250
Investigation of Amorphous Solid Dispersions of Poorly Water-soluble Drugs in Poly(2-Hydroxyethyl Methacrylate) Hydrogels for Enhanced Solubility and Controlled Release by Dajun Sun A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Pharmaceutical Sciences University of Toronto © Copyright by Dajun Sun (2014)

Investigation of Amorphous Solid Dispersions of Poorly ......soluble drugs. The first part of the study identifies physicochemical properties affecting the solid state and physical

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

  • Investigation of Amorphous Solid Dispersions of Poorly Water-soluble Drugs in Poly(2-Hydroxyethyl Methacrylate) Hydrogels for Enhanced Solubility and Controlled Release

    by

    Dajun Sun

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

    Pharmaceutical Sciences University of Toronto

    © Copyright by Dajun Sun (2014)

  • ii

    Amorphous Solid Dispersions of Poorly Water-soluble Drugs in Poly(2-Hydroxyethyl Methacrylate) Hydrogels for Enhanced Solubility and

    Controlled Release

    Dajun Sun

    Doctor of Philosophy

    Pharmaceutical Sciences

    University of Toronto

    2014

    Abstract

    The purpose of this study was to investigate the potential of applying amorphous solid dispersions

    (ASD) in crosslinked PHEMA hydrogels to enhance the dissolution behavior of poorly water-

    soluble drugs. The first part of the study identifies physicochemical properties affecting the solid

    state and physical stability of ASD of the model drug indomethacin (IND) in PHEMA hydrogels.

    The results of the second part show that ASD based on water-insoluble crosslinked PHEMA can

    maintain a high level of supersaturation over a prolonged duration via a diffusion-controlled

    feedback mechanism, thus avoiding the initial surge of supersaturation followed by a sharp decline

    in drug concentration, which is typically encountered with ASD based on water-soluble polymers

    (e.g., PVP, HPMCAS) under nonsink dissolution conditions. A subsequent study examines the

    effect of supersaturation generation rate on the resulting kinetic solubility profiles of amorphous

    pharmaceuticals and delineates the interplay between dissolution and precipitation processes from

  • iii

    a mechanistic viewpoint. In the absence of any dissolved polymer to inhibit drug precipitation

    from the supersaturated state, both our experimental and predicted results confirm that the faster

    rise of the kinetic solubility profile of an amorphous drug will inevitably lead to an earlier but

    higher maximum kinetic solubility and a sharper decline in the de-supersaturation phase, and vice

    versa. The relationship between the achievable maximum supersaturation and the rate of

    supersaturation generation in the observed kinetic solubility profiles has been described for the

    first time by our comprehensive mechanistic model taking into account the role of supersaturation

    in both the nucleation and crystallization processes as well as the associated competitive particle

    growth and ripening effects. Finally, this theoretical framework was further employed to semi-

    quantitatively predict the evolution of supersaturation of amorphous pharmaceuticals generated

    from nonlinear dissolution profiles. The effects of initial degree of supersaturation, dissolution of

    amorphous drug and that from the IND-PHEMA ASD under nonsink dissolution conditions were

    subsequently examined in detail. The comparison of dissolution behaviors between amorphous

    IND and IND-PHEMA ASD demonstrates the advantage of the diffusion-controlled feedback

    mechanism that makes crosslinked PHEMA a unique and desirable carrier for amorphous drug

    delivery systems.

  • iv

    Acknowledgements

    For those who studied in the disciplines of science and engineering, the term “problem set” should

    sound familiar. The journey of my Ph.D. study feels like a very long problem set in which an

    answer to a question indefinitely leads to and expands to more, larger and harder questions. This

    daunting task would not have been achievable without a great amount of guidance, support and

    assistance from many wonderful individuals. First and foremost, I would like to express my

    greatest gratitude to my Ph.D. supervisor Professor Ping Lee for his endless advice, support and

    supervision throughout my research project. His scientific insights and comprehensive expertise

    in pharmaceutical science has not only provided immense leadership to this project but also shaped

    me as an independent researcher. I have learnt many great personal characteristics from him,

    especially diligence and professionalism. I will always remain thankful to his generous education.

    Also, I am greatly thankful to my advisory committee members, Professors Edgar Acosta,

    Christine Allen and Shirley Wu, for providing me with valuable guidance and direction to my

    research as well as precious career advice. My thanks to Professor Carolyn Cummins for

    completing my experience as a graduate student by giving me a teaching opportunity. I am grateful

    to Dr. Rob Ju from Abbvie for a delightful experience of academia-industry collaboration. Special

    thanks are due to the helpful staff and students in the Department of Geology, the Department of

    Chemistry, University Health Network pre-formulation lab and Professor Allen’s lab for kindly

    assisting me to use their laboratory equipment.

    This research work was supported by research funding from Abbvie and Natural Sciences and

    Engineering Research Council of Canada (NSERC), and I was also supported by a University of

    Toronto Fellowship Award.

  • v

    I would like to thank the past and present colleagues in the PIL research group, Dr. Beibei Qu, Dr.

    Hui Zhao, Dr. Yan Li, Dr. Hongliang Jiang, Dr. Yanhong Luo, Sammi Liu, Arthur Li, Giovanna

    Medeiros and many others, for their continuous assistance and advice. Working together with them

    is always a nice and memorable experience. Lastly, I would like to thanks my family and friends,

    especially my parents, my brother Peter and Che-chien Wang for their unreserved faith in me. I

    could not have reached my goals without their unconditional love and support.

  • vi

    Publications

    - Dajun D. Sun and Ping I. Lee “Crosslinked hydrogel – a promising class of insoluble solid

    molecular dispersion carriers for enhancing the delivery of poorly soluble drugs” Acta

    Pharmaceutica Sinica B, Volume 4, Issue 1, pp 26-36 (2014). [Invited Review and Cover Story]

    - Dajun D. Sun and Ping Lee, “Evolution of supersaturation of amorphous pharmaceuticals: the

    effect of rate of supersaturation generation” Molecular Pharmaceutics, Volume 10, Issue 11, pp.

    4330-4346 (2013).

    - Dajun D. Sun, Tzu-chi Rob Ju, Ping I. Lee, “Enhanced kinetic solubility profiles of indomethacin

    amorphous solid dispersions in poly(2-hydroxyethyl methacrylate) hydrogels” European Journal

    of Pharmaceutics and Biopharmaceutics, Volume 81, Issue 1, pp. 149-158 (2012).

  • vii

    Table of Contents

    Abstract ii

    Acknowledgments iv

    Publications vi

    Table of contents vii

    List of tables xii

    List of figures xiv

    List of appendices xxi

    List of symbols xxiii

    List of abbreviations xxv

    Chapter 1: Introduction

    1.1 Solubility enhancement of poorly water-soluble drugs for oral drug delivery 1

    1.2 Pharmaceutical significance of amorphous solid dispersions in polymeric carriers 3

    1.2.1 Water-soluble carriers 7

    1.2.2 Water-insoluble carriers 14

    1.3 Crosslinked PHEMA hydrogels for amorphous solid dispersions carriers 16

    1.4 Crystallization of amorphous pharmaceuticals in the solid state 19

    1.4.1 Solubility advantage of amorphous solids 19

    1.4.2 Classical nucleation theory (solid state) 21

    1.4.3 Crystal growth (solid state) 23

    1.4.4 Kolmogorov-Johnson-Mehl-Avrami (KJMA) theory 24

    1.5 Crystallization of supersaturated drug solutions 25

    1.5.1 Classical nucleation theory (solution state) 25

    1.5.2 Crystal growth (solution state) 26

  • viii

    1.6 Overview of the Ph.D. research 27

    1.6.1 Hypothesis 27

    1.6.2 Research objectives 27

    Chapter 2: Indomethacin amorphous solid dispersions in PHEMA

    2.1 Introduction 29

    2.2 Materials and methods 31

    2.2.1 Materials 31

    2.2.2 Synthesis of PHEMA hydrogel beads 31

    2.2.3 Preparation of amorphous solid dispersion systems 34

    2.2.4 Scanning electron microscopy (SEM) 36

    2.2.5 X-ray diffraction (XRD) 36

    2.2.6 Differential scanning calorimetry (DSC) 37

    2.2.7 Fourier-transformed infrared (FTIR) spectroscopy 37

    2.2.8 Solubility parameter estimation 37

    2.3 Results and discussion 40

    2.3.1 PHEMA hydrogel beads synthesis 40

    2.3.2 Physical properties of ASD IND in PHEMA, PVP and HPMCAS 44

    2.3.3 IND-polymer interactions 55

    2.3.4 Solubility parameters 57

    2.4 Conclusion 59

    Chapter 3: Physical stability of amorphous indomethacin in PHEMA

    3.1 Introduction 60

    3.2 Materials and methods 62

    3.2.1 Materials 62

    3.2.2 Preparation of amorphous solid dispersion systems 63

    3.2.3 Stability study 63

    3.2.4 Preparation of physical mixtures of amorphous, crystalline - and -indomethacin

    in polymeric carriers 63

    3.2.5 Raman spectroscopy 64

    3.2.6 Multivariate data analysis 65

  • ix

    3.2.7 Isothermal crystallization kinetics 65

    3.2.8 Water sorption isotherm 66

    3.3 Results and discussion 66

    3.3.1 Storage stability study 66

    3.3.2 Quantification of ternary mixtures of different solid-state forms of indomethacin in

    polymeric carriers 68

    3.3.3 Crystallization kinetics of amorphous indomethacin in polymeric carriers 74

    3.3.4 Estimation of drug solubility in polymers 78

    3.3.5 Intermolecular forces between amorphous indomethacin and polymers 80

    3.3.6 Analysis of isothermal water vapor absorption 82

    3.4 Conclusion 86

    Chapter 4: Enhanced kinetic solubility profiles of amorphous indomethacin in PHEMA

