempted by

    different researchers [56-58]. However, these approaches have often led to reduced anti-cancer

    efficacy, preventing effective anti-cancer outcomes.

    Because anticancer drugs often show a steep dose-response curve, it is has been suggested by

    some clinicians that aggressive chemotherapy is associated with greater therapeutic efficacy [59],

    but high doses of chemotherapeutics are generally associated with high levels of acute and

    chronic toxicity. In order to improve therapeutic efficacy, the drug associated systemic toxicity

    needs to be reduced. This may be achieved through limiting the drug exposure to the healthy

  • 13

    cells while increasing the drug concentration in the cancer cells. This rational forms the basis for

    the use of nanotechnology as drug delivery vector for cancer chemotherapy.

    1.1.5 Nanoparticulate Systems in Cancer Therapy

    Despite outstanding progress in the field of cancer biology and understanding the fundamentals

    of the disease, challenges remain in our ability to translate our understanding of fundamental

    cancer biology to the clinic. Classical chemotherapies suffer many challenges which include

    dose limiting systemic toxicity, hypoxia, relatively low intra-tumoral drug levels, and MDR

    phenotypes of cancer cells, often limiting their therapeutic potentials. Over the past two decades,

    nanotechnology-based approaches have emerged as an exciting field with promises to remedy

    these limitations. First, a well-designed nanoparticulate system with optimum size and

    circulation half-life would facilitate controlled and tumor specific drug accumulation and release,

    thus reducing the systemic toxicity that is often associated with classical chemotherapeutics and

    increasing the possibility of more aggressive chemotherapy and possibly better clinical outcome

    [60]. Second, the encapsulated drug is protected from the harsh environments of the body, drug

    metabolizing enzymes, and extensive binding to serum proteins while in the circulation leading

    to improved therapeutic efficacy [61]. Third, due to leaky nature of the tumor vasculature and its

    poor lymphatic drainage, drug carriers with optimum size (50-300 nm) and prolonged blood

    circulation will preferentially accumulate in the tumor, a phenomenon known as Enhanced

    Permeation and Retention effect (EPR) (Figure 1.3) [18, 19]. Finally, the sub-micron size of the

    drug carriers affords the ability to enter deep within the cell and overcome the MDR efflux pump

    mechanisms which are currently one of the major causes of treatment failure in the clinic (Figure

    1.4) [60, 62-67]. By designing an optimum nanoparticulate drug delivery system, the right drug

    can be delivered to the right site in the body at the right time.

  • 14

    Figure 1.3 Passive targeting of the tumor by EPR. The EPR effect is the balance of enhanced

    tumor permeability with poor tumor interstitial fluid drainage, resulting in the selective uptake

    and retention of nanoparticles in the tumor tissue [18, 19].

    Figure 1.4 Drug loaded nanoparticles can overcome MDR cancer cells. Endocytosis of the drug

    loaded nanoparticles in membrane bound vesicles protects the drug from the action of the

    membrane efflux pumps. The nanoparticles release the drug deep inside the cell and the drug can

    gain access to its cellular target site (e.g. DNA)

    Enhanced Permeability & Retention Effect (EPR)

    1. Tumor-associated leaky vasculature2. Poor lymphatic drainage

    Tumor

    Neovasculature

    Loose

    Interendothelial

    Junctions

    (Fenestrations)

    Blood

    Flow

    Lymph

    Flow Tumor

    ~ < 300 nm

    Blood

    pH =7.4

    Lysosomal

    pH < 6

    Tumor

    pH < 7Drug

    Dox efflux by

    P-gp pumps

    Leaky tumor

    vasculatureActive uptake of NPs

    by tumor cells

    Dox

  • 15

    1.1.5.1 Requirements of an Ideal Nanoparticulate Drug Delivery System for Cancer Chemotherapy

    Ideally, a drug carrier should have as many of the following properties as possible: (1) good

    biocompatibility profile, (2) biodegradable with non-toxic degradation products, (3) convenient,

    cost-effective, and reproducible preparation, (4) ability to efficiently load the drug at high

    contents, (5) controlled and tumor specific drug release kinetics, (6) optimum size and

    circulation half-life, (7) passive and/or active tumor targeting capabilities, and (8)

    “upgradability”, i.e., allowing further surface modifications.