    4.1 Introduction 87

    4.2 Materials and methods 88

    4.2.1 Materials 88

    4.2.2 Dissolution testing of amorphous solid dispersion systems under nonsink

    dissolution conditions 88

    4.3 Results and discussion 90

    4.3.1 Comparing solubility advantages of amorphous IND in PHEMA with that in water-

    soluble polymers 90

    4.3.2 Diffusion-controlled release of amorphous IND from PHEMA hydrogels 96

    4.4 Conclusion 101

    Chapter 5: Effect of rate of supersaturation generation on the kinetic solubility profiles

    5.1 Introduction 103

    5.2 Theory 106

    5.3 Materials and methods 111

    5.3.1 Materials 111

    5.3.2 Measurement of kinetic solubility profiles 111

    5.3.3 X-ray diffraction (XRD) 113

    5.3.4 Differential scanning calorimetry (DSC) 114

  • x

    5.3.5 Scanning electron microscopy (SEM) 114

    5.3.6 Particle size distribution 114

    5.3.7 Simulation of modeling equations 115

    5.4 Results and discussion 115

    5.4.1 Effect of rate of supersaturation generation on the kinetic solubility profiles 115

    5.4.2 Amorphous solid dispersions in polymeric carriers 127

    5.4.3 Kinetic solubility advantage of amorphous solids 129

    5.4.4 Crystallization kinetics and particle size distribution 132

    5.4.5 Evolution of concentration-time profiles due to dissolution and recrystallization

    processes 140

    5.5 Conclusion 142

    Chapter 6: Semi-quantitative prediction of the kinetic solubility profiles of amorphous

    indomethacin

    6.1 Introduction 144

    6.2 Theory 146

    6.2.1 Effect of linear rate of supersaturation generation 146

    6.2.2 Effect of initial degree of supersaturation 146

    6.2.3 Effect of first-order supersaturation generation 147

    6.2.4 Supersaturation generation with diffusion-controlled feedback mechanism –

    Dissolution of amorphous solid dispersions from PHEMA hydrogel beads 148

    6.3 Materials and methods 152

    6.3.1 Materials 152

    6.3.2 Preparation of amorphous indomethacin 152

    6.3.3 Dissolution testing of amorphous IND and ASD IND-PHEMA under nonsink

    dissolution conditions 152

    6.3.4 Dissolution testing of solid-state amorphous IND under sink dissolution

    conditions 153

    6.4 Results and discussion 154

    6.4.1 Effect of initial degree of supersaturation 154

    6.4.2 Dissolution of amorphous IND from the solid state 158

    6.4.3 Dissolution of amorphous IND from PHEMA hydrogels 163

  • xi

    6.5 Conclusion 170

    Chapter 7: Summary and future research directions

    7.1 Summary 172

    7.2 Future research directions 175

    Appendices 177

    References 202

  • xii

    List of Tables

    Chapter 1

    Table 1.1: Selected studies of water-soluble carriers for amorphous drugs... 11

    Table 1.2: Selected examples of insoluble carriers for amorphous drugs …. 13

    Chapter 2

    Table 2.1: Chemical structures of IND, PVP and HPMCAS and their

    potential hydrogen bonding sites ………………………………. 36

    Table 2.2: Solubility parameters of component group contribution from

    van Krevenlen/Hoftyzer and Hoy’s methods ………………….. 40

    Table 2.3: Measured and calculated solubility parameters of IND, PVP,

    HPMCAS, PHEMA and PHEMA copolymers ………………… 58

    Chapter 3

    Table 3.1: Comparison of existing technology of quantification of drug

    crystallinity in ASD ……………………………………………. 62

    Table 3.2 Physical mixtures of crystalline -, - and amorphous IND …… 64

    Table 3.3 Stability study of IND solid dispersions in PHEMA beads …..... 67

  • xiii

    Table 3.4 Raman molecular assignment for the C=O stretching group …... 70

    Table 3.5 KJMA isothermal crystallization parameters for amorphous

    IND ……………………………………………………………... 77

    Table 3.6 IND solubility in PHEMA, PVP and HPMCAS ……………….. 80

    Table 3.7 Characteristics of the model drug and polymers relevant for

    hydrogen bonding ……………………………………………… 81

    Chapter 5

    Table 5.1: Physicochemical properties of model poorly water-soluble IND,

    NAP and PIR …………………………………………………… 112

    Table 5.2 Summary of relevant physical constants for the numerical

    simulation of IND crystallization kinetics …………………… 120

    Chapter 6

    Table 6.1: Dissolution rate constants of amorphous IND of various particle

    size ranges ……………………………………………………… 161

  • xiv

    List of Figures

    Chapter 1

    Figure 1.1: Biopharmaceutics Classification System (BCS) of drugs ………………. 2

    Figure 1.2: Thermodynamic description of different solid states …………………… 4

    Figure 1.3: Gibb’s free energy levels of the amorphous state (metastable), crystalline

    (stable) and unstable state …..………………..………………………….. 5

    Figure 1.4: Schematics of substitutional, interstitial and polymeric solid solution ….. 6

    Figure 1.5: Classification of solid dispersion/solution of drug molecules in

    polymeric carrier matrix …………………………………………………. 7

    Figure 1.6: Dissolution performance of ASD containing a model poorly water-

    soluble compound ………………………………………………………. 9

    Figure 1.7 “Spring” and “parachute” dissolution behaviors ………………………… 10

    Figure 1.8: Free energy diagram for nucleation process ……………………………. 22

    Chapter 2

    Figure 2.1: Experimental apparatus for PHEMA hydrogel beads synthesis ……….... 33

    Figure 2.2: Microscopic images of PHEMA hydrogel beads ……………………….. 34

  • xv

    Figure 2.3: Quantification of IND by UV-spectrometer …………………………….. 35

    Figure 2.4: Particle size distributions of PHEMA hydrogel beads .………………….. 42

    Figure 2.5: Images of failed batches of crosslinked PHEMA hydrogels…………….. 43

    Figure 2.6: Equilibrium solvent content of PHEMA and IND solubility in

    DMSO/ethanol mixtures for IND loading process ……………………… 45

    Figure 2.7: IND loading levels in PHEMA hydrogel beads as a function of loading

    solution concentration ………………………….. ……………………… 46

    Figure 2.8: Residual solvent content in crosslinked PHMEA hydrogel after

    equilibrium sorption …………………………………………………….. 47

    Figure 2.9: Microscopic images of IND-loaded PHEMA hydrogel beads ………….. 48

    Figure 2.10: Microscopic images of IND-loaded PHEMA-co-MMA, -co-EMA and -

    co-BMA hydrogel beads ………………………………………………… 49

    Figure 2.11: Microscopic images of IND-loaded PVP and HPMCAS cast films ……. 50

    Figure 2.12: SEM images of surface morphology of IND-loaded PHEMA hydrogel

    beads …………………............................................................................... 51

    Figure 2.13: SEM images of cross-section of IND-loaded PHEMA hydrogel beads … 51

    Figure 2.14: XRD spectra of IND-loaded PHEMA, PVP and HPMCAS …………….. 52

    Figure 2.15: DSC thermograms of IND-loaded PHEMA, PVP and HPMCAS ………. 54

    Figure 2.16: Carbonyl stretch region of FTIR spectra of ASD and physical mixture

    IND-loaded PHEMA, PVP and HPMCAS ……………………………… 56

    Figure 2.17: Determination of solubility parameters of PHEMA hydrogels by Gee’s

    method of equilibrium swelling …………………………………………. 58

  • xvi

    Chapter 3

    Figure 3.1: Evolution of physical appearance of IND-PHEMA discs ………………. 67

    Figure 3.2: XRD spectra of IND-PHEMA ASD after 8-month stability study ……… 68

    Figure 3.3: XRD spectra of IND of different solid state forms and polymer carriers .. 69

    Figure 3.4: Raman spectra of IND of different solid state forms and polymer carriers 70

    Figure 3.5: Raman spectra and their statistical peak fittings of physical mixtures of

    crystalline -. - and amorphous IND in PHEMA, PVP and HPMCAS in

    the carbonyl stretching region ………………………………………….. 71

    Figure 3.6: Raman spectra and their statistical peak fittings of physical mixtures of

    different amount of -IND and amorphous IND in PHEMA …………… 72

    Figure 3.7: Calibration standard curves of intensity ratio between IND polymorphs

    and polymers …………………………………………………………….. 73

    Figure 3.8: Effect of relative humidity on crystallization kinetics of IND in PHEMA

    hydrogel (0-54% RH) ……………………………………………………. 75

    Figure 3.9: Effect of relative humidity on crystallization kinetics of IND in PHEMA

    hydrogel discs (76 and 95% RH) ……………………………………….. 75

    Figure 3.10: Effect of temperature on crystallization kinetics of IND in PHEMA …… 76

    Figure 3.11: Effect of the polymer type on crystallization kinetics of IND ………….. 78

    Figure 3.12: Water sorption isotherms of -form crystalline and amorphous IND,

    PHEMA, PVP and HPMCAS …………………………………………… 83

    Figure 3.13: Water sorption isotherm of ASD of IND (25% drug loading) in polymers 84

    Figure 3.14: DSC endotherms of polymers in various RH storage conditions ……….. 85

  • xvii

    Chapter 4

    Figure 4.1: Effect of drug loading on kinetic solubility profiles of ASD IND in

    PHEMA, PVP and HPMCAS …………………………………………… 91

    Figure 4.2: Effect of polymer type on kinetic solubility profiles of ASD IND in

    PHEMA, PVP and HPMCAS …………………………………………… 94

    Figure 4.3: Effect of particle size on kinetic solubility profiles of ASD IND-

    PHEMA …………………………………………………………………. 95

    Figure 4.4: Distribution of IND in the hydrogel and dissolution medium after 24 h of

    dissolution from IND-PHEMA ASD …………………………………… 97

    Figure 4.5: Effect of crosslinked PHEMA hydrogels in the dissolution medium on

    the equilibrium solubility of crystalline IND.…………………………… 98

    Figure 4.6: Amorphous drug release mechanism: crosslinked hydrogels vs. water-

    soluble polymers ………………………………………………………… 100

    Chapter 5

    Figure 5.1: Schematic depiction of the nucleation and crystallization events due to

    supersaturation generation by the infusion experiment …………………. 106

    Figure 5.2: Representative UV absorbance spectra and second derivative spectra for

    IND, NAP and PIR ………………………………………………………. 116

    Figure 5.3 Calibration curve of drug concentration and the second-derivative UV

    spectra for model drugs ………………………………………………….. 117

    Figure 5.4: Elimination of spectral interference by second-derivate UV technique ….. 118

    Figure 5.5: Experimental kinetic solubility profiles of IND, NAP and PIR as a

    function of supersaturation rate generated with infusion rates ………….. 119

  • xviii

    Figure 5.6: Comparison of experimental data and simulation results of kinetic

    solubility profiles of IND as a function of supersaturation rate …………. 123

    Figure 5.7: Predicted kinetic solubility profiles of IND as a function of

    supersaturation rate ……………………………………………………… 124

    Figure 5.8: Effect of rate of supersaturation generation on Cmax and AUC of kinetic

    solubility profiles …………………………….………………………….. 125

    Figure 5.9: Effect of rate schedule of supersaturation on kinetic solubility profiles.... 127

    Figure 5.10: Effect of rate schedule of solid addition on kinetic solubility profiles of

    IND-PVP ASD …………………………………………………………... 129

    Figure 5.11: Cmax measured during the infusion experiment as a function of the

    inverse of infusion rate raised to an exponent α …………………………. 130

    Figure 5.12: Crystallization kinetics of IND, NAP and PIR as a function of

    supersaturation rate generated with various drug solution infusion rates .. 133

    Figure 5.13: XRD spectra of precipitated IND from the infusion experiment ..………. 135

    Figure 5.14: DSC thermograms of precipitated IND from the infusion experiment ..… 136

    Figure 5.15: SEM images of precipitated IND from the infusion experiment ..………. 137

    Figure 5.16: Particle size distributions of precipitated IND from the infusion

    experiment ……………………………………………………………….. 139

    Figure 5.17: Simulated time-dependent growth of IND particle size under various

    rates of supersaturation generation ………………………………………. 139

    Figure 5.18: Conceptual concentration-time profile during dissolution and

    precipitation of amorphous sparingly soluble drugs …………………….. 141

  • xix

    Chapter 6

    Figure 6.1: Schematics of diffusion-controlled release of amorphous drug from a

    PHEMA hydrogel bead into a finite volume of dissolution medium ……. 149

    Figure 6.2: Kinetic solubility profiles of IND at various initial degrees of

    concentration …………………………………………………………….. 154

    Figure 6.3: COMSOL simulation of kinetic solubility profiles of various initial

    degrees of supersaturation ……………………………………………….. 156

    Figure 6.4: Cmax and AUC of in vitro dissolution data and simulation results of the

    kinetic solubility profiles of various initial degrees of supersaturation ….. 158

    Figure 6.5: Kinetic solubility profiles of amorphous IND of different particle size

    ranges under nonsink dissolution conditions…………………………….. 159

    Figure 6.6: Dissolution of amorphous IND with various particle size ranges under a

    sink dissolution condition ……………………………………………….. 160

    Figure 6.7: COMSOL simulation of kinetic solubility profiles of amorphous IND

    with various particle size ranges under nonsink dissolution conditions ... 162

    Figure 6.8: Comparison of IND dissolution profiles from purely amorphous IND

    versus that of IND-PHEMA ASD (10 wt % IND loading) ……………… 164

    Figure 6.9: Dissolution profiles over an extended time period from IND-PHEMA

    ASD (10 wt% drug loading) and crystalline IND ………..……………… 165

    Figure 6.10: Determination of the partition coefficient p between PHEMA hydrogel

    and external dissolution medium ………………………………………... 166

    Figure 6.11: Illustration of one-dimensional lengths for one hydrogel bead and the

    external dissolution medium in COMSOL simulation …………….….. 166

    Figure 6.12: Diffusion coefficient of IND in PHEMA hydrogels …….………………. 168

  • xx

    Figure 6.13: Comparison of COMSOL simulation results of kinetic solubility profiles

    between amorphous IND solids and IND-PHEMA ASD system ……….. 170

  • xxi

    List of Appendices

    Chapter 2

    Figure A2.1 XRD spectra of physical mixtures of IND and PHEMA with

    various IND weight percentage ………………………………. 177

    Figure A2.2 DSC thermograms of physical mixtures of IND and PHEMA

    with various IND weight percentage …………………………. 178

    Chapter 5

    Table A5.1 Parameter input in COMSOL 3.5a (Figures 5.6 and 5.7) ….…. 180

    Chapter 6

    Table A6.1 Parameter input in COMSOL 3.5a (Figure 6.3) ……………… 184

    Table A6.2 Parameter input in COMSOL 3.5a (Figure 6.7) ……………… 188

    Table A6.3 Parameter input in COMSOL 3.5a (Figure 6.13) …………….. 192

    Figure A6.1 COMSOL simulation results of time-dependent concentration

    profiles of IND inside PHEMA hydrogel beads under various

    sink conditions ……………………………………………….. 196

  • xxii

    Figure A6.2 IND release profiles in the external dissolution medium from

    Figure A6.2 …………………………………………………… 199

  • xxiii

    List of Symbols

    C concentration

    Cb bulk concentration

    Ceq concentration in equilibrium with a precipitate particle of radius r

    Cmax maximum concentration

    CS equilibrium solubility

    Ct concentration assuming all drugs have dissolved

    D diffusion coefficient

    d molecular size (twice the molecular diameter)

    Dose dose required

    f molar concentration of precipitate

    Fi molar attraction constant

    G crystallization kinetics coefficient

    J nucleation rate

    Jc nucleation rate constant

    k rate constant

    kB Boltzmann’s constant

    mo repeating monomer’s MW

    MW molecular weight

    N average particle number density

    No Avogadro’s number

    p partition coefficient

    Q equilibrium swelling ratio

    r particle size

  • xxiv

    R rate of drug input

    rn critical particle size

    s supersaturation

    s’ measured supersaturation at time t”

    T temperature

    t time

    t’ time in which addition of dissolved drugs in solvent stops

    t” time in which dissolution stops

    u precipitation rate

    V (molar) volume

    wd dry weight of hydrogels

    ws swollen weight of hydrogels

    δ solubility parameter

    ε thickness of Gibbs surface

    ρ density

    σ interfacial tension

    Σ average surface of the precipitate particles per unit volume

    υ molar volume of precipitate

    Φ average radius of the particles per unit volume

    ϕ volumetric fraction

    Ψ radius dependence of the surface tension

    ω capillary length

  • xxv

    List of Abbreviations

    AIBN Azobisisobutyronitrile [2,2′-Azobis(2-methylpropionitrile), 2-(azo(1-cyano-1-

    methylethyl))-2-methylpropane nitrile]