    As mentioned previously, chemotherapeutics typically show a steep dose-response curve. To

    ensure therapeutic success sufficiently high dose intensity has to be used. In other words,

    reasonably high drug loading capacity and sufficient drug release at the diseased site are required

    or else an unreasonably large quantity of the drug carrier has to be administered for effective

    cancer treatment [68]. Cytotoxic drugs generally do not discriminate between cancer cells and

    healthy cells in the body; hence, an ideal drug delivery system should exhibit minimal non-

    specific drug release while in the circulation followed by increased release rate upon

    accumulation in the tumor. Size, blood circulation time, and colloidal stability of the

    nanoparticles are all important characteristics of the drug carrier. Nanoparticles which are

    smaller than 10 nm are rapidly cleared by the renal route preventing adequate time for tumor

    accumulation, while passive targeting by the EPR effect is significantly reduced with particles

    larger than 300 nm [69, 70]. The size range of 50-200 nm has found to be optimal in promoting

    the passive targeting of the nano-carriers to the tumor site. However, it has to be pointed out that

    the range of acceptable nanoparticle sizes for optimized chemotherapy is highly material

    dependent and will change from polymeric to inorganic to lipid based formulations. The long

    circulation of the nanoparticulate drug delivery system is required to increase the number of

  • 16

    exposures of tumor tissue to the drug delivery system, promoting the passive tumor targeting by

    EPR effect. In fact, it is believed that EPR effect-mediated passive tumor uptake is optimized

    with circulation times of at least 4 hours [61, 70]. Surface grafting of some hydrophilic polymers

    to the nanoparticles can prolong the circulation half-life of nanoparticles, and will be reviewed in

    the next section.

    1.1.5.2 Surface Modification to Prolong Nanoparticles Blood Circulation

    Plasma proteins, including immunoglobulins and complement proteins readily bind to foreign

    matters in the blood [61, 70]. These plasma proteins, or opsonins, effectively, tag the foreign

    particles for removal from the circulation by resident tissue macrophages of the liver, lymph, and

    spleen, known as the mononuclear phagocytic system (MPS) [71]. The removal of opsonised

    nanoparticles from the systemic circulation reduces the longevity of nanoparticle circulation and

    hence compromises its therapeutic efficacy. The surface of the nanoparticles can be modified to

    prevent this rapid uptake producing long circulating or “stealth” particles which promote both

    passive and/or active targeting. Vonarbourg et al. have reviewed the general conditions for the

    stealthiness of colloidal drug carriers [72]. These include having a small size, with a neutral and

    hydrophilic surface, as well as a thick, well-anchored and flexible coating.

    The use of hydrophilic polymer coatings to nanoparticulates is a common practice to achieve a

    long circulation time. The most commonly used polymer for this purpose is poly(ethylene

    glycol) (PEG) [73]. It is generally accepted that the PEGylation of the particle surface prevents

    opsonisation by enhancing the steric repulsion between nanoparticles’ surfaces and blood

    components, and by forming an inert, impenetrable polymer layer over the surface of the

    nanoparticles with the effectiveness mainly depending on the degree of surface coverage and

    molar mass [74]. Additionally, it has been suggested that the PEG coats allow the selective

  • 17

    adsorption of some serum proteins that are sometimes known as dysopsonins, i.e. plasma

    components which are believed to prevent opsonization. In other words, it may be the selective

    adsorption, and not the prevention of adsorption, that alters the pharmacokinetics and fate of

    nanoparticles [75, 76]. Nevertheless, it is the general consensus that the surface PEGylation of

    the nanoparticles prolongs their blood circulation.

    However, PEG is not biodegradable; also, chemical attachment of additional groups such as

    targeting moieties, and/or metal chelators, to PEGylated surfaces is difficult and involves

    complicated reaction schemes mostly due to the absence of reactive groups on the PEGylated

    surfaces. This is one of the reasons why polysaccharide coatings have been considered as an

    alternative to the PEG coatings. Additionally, oligo and polysaccharides may achieve active

    targeting per se since they have specific receptors in certain cells or tissues [77]. Moreover,

    polysaccharides display well-documented biocompatibilities and biodegradabilities, which are

    the desired basic characteristics for polymers used as biomaterials. Polysaccharides have been

    suggested as biocompatible polymer coatings for nanoparticles. For example, heparin coating of

    poly(methyl methacrylate), PMMA, nanoparticles exhibited an increased circulation half-life of

    5 h compared to only a few minutes for the bare nanoparticles [78]. Similarly, coating of

    superparamagnetic iron oxide nanoparticles (SPIONs) with dextran increased their half-life up to

    4.5 h [79]. However, there have been reports of hypersensitivity reactions upon administration of

    dextran in the clinic [80]. Hydroxyethyl starch (HES) is currently investigated at the industrial

    level as a biodegradable substitute for PEG, so that HESylation of proteins could substitute

    PEGylation [81]. In one study poly(lactic-co-glycolic acid), PLGA, nanoparticles stabilized with

    HES exhibited reduced human serum albumin (HSA) and fibrinogen (FBG) adsorption, and

    uptake by phagocytic cells in vitro [82].

  • 18

    In summary, the polysaccharide coating may provide steric protection against protein adsorption

    and macrophage uptake. Additionally, as polysaccharides offer many available reactive groups,

    active targeting could be obtained by grafting ligands onto the nanoparticle surface. Considering

    the very large variety of polysaccharides in nature, one could imagine the wide array of surface-

    engineered nanoparticles adapted to a given therapeutic and monitoring purpose.

    1.1.5.3 pH-responsive Nanoparticles in Cancer Chemotherapy

    As discussed previously, tumor environment is acidic with average pH of 6.9 but