    API Active pharmaceutical ingredient

    ASD Amorphous solid dispersion

    AUC Area under the curve

    BCS Biopharmaceutics classification system

    BMA Butyl methacrylate

    CAP Cellulose acetate phthalate

    CNT Classical nucleation theory

    DSC Differential scanning calorimetry

    DMSO Dimethyl sulfoxide

    ESC Equilibrium solvent content

    EMA Ethyl methacrylate

    EDGMA Ethylene glycol dimethylacrylate [2-(2-Methyl-acryloyloxy)ethyl 2-methyl-

    acrylate]

    FDA Food and Drug Administration

    FTIR Fourier-transform infrared

    GFA Glass forming ability

    GI Gastrointestinal

    GS Glass stability

    ICH International conference on harmonization

    Tg Glass transition temperature

    HEMA Hydroxyethyl methacrylate [2-Hydroxyethyl 2-methylprop-2-enoate]

  • xxvi

    HPMC Hydroxypropyl methylcellulose

    HPMCAS Hydroxypropyl methylcellulose acetate succinate

    HPMCP Hydroxypropylmethyl cellulose phthalate

    IND Indomethacin [2-{1-[(4-chlorophenyl)carbonyl]-5-methoxy-2-methyl-1H-indol-3-

    yl}acetic acid]

    IR Infrared

    MMA Methyl methacrylate

    MW Molecular weight

    NAP Naproxen [(+)-(S)-2-(6-methoxynaphthalen-2-yl) propanoic acid]

    NMR Nuclear magnetic resonance

    ODE Ordinary differential equation

    PDE Partial differential equation

    PHEMA Poly(2-hydroxyethyl methacrylate)

    PIR Piroxicam [(8E)-8-[hydroxy-(pyridin-2-ylamino)methylidene]-9-methyl-10,10-

    dioxo-10λ6-thia-9-azabicyclo[4.4.0]deca-1,3,5-trien-7-one]

    PLS Partial least square

    PVP Polyvinylpyrrolidone

    RH Relative humidity

    SEM Scanning electron microscopy

    SI Sink index

    SNV Standard normal variate

    ss Solid state

    UV Ultraviolet

    XRD X-ray diffraction

  • 1

    Chapter 1

    Introduction

    1.1 Solubility enhancement of poorly water-soluble drugs for oral drug

    delivery

    Recent advances in combinatorial chemistry, automated synthesis and high-throughput screening

    have significantly improved effectiveness of the drug discovery process (White, 2000). Although

    many of these new chemical entities exhibit promising therapeutic potential, one major dilemma

    in developing these candidate compounds into oral dosage forms, the most popular route of

    administration, is their poor aqueous solubility and/or permeability across the intestinal villi in the

    gastrointestinal (GI) tract. According to the Biopharmaceutics Classification System (BCS), drug

    substances are classified into four categories according to their solubility and permeability

    properties, as shown in Figure 1.1. Absorption of orally administered medications involves

    solubilization of drug molecules in the GI fluid and transport across membranes of the epithelial

    cells in the GI tract. Therefore, poorly water-soluble and/or poorly permeable drugs with

    suboptimal bioavailability require enabling formulation approaches to develop oral dosage forms

    with an enhanced solubility and/or permeability to reach systemic circulation. Improving drug

    permeability often requires chemical modifications of the drug, which may fundamentally alter its

    toxicity profile and therapeutic potential. On the other hand, various techniques are available for

    the enhancement of the aqueous solubility of poorly water-soluble drugs without chemically

    altering the molecular structure of the drug. It is estimated that approximately 60-70% of newly

    discovered therapeutic compounds are classified in BCS II (high permeability, low solubility)

    during the drug discovery process (Douroumis and Fahr, 2012; Lipinski et al., 1997). Hence, a

  • 2

    significant amount of current research focuses on enhancing the aqueous solubility of these BCS

    II, small molecular weight (MW less than 1000 Da) active pharmaceutical ingredients (API).

    Figure 1.1: The Biopharmaceutics Classification System (BCS) of drugs is based on the solubility

    threshold above which the highest required dose is soluble in less than 250-mL aqueous media

    over the pH 1.0 to 7.5 and the permeability threshold above which the extent of absorption in

    humans is determined to be more than 90% of the dose administered by i.v.

    (http://www.fda.gov/AboutFDA/CentersOffices/OfficeofMedicalProductsandTobacco/CDER/uc

    m128219.htm).

    In previous attempts to address this issue, methods commonly used to increase drug solubility have

    practical limitations and may not always accomplish the desired enhancement in drug solubility

    and bioavailability (Fahr and Liu, 2007; Pouton, 2006). For example, particle size reduction (or

    increase in surface area) often has a threshold of achievable size reduction; creating stable salt

    forms or pro-drug of therapeutic agents is not always feasible; introducing surfactants or co-

    solvents may lead to liquid formulations that are usually poor in patient acceptability and

    undesirable for commercialization (Serajuddin, 1999). Among various known approaches,

    incorporating a poorly water-soluble compound in a suitable polymeric carrier to form an

    amorphous solid dispersion (ASD) has become an increasingly important strategy in the solubility

    and bioavailability enhancement for oral delivery of poorly water-soluble compounds. Various

    approaches for the preparation, characterization and stabilization of ASDs for oral drug delivery

    http://www.fda.gov/AboutFDA/CentersOffices/OfficeofMedicalProductsandTobacco/CDER/ucm128219.htmhttp://www.fda.gov/AboutFDA/CentersOffices/OfficeofMedicalProductsandTobacco/CDER/ucm128219.htm

  • 3

    have been reviewed comprehensively (Chiou and Riegelman, 1971; Craig, 2002; Kawakami, 2009;

    Leuner and Dressman, 2000; Serajuddin, 1999; Yu, 2000). Poorly water-soluble drugs in their

    stabilized amorphous form can generate a transient but highly supersaturated solution

    concentration (i.e., kinetic solubility) significantly greater than the equilibrium saturation

    concentration of their crystalline counterparts. Since drug supersaturation increases the driving

    force for oral absorption, maintaining an elevated and sustained level of drug supersaturation is

    critical to improving the bioavailability of poorly water-soluble drugs. The causality between

    increased kinetic solubility from ASDs and improved oral bioavailability has been demonstrated

    in many in vivo studies (Kohri et al., 1999; Law et al., 2004; Newa et al., 2007; Six et al., 2005;

    Verreck et al., 2004; Yüksel et al., 2003).

    1.2 Pharmaceutical significance of amorphous solid dispersions in polymeric

    carriers

    Solids in the amorphous state (also referred to as “unstable form”, “high-energy state”) are

    structurally defined as the lack of a long-range order of molecular packing or the lack of a

    crystalline state. The amorphous state has a higher internal energy and specific volume compared

    to the crystalline state (Figure 1.2). Due to “loose” molecular packing, the amorphous state has a

    high level of free energy, which leads to various physicochemical characteristics such as an

    elevated aqueous solubility, higher vapor pressure, greater molecular mobility and higher chemical

    reactivity than their crystalline counterpart (Figure 1.3). In contrast to the equilibrium solubility,

    which is an intrinsic thermodynamic property of the crystalline drug, the dissolution of amorphous

    pharmaceuticals achieves a transient supersaturation (i.e., a kinetic drug solubility that is higher

    than the equilibrium solubility), as a result of the lack of crystalline lattice energy of the solids.

    High-energy formulations based on amorphous pharmaceuticals can therefore improve oral

    bioavailability of poorly water-soluble drugs by generating supersaturated drug solutions in the GI

    tract. However, the metastable amorphous structure will be eventually converted to the equilibrium

    crystalline state, provided that there is a thermodynamic driving force sufficient to overcome the

    Gibb’s free energy barrier. The threshold drug loading level in ASD systems in different polymeric

    matrices above which an amorphous-to-crystalline transition tends to occur has typically been

    identified empirically. Stressed conditions including high relative humidity, elevated temperature

    and aging are known to accelerate the kinetics of transformation from metastable amorphous drugs

  • 4

    into the crystalline state (Konno and Taylor, 2008). Unfortunately, questions on their solid-state

    structure, mechanisms of dissolution enhancement, and criteria of solid dispersion stability upon

    storage have remained mostly unanswered (Craig, 2002). Recurrence of crystallization in many of

    these systems still represents the primary factor affecting product stability. The poor predictability

    of ASD stability is due to the lack of a more fundamental understanding of their physical properties

    and parameters which govern the stability of amorphous structures in retarding the initiation of

    nucleation and the propagation of drug crystallization. The formation of stabilized ASD in an inert

    carrier has been shown to be very effective in delaying nucleation and crystallization of amorphous

    drugs to achieve a reasonable shelf life of pharmaceutical products.

    Figure 1.2: Thermodynamic description of different solid states (crystalline, amorphous and

    supercooled liquid).

  • 5

    Figure 1.3: Gibb’s free energy levels of the amorphous state (metastable), crystalline state (stable)

    and unstable state.

    From a list of pharmaceutically acceptable excipients that are FDA-approved in human oral

    delivery, ideal carrier matrix systems have to demonstrate the ability to maintain amorphous drug

    in solid dispersions, enhancing drug dissolutions and subsequent bioavailability, and a potential of

    programmable release rates through the GI tract. Earlier investigations of incorporating poorly

    water-soluble drugs in solubilizing agents were mostly concerned with identifying rapidly

    dissolving small MW carriers (e.g., sulfathiazole-urea (Sekiguchi and Obi, 1961)) and polymeric

    carriers (e.g., sulfathiazole-polyvinylpyrrolidone PVP (Simonelli et al., 1969)) to improve the drug

    dissolution rate, the degree of supersaturation and the oral bioavailability. The solute molecules

    (amorphous drug) can either be substituted for solvent molecules or be fitted into the interstices

    between the solvent molecules in a eutectic mixture (Figure 1.4 A&B) (Leuner and Dressman,

    2000). Recent research interests have gradually been shifting to relevant physicochemical

    properties of drug-carrier composites that can enhance the stabilization of metastable amorphous

    drugs, particularly in the form of solid solutions in polymeric carriers because of their ability to

    inhibit nucleation and crystal growth in the solid state (Figure 1.4 C) (Bhugra and Pikal, 2008;

    Khougaz and Clas, 2000b; Konno and Taylor, 2007). This generally involves the preparation of

    solid solutions in polymers that have a glass transition temperature (Tg) higher than room

    temperature as the low molecular mobility of which contributes to a slow crystallization rate at

    ambient conditions. A variety of these carrier matrices can be categorized as hydrophilic,

  • 6

    hydrophobic, enteric and an association polymer system. In some cases, a ternary (polymer A-

    polymer B-drug) system might be necessary to achieve both stabilization of an amorphous solid

    state and controlled release. More recently, refinement of the solid dispersion approach has been

    pursued to include surfactants, plasticizer, alkalizer, copolymer and disintegratants to form the

    desired carrier matrices (Broman et al, 2001; Ghebremeskel et al, 2006; Tran et al, 2009).

    Figure 1.4: Schematics of (A) substitutional (B) interstitial and (C) polymeric solid solution. Figure

    adapted from Leuner and Dressman (Leuner and Dressman, 2000) (reproduced with permission

    from the European Journal of Pharmaceutics and Biopharmaceutics, Copyright Elsevier 2000).

    Entrapped drug molecules in polymeric carrier matrices can be classified as crystalline solid

    dispersion (2-phase system), amorphous solid dispersion (2-phase system) and solid solution (1-

    phase system) as illustrated in Figure 1.5. Compared to the 1-phase solid solution in which drug

    molecules are molecularly dispersed in the carrier, 2-phase solid dispersion systems contain a

    separate phase of either crystalline (long-range molecular order) or amorphous drugs (short-range

    molecular order). Pharmaceutical solid dispersion systems are rarely in a completely amorphous

    state. In most cases, pharmaceutical solids are in a state between crystalline and amorphous solids

    (i.e., a mixture or hybrid of both the 2-phase and 1-phase systems). The magnitude of crystallinity

    can be measured by x-ray diffraction (XRD) analysis, spectroscopic analysis (e.g., IR and Raman)

    and thermal analysis (e.g., differential scanning calorimetry DSC and thermally stimulated current

    TSC).

    (A) (B) (C)

  • 7

    Figure 1.5: Classification of solid dispersion/solution of drug molecules in polymeric carrier

    matrix. Figure adapted from Sun and Lee (Sun and Lee, 2014) (reproduced with permission from

    Acta Pharmaceutica Sinica B, Copyright Elsevier 2014).

    1.2.1 Water-soluble carriers

    It is a common practice to employ water-soluble polymers as carriers in conventional ASDs to

    enhance solubility and dissolution rate. The conventional design of oral dosage forms based on

    amorphous solid dispersions for poorly water-soluble drugs typically focuses on increasing the

    dissolution rates, elevating the degree of supersaturation and extending its duration following the

    dissolution of various ASD systems. Table 1.1 summarizes recent studies of water-soluble ASD

    carriers such as hydrophilic polymers and enteric polymers utilized to convert poorly water-soluble

    model drugs into amorphous formulations. In this case, hydrophilic or hydrocolloid matrices form

    solid dispersion systems from which water-soluble polymers and entrapped amorphous drugs can

    dissolve in an aqueous medium. Available references commonly show enhanced dissolution rates

    of poorly soluble drugs from hydrophilic polymers including PVP and derivatives of cellulose

    such as hydroxypropyl methyl cellulose (HPMC or hypromellose), methylcellulose (MC),

    carboxymethyl cellulose sodium, and hydroxyethyl cellulose (HEC). In addition, enteric polymers

    are preferentially soluble in a pH environment above pH 5.5 to pH 6 of the intestine relative to the

    acidic gastric fluid. For instance, rapid disintegration of enteric polymers like hydroxypropyl

  • 8

    methylcellulose phthalate (HPMCP) and hydroxypropyl methylcellulose acetate succinate

    (HPMCAS) can be achieved in an elevated pH environment. Moreover, it is worth noting that the

    frequently employed solid dispersion carrier polyethylene glycol (PEG) has a low glass transition

    temperature (Tg) and is therefore more “rubbery” at ambient temperature. This unfortunately does

    not offer much retardation in the rates of drug diffusion and crystallization. In fact, most PEG-

    based solid dispersions are truly dispersions of micro-crystalline drug particles. A typical example

    of this micro-crystalline drug dispersion is a product called Gris-PEG®, a solid crystalline

    dispersion of griseofulvin in PEG. Ideal candidates of polymeric carriers should prevent

    amorphous pharmaceutical solids from nucleating and becoming crystalline in order to achieve the

    desired stability during shelf life.

    Over the past decade, numerous in vitro and in vivo studies have displayed a variety of

    combinations of carrier polymers and solid dispersion systems of poorly water-soluble drugs to

    demonstrate the effectiveness of amorphous pharmaceuticals in solubility and bioavailability

    enhancement. Regrettably, almost none of the research results drew any proper mechanistic

    conclusions. Many of the literature reports investigated methods of preparation and evaluated

    concomitant pharmacokinetic improvement of the drug-polymer amorphous solid dispersions by

    means of in vivo studies. Well-characterized water-insoluble drugs such as indomethacin,

    nifedipine, felodipine, and itraconazole were frequently investigated for their enhanced dissolution

    rates in various polymer carriers. The common methods can generally be classified as solvent-

    based (e.g., dissolution in co-solvent followed by solvent evaporation, spray-drying,

    electrospinning), temperature-based (e.g., hot-stage melting, hot-melt extrusion), and physically

    mixing (e.g., granulation, milling) in order to achieve a homogenous binary, ternary (Janssens et

    al., 2008) or multi-component solid dispersion system (Yoo et al., 2009).

  • 9

    Figure 1.6: Dissolution performance of ASD containing a model poorly water-soluble compound

    in various polymers at 10% drug loading. Dissolution was carried out using the microcentrifuge

    tube under nonsink dissolution conditions in PBS at 37oC with 200 g/mL total drug concentration

    (dissolved plus undissolved drug). Figure adapted from Curatolo et al. (Curatolo et al., 2009)

    (reproduced with permission from Pharmaceutical Research, Copyright Springer 2009).

    Typically, the dissolution of ASDs based on water-soluble polymers under nonsink dissolution

    conditions is very rapid, resulting in an initial surge of drug concentration in the dissolution

    medium followed by a decline in drug concentration due to the nucleation and crystallization

    events triggered by the rapid buildup of drug supersaturation. Depending on the ability of the

    dissolved polymer to inhibit drug precipitation from the supersaturated state, such a decline in drug

    concentration can be retarded to different degrees. In general, the more gradual the decline in drug

    concentration, the greater its effectiveness in inhibiting drug precipitation and in maintaining drug

    supersaturation (Alonzo et al., 2011; Alonzo et al., 2010). In this regard, amphiphilic HPMCAS

    has been identified as the most effective in achieving and maintaining drug supersaturation among

    several available water-soluble polymers commonly employed in ASD-based oral drug products

    (Figure 1.6) (Curatolo et al., 2009; Friesen et al., 2008). Typical dissolution profiles of ASDs

    showing a rapid initial buildup of drug supersaturation and subsequent retardation of precipitation

    have been qualitatively characterized as the “spring and parachute” approach (Figure 1.7)

    (Brouwers et al., 2009; Guzman et al, 2007; Warren et al., 2010). This combination of a rapidly

    dissolving and supersaturating “spring” with a precipitation-retarding “parachute” has been

    pursued as an effective formulation strategy to enhance the rate and extent of oral absorption.

  • 10

    Although such “spring and parachute” dissolution data have been fitted to empirical rate equations

    to estimate the time constants for the “spring” and “parachute” portions of the dissolution profiles

    (Kawakami, 2012), the interplay between these two rate processes in achieving and maintaining

    supersaturation remains inadequately understood.

    Figure 1.7: Illustration of the time evolution of kinetic solubility profiles of a crystalline drug,

    amorphous drug and ASD in water-soluble carriers under nonsink dissolution conditions. Profile

    2 represents the dissolution of a higher energy “spring” form of the drug in absence of any

    crystallization inhibitor; Profile 3 displays the combination of the rapidly supersaturating “spring”

    form and precipitation-inhibiting “parachute” form. Cs indicates the equilibrium solubility. Figure

    adapted from Brouwers et al. (Brouwers et al., 2009) (reproduced with permission from the

    Journal of Pharmaceutical Sciences, Copyright John Wiley and Sons, 2009).

  • 11

    Model Drug Carrier* Additives (copolymer,

    surfactant, plasticizers) Preparation method**

    Ref

    Felodipine PVP None C/SE (Marsac et al., 2010)

    Tacrolimus HPMC

    PEG6000, PEG4000, Poloxamer 188, Poloxamer 407, SDS SD (Park et al., 2009)

    Indomethacin PVP None C/SE (Telang et al., 2009)

    AMG 517 (VR1 antagonist) HPMCAS, HPMC None SD

    (Kennedy et al., 2008)

    Compound C35H35N5O3 PVP poloxamer 188 C/SE, HE

    (Lakshman et al., 2008)

    Itraconazole mannitol/lecithin None FD (URF), PM (Yang et al., 2008)

    Itraconazole PEG / HPMC None SD (Janssens et al., 2008)

    UC 781 (anti-HIV) PVP/VA, HPMC

    TPGS (d-alpha-tocopheryl polyethylene glycol 1000 succinate) C/SE

    (Goddeeris et al., 2008)

    Felodipine PVP, HPMCAS, HPMC None C/SE

    (Konno et al., 2008)

    Nifedipine PVP, HPMC, PHPA None SD (Aso et al., 2007)

    Ibuprofen

    PVP (Kollidon 25, Kollidon 30, Kollidon VA64, Kollidon CL) None SD (Xu et al., 2007)

    Itraconazole HPMC, HPMCP

    Polysorbate 80, anhydrous silicic acid, croscarmellose sodium, magnesium stearate C/SE, HE

    (Oshima et al., 2007)

    Itraconazole PEG, HPMC None SD, HE (Janssens et al., 2007)

    Itraconazole HPMCP, Eudragit (L100) None FD (URF)

    (Overhoff et al., 2007)

    Piroxicam

    Polyoxyethylene 40 Stearate, Eudragit (E100) Mannitol, dextrin C/SE

    (Valizadeh et al., 2007)

    Compound (MW400)

    PVP, Plasdone (S630), HPMC

    Tween-80 and Docusate sodium HE

    (Ghebremeskel et al., 2007)

    Meloxicam PEG None C/SE, PM (Kumar and Mishra, 2006)

    KRN633 (VEGF tyrosine kinase inhibitor) PVP None C/SE

    (Matsunaga et al., 2006)

    Nifedipine / Felodipine PVP None C/SE

    (Marsac et al., 2006)

    Acetaminophen HPMCP, chitosan None SD (Chen et al., 2006)

    Bicalutamide PVP None C/SE, PM (Ren et al., 2006)

    Felodipine PVP, HPMC Poloxamer 127 C/SE, PM (Kim et al., 2006)

    Felodipine PVP, PEG sodium docusate C/SE (Karavas et al., 2005)

    Nitrendipine HPMCP None HE (Wang et al., 2005)

  • 12

    Model Drug Carrier* Additives (copolymer,

    surfactant, plasticizers) Preparation method** Ref

    Felodipine HPMC

    Poloxamer 188, Poloxamer 407, HCO-60 C/SE (Won et al., 2005)

    Nifedipine HPMCAS, HPMCP, MAEA, PVP None C/SE (Tanno et al., 2004)

    Nifedipine / Phenobarbital PVP None C/SE (Aso et al., 2004)

    Ketoprofen PEO CrosPVP, SDS PM, ME (Schachter et al., 2004)

    Ritonavir PEG None C/SE (Law et al., 2004)

    Indomethacin PVP None PM (Watanabe et al., 2003)

    Itraconazole HPMC None PM, HE (Verreck et al. 2003a,b,c)

    Indomethacin Na-indomethacin / PVP None PM, C/SE

    (Tong and Zografi, 2001)

    Probucol PVP, PAA, PEO Tween 80 C/SE, ME (Broman et al., 2001b)

    Furosemide PVP calcium alginate Floating multi-unit system

    (Iannuccelli et al., 2000)

    Itraconazole lactose Disintegrants: Primogel, Kollidon CL, and Ac-Di-Sol PM

    (Chowdary and Rao, 2000)

    *Carrier: HPMC hydroxypropyl methyl cellulose; HPMCAS hydroxypropyl methyl cellulose acetate succinate; HPMCP hydroxypropyl methyl

    cellulose phthalate; MAEA methacrylic acid ethyl acrylate; PAA poly(acrylic acid); PEG polyethylene glycol; PEO polyethylene oxide; PHPA

    poly(N-5-hydroxypentyl) aspartamide; PVP polyvinyl pyrrolidone; PVP/VA polyvinyl pyrrolidone/vinyl acetate; SDS sodium dodecyl sulfate.

    **Preparation method: C/SE cosolvent/solvent evaporation; SD spray-drying; FD freeze-drying; URF ultra-rapid freezing; ME melting; HE hot-

    melt extrusion; PM physical mixing

    Table 1.1: Selected studies of water-soluble carriers for amorphous drugs.

  • 13

    Category Examples Carrier* Model Drug Preparation

    method** Release mode***

    Ref

    a. Crosslinked hydrogel

    cr-PHEMA hydrogel

    diclofenac sodium, naproxen, piroxicam, indomethacin

    S/SE IR/CR (Sun et al., 2012; Zahedi and Lee, 2007)

    cr-PEO hydrogel

    progesterone S/SE CR (Carelli et al., 1993)

    Carbopol® phenacetin S/SE CR (Ozeki et al, 2000)

    I. Non-porous

    b. Water-insoluble polymer

    Ethylcellulose indomethacin C/SE CR (Ohara et al, 2005)

    Eudragit® RS, RL

    indomethacin, dipyridamole

    C/SE CR (Beten et al, 1994; Oth et al, 1989)

    c. Lipid

    Labrasol and Gelucire 44/14

    piroxicam HM IR (Yüksel et al, 2003)

    Gelucire 44/14 and Gelucire 50/13

    gilbenclamide C/SE IR (Chauhan et al, 2005)

    “popcorn” cr-PVP

    griseofulvin, indomethacin I/SE, HG IR

    (Fujii et al, 2005; Shibata et al, 2007; Carli et al, 1986)

    II. Porous

    silica

    fenofibrate, carbamazepine, cinnarizine, danazol, ibuprofen diazepam, griseofulvin, indomethacin, ketoconazole, nifedipine, phenylbutazone

    I/SE, SD IR

    (Van Speybroeck et al, 2009; Van Speybroeck et al, 2010; Shen et al, 2010)

    starch foam lovastatin I/SE IR (Wu et al, 2011)

    carbon celecoxib I/SE IR (Zhao et al, 2012)

    *Carrier: cr-PHEMA poly(2-hydroxyethyl methacrylate) crosslinked with ethylene glycol dimethacrylate; Carbopol® (910, 971P, 934P, 974P,

    940) polyacrylic acid (lightly crosslinked with allyl sucrose or allyl pentacrythritol); cr-PEO poly(ethylene oxide) crosslinked with hexamethylene

    diisocyanate (HMDIC) or tolylene-2,4-diisocyanate (TDIC); cr-PVP crosslinked polyvinylpyrrolidone; Gelucire 44/14, 50/13 polyethylene glycol

    glycerides; Labrasol caprylocaproyl macrogolglycerides.

    **Preparation method: S/SE swelling/solvent evaporation; C/SE cosolvent/solvent evaporation; I/SE immersion/solvent evaporation, HM hot melt;

    HG heated granulation (

  • 14

    1.2.2 Water-insoluble carriers

    In contrast to studies of water-soluble ASD carriers, previous efforts in applying water-insoluble

    polymers as ASD carriers have primarily aimed to achieved controlled release (Tran et al. 2011;

    Zhu et al. 2006). Table 1.2 summarizes available water-insoluble carriers utilized to convert poorly

    water-soluble model drugs into amorphous formulations categorized by their physicochemical

    characteristics such as porosity, location of amorphous drug and carrier chemical composition. In

    the case of nonporous ASD carriers where amorphous drug molecules are completely dissolved

    (i.e., one-phase solid solution), carriers with extremely low drug diffusivity such as “glassy

    polymers” (i.e., Tg much higher than ambient temperature) will exhibit better ASD stability due

    to hindered drug diffusion and inhibition of drug precipitation in the glassy matrix. The presence

    of specific drug-carrier intermolecular interactions due to hydrogen bonding, dipole-dipole

    attraction and van der Waals forces further stabilizes the entrapped amorphous drug, preventing it

    from nucleating and becoming crystallite. By contrast, in porous carriers where the incorporated

    amorphous drug is localized in interstitial pore space (i.e., not molecularly dispersed), the

    nucleation and crystallization of this aggregated amorphous drug can give rise to stability issues.

    However, the drug nucleation and crystallization rates can be reduced if the size of the pore is

    sufficiently small compared to that of the critical nuclei, rendering it energetically unfavorable for

    nuclei to grow (Rengarajan et al., 2008; Van Speybroeck et al., 2009; Van Speybroeck et al., 2010).

    For inorganic mesoporous materials such as silica and silicon (Xu et al., 2012), their large surface

    area and pore volume can accommodate a large amount of drug payload. Drug nucleation and

    crystal growth in the porous channels will be energetically unfavorable if the size of the critical

    nucleus is larger than that of the pore (Jackson and McKenna, 1996; Wang et al., 2006). In other

    words, the spatial constraint of a capillary imposed on the amorphous drug below the critical

    nucleus size has a stabilizing effect. In addition, a strong interaction with the pore walls (e.g.,

    through hydrogen bonding) further stabilizes the confined drug molecules in the amorphous state

    and this typically occurs in nano-sized pores with a pore diameter smaller than approximately 10

    nm (Rengarajan et al., 2008). The drug loading process in such mesoporous materials commonly

    involves immersing the carriers in a concentrated drug solution to fill the pores followed by the

    evaporative removal of solvents. One study claims to have successfully accomplished solid

    dispersion by means of a melt-mixing method, during which a physical mixture of nitrendipine

    and mesoporous silica particles is heated above the melting point of the drug (Wang et al., 2006).

  • 15

    Nonetheless, in this case the high viscosity of the melted drug can interfere with the capillary

    action within the pores, potentially causing an incomplete drug loading. Similar constraints also

    apply to macroporous polymeric carriers such as crosslinked polyvinylpyrrolidone (crospovidone)

    which exhibits a “popcorn” structure, containing many macroscopic cavities, as clearly shown

    under the SEM (Carli et al., 1986; Fujii et al., 2005; Shibata et al., 2007). In this case, in addition

    to the conventional loading method of sorption from a concentrated drug solution, an alternative

    method of drug loading based on blending or milling of the drug and porous crospovidone particles

    without solvent under high mechanical shear for an extended period of time has been proposed

    (Fujii et al., 2005; Shibata et al., 2007). However, the “amorphous” material produced in this could

    be a result of the well-known shear-induced phase transformation at the shear-fractured surfaces

    on drug crystals (Greco and Bogner, 2010; Koike et al., 1990). As such, the long-term stability of

    such mechanically generated amorphous systems is questionable since without any mechanism for

    crystallization inhibition this exposed surface amorphous content can easily be converted to the

    crystalline phase under accelerated temperature and/or in higher humidity.

    Alternatively, in a nonporous insoluble carrier system, amorphous drugs can exist as surface

    adsorbed or molecularly dissolved/dispersed in the matrix depending on the drug loading process.

    For example, solvent or melt granulation of a crystalline drug with a nonporous insoluble carrier

    typically results in an amorphous drug either adsorbed on the carrier surface or unevenly

    distributed throughout the carrier-drug granule (i.e., a two-phase amorphous solid dispersion).

    Such pure amorphous drug, either surface adsorbed or unevenly distributed lacks the benefit of

    crystallization inhibition conferred by dissolving or molecularly dispersing the drug in a protective

    carrier matrix and, therefore, is prone to undesirable nucleation and crystallization in the solid

    state. Nonetheless, the creation of uniformly dissolved or molecularly dispersed drug in ASD

    polymeric carriers (i.e., one-phase solid solution) through either drug sorption (e.g., from a good

    swelling solvent) or co-precipitating from a common organic solvent (e.g., via spray drying or

    freeze drying) is understandably more advantageous than producing dispersed two-phase ASD

    systems in terms of its effectiveness in crystallization inhibition and its effect in enhancing the

    physical stability. Therefore, applicable nonporous water-insoluble carriers include crosslinked

    polymers (e.g., PHEMA), cellulose derivatives (e.g., ethylcellulose), and lipids (e.g., PEG-

    glycerides). The common methods of preparation of ASD based on water-insoluble carriers can

    be generally classified as supercooling of the melt (e.g., hot-stage melting, hot-melt extrusion), co-

  • 16

    precipitating from a common organic solvent (e.g., spray-drying, freeze-drying), and equilibrium

    sorption (e.g., from a concentrated drug solution prepared in a good swelling solvent) to produce

    a homogenous binary one-phase system. In this case, the carrier plays an essential role in

    preserving the entrapped drug molecules in the amorphous state despite the fact that the exact

    mechanisms of crystallization inhibition in solid molecular dispersions are still not completely

    understood. Various factors such as molecular mobility, thermodynamic properties, and drug-

    polymer interactions have been identified as responsible for inhibiting drug crystallization from

    the amorphous state (Bhugra and Pikal, 2008; Hancock and Zografi, 1997; Khougaz and Clas,

    2000). Stressed conditions involving high relative humidity and high temperature in combination

    with ageing may accelerate the transformation of metastable amorphous drugs in an ASD into a

    more thermodynamically stable crystalline state (Hancock and Zografi, 1997; Konno and Taylor,

    2008).

    1.3 Crosslinked PHEMA hydrogels for amorphous solid dispersions carriers

    Poly (2-hydroxyethyl methacrylate) (PHEMA) has been widely utilized in controlled drug delivery

    systems particularly for water-soluble drugs and in biomedical application such as contact lenses,

    wound dressing and tissue engineering. Crosslinked PHEMA is probably the most extensively

    studied gel-forming water-insoluble polymer among several other synthetic hydrogels based on

    nonionic hydrophilic monomers (e.g., hydroxyalkyl acrylate, N-substituted methacrylamides and

    N-vinyl-2-pyrrolidone) commonly used in swelling-controlled oral drug delivery (Gehrke, 2000).

    The drug release kinetics and release rate modulation from such a spherical monolithic diffusion-

    controlled system have previously been characterized in detail (Lee and Kim, 1991; Lee and

    Peppas, 1987). Chemically, PHEMA has hydrophilic side groups which can promote the sorption

    of a significant amount of polar solvent, thereby providing desired conditions to create solid

    dispersions with high drug loading. PHEMA can also be easily copolymerized with other

    hydrophobic monomers, such as methyl methacrylate (MMA) or its homologs, to improve the

    compatibility of the hydrogel matrix with hydrophobic drugs. Furthermore, the crosslinker chain

    length and crosslinking density can be adjusted to affect the network mesh size which provides

    further control of the diffusional drug release (Hoare and Kohane, 2008). Thus, PHEMA is a good

    candidate drug carrier because of the ease in regulating drug release by controlling particle size,

  • 17

    crosslinking density, chemical composition (e.g., copolymers) and water sorption rate to achieve

    different drug loading levels and release profiles suitable for oral drug delivery.

    Orally administered excipients must be stable, non-toxic, biocompatible in the GI tract and

    compatible to API and other excipients in the formulation in order to satisfy regulatory

    requirements. Full in vivo studies such as biodistribution, immune response, clearance, and chronic

    toxicology studies for different durations are generally required to demonstrate the excipient’s

    safety for use in human drug products. The approval process to introduce a new pharmaceutical

    excipient onto the market requires expensive toxicological investigations and typically takes a

    similar period of time as that of a new API (Katdare and Chaubal, 2006). Due to the complexity

    and high cost of long-term toxicological and clinical testing, PHEMA has not yet been used as an

    oral excipient in the FDA approved drug products (note: PHEMA is a FDA-approved inactive

    ingredient for only topical and dressing application as on September 16, 2013). Nevertheless,

    PHEMA hydrogel has a long history of human use as biomedical implants and for controlled drug

    release from medical devices. A number of in vivo tests using PHEMA hydrogel delivery systems

    have produced favorable results in blood and tissue compatibility (Imai and Masuhara, 1982).

    Moreover, no toxic abnormalities were reported when crosslinked PHEMA hydrogels loaded with

    diclofenac sodium and theophylline were administered orally and rectally, respectively, to human

    subjects in clinical pharmacokinetic studies (De Leede et al., 1986; Thakker et al., 1992). Since

    crosslinked hydrogels for the purpose of oral delivery are generally of micron sizes and insoluble

    in the GI fluid, cellular internalization or absorption through the mucosal membrane of the GI tract

    is very unlikely. Therefore, the above mentioned lines of evidence suggest that crosslinked

    PHEMA hydrogel beads would be an attractive candidate excipient for oral drug delivery. To

    promote the practical application of PHEMA hydrogels as an excipient for oral dosage forms,

    long-term oral toxicity studies should be further investigated and established.

    In selecting an appropriate polymer matrix for an ASD system, the carrier polymer must provide

    sufficient stabilization effect in the solid state to prevent the entrapped amorphous drug from

    crystallizing, thereby achieving a reasonable shelf life. From a mechanistic point of view,

    nucleation and subsequent crystallization of an amorphous drug in a polymer matrix deplete the

    local drug concentration around the crystallite and thus triggering drug diffusion from the

    surrounding amorphous drug region to the crystallite surface to sustain crystal growth. The reduced

    segmental mobility of polymer chains in glassy PHEMA hydrogels at ambient temperature well

  • 18

    below its Tg of 115 oC (Roorda et al., 1988) should lengthen the polymer molecular relaxation

    time and retard the diffusional mobility as well as reduce the nucleation and crystallization rates

    of the entrapped molecularly dispersed drug, thus resulting in enhanced stability of such hydrogel-

    based solid dispersion systems. An additional advantage is that PHEMA hydrogels exist as

    optically clear glassy solids at ambient temperature, which is suitable for direct microscopic

    examination for any phase transformation and for the determination of the drug loading threshold

    in the hydrogel matrix above in which amorphous-to-crystalline transition may occur.

    In addition to the physical stability of ASDs, the extent of solubility and dissolution enhancement

    is another key criterion to consider in selecting an appropriate ASD carrier. For a long while, the

    kinetics of drug release from crosslinked water-insoluble hydrogels was characterized only in the

    context of achieving controlled drug release under perfect sink dissolution conditions (Lee, 1985).

    The swelling kinetics and diffusional drug release kinetics are important physical characteristics

    of hydrogel-based drug delivery. The swelling kinetics of crosslinked hydrogels is important not

    only to reach an adequate drug loading but also to regulate the rate of drug release. During drug

    release, the time evolution of the solvent-penetrating front and normalized dimensional changes

    of drug-loaded PHEMA hydrogel beads is a function of crosslinking density, drug solubility and

    initial drug loading. For example, two moving boundaries (swelling front and drug dissolution

    front) were generally observed for poorly water-soluble diclofenac sodium in PHEMA hydrogels

    at a high drug loading range of 25-30% w/w as the aqueous medium penetrates into the matrix,

    whereas a single fast-moving swelling front was observed below this range of drug loading (Lee

    and Kim, 1991). In terms of dimensional transformation, the swelling PHEMA beads generally

    first expand to a maximum in radius followed by a gradual decrease to an equilibrium dimension

    during the drug release (Lee, 1983; Lee and Kim, 1991). Moreover, the dissolution of incorporated

    drug from crosslinked PHEMA hydrogels is regulated by diffusion where the hydrogel matrix is

    insoluble and the dissolved or dispersed drug slowly diffuses out of the hydrogel network.

    Although the application of PHEMA hydrogel as a carrier for ASDs of model poorly water-soluble

    drugs such as diclofenac sodium, naproxen and piroxicam have been proposed recently (Zahedi

    and Lee, 2007), the effect of the hydrogel matrix on the maintenance of supersaturation was

    virtually unknown. Crosslinked hydrogels with a unique diffusion-controlled mechanism of drug

    release will hopefully provide an adequate alternative to the rapidly dissolving ASD carriers in

    enhancing the solubility and dissolution rate of poorly water-soluble drugs.

  • 19

    1.4 Crystallization of amorphous pharmaceuticals in the solid state

    1.4.1 Solubility advantage of amorphous solids

    Several theoretical models present amorphous solids as having “excess” (Yu, 2000)

    thermodynamic properties including higher volume, enthalpy, entropy and free energy, as

    compared to its crystalline state. The change in free energy levels between amorphous and

    crystalline states accounts for their significant differences in physiochemical properties such as the

    apparent solubility (or the kinetic solubility) in water. Parks et al (Parks et al, 1934) for the first

    time theoretically estimated the difference in kinetic solubility between amorphous and crystalline

    glucose based on thermodynamic analysis, and experimentally confirmed their theoretical

    predictions. The free energy difference between the amorphous and crystalline phases is expressed

    by:

    ∆𝐺𝑎,𝑐 = −𝑅𝑇𝑙𝑛 (𝑎𝑎

    𝑎𝑐) = −𝑅𝑇𝑙𝑛 (

    𝛾𝑎𝐶𝑆𝑎

    𝛾𝑐𝐶𝑆𝑐 ) Equation 1.1

    where a is activity of the solute in the saturated solution (superscript: a for amorphous and c for

    crystalline), R the gas constant, T the temperature, 𝛾 the activity coefficient (close to unity when

    the system is in a diluted form) and Cs the concentration at saturation (i.e., solubility). One

    important assumption of this equation is that an equilibrium solubility exits in both crystalline and

    amorphous states. For crystalline materials, the equilibrium solubility refers to the physical state

    when the molecules in the solid state are in chemical equilibrium with those in the solution state

    in a close system. In other words, the rates of dissolution and precipitation of individual molecules

    between the solid and solution phases are equal to each other. However, amorphous solids are in

    a non-equilibrium state (i.e., metastable) at which the disordered molecular structure does not

    require the breaking of crystal lattice upon dissolution. The apparent solubility of amorphous solids

    generated by time-dependent rates of dissolution and precipitation is in fact a kinetic property.

    Nevertheless, Parks et al (Parks et al, 1934) treated amorphous drug as either an equilibrium

    supercooled liquid or a pseudo-equilibrium glass and estimated the ratio of solubility enhancement

    between the amorphous and crystalline states by calculating the difference in Gibbs free energy

    expressed in enthalpy (H) and entropy (S):

  • 20

    ∆𝐺𝑎,𝑐 = ∆𝐻𝑎,𝑐 − 𝑇∆𝑆𝑎,𝑐 Equation 1.2

    According to Kirchhoff’s law, ∆𝐻𝑎,𝑐and ∆𝑆𝑎,𝑐 can be calculated by:

    ∆𝐻𝑎,𝑐 = ∆𝐻𝑓𝑐 + ∫ ∆𝐶𝑝

    𝑇

    𝑇𝑚𝑑𝑇 = ∆𝐻𝑓

    𝑐 − (𝐶𝑝𝑎 − 𝐶𝑝

    𝑐) × (𝑇𝑚 − 𝑇) Equation 1.3

    ∆𝑆𝑎,𝑐 = ∆𝑆𝑓𝑐 + ∫

    ∆𝐶𝑝

    𝑇

    𝑇

    𝑇𝑚𝑑𝑇 = ∆𝑆𝑓

    𝑐 − (𝐶𝑝𝑎 − 𝐶𝑝

    𝑐) × 𝑙𝑛 (𝑇𝑚

    𝑇) Equation 1.4

    where ∆𝐻𝑓𝑐 is the entropy of fusion, Cp the isobaric heat capacity, Tm melting temperature and

    ∆𝑆𝑓𝑐 entropy of fusion (calculated by ∆𝐻𝑓

    𝑐/𝑇𝑚). The difference in isobaric heat capacity (∆𝐶𝑝)

    has been estimated to be (1) zero (i.e., ∆𝐶𝑝 ≈ 0), (2) the heat capacity gained upon going through

    the glass transition (i.e., ∆𝐶𝑝 ≈ ∆𝐶𝑝,𝑇𝑔), (3) the entropy of fusion (i.e., ∆𝐶𝑝 ≈ ∆𝑆𝑓), and (4) a non-

    zero constant in which the difference in enthalpy between the two states vanishes at a temperature

    T∞, slightly below Tg (i.e., ∆𝐶𝑝 ≈∆𝐻𝑓

    𝑇𝑚−𝑇∞). The last approach can be further utilized to derive a

    simple mathematical express of the Gibb’s free energy difference, also known as the Hoffman

    equation:

    ∆𝐺𝑎,𝑐 =∆𝐻𝑓(𝑇𝑚−𝑇) 𝑇

    𝑇𝑚2 Equation 1.5

    This theoretical consideration has been the basis for many attempts on the thermodynamic

    prediction of solubility enhancement ratios between the amorphous and crystalline forms of poorly

    water-soluble drugs (Hancock and Parks, 2000; Alonzo et al, 2010; Alonzo et al, 2011). Additional

    rigorous modifications including the consideration of configurational heat capacity, water sorption

    during dissolution and different degrees of drug ionization in amorphous and crystalline forms

    have provided further improvements in producing a closer agreement with the measured data

    (Aceves‐Hernandez et al., 2009; Hoffman, 1958; Matteucci et al., 2008; Murdande et al., 2010a,

    b, 2011a, b; Ogino et al., 1990). Nevertheless, both the theoretical prediction of the solubility

    enhancement using thermodynamic analysis and the accurate experimental measurements of the

    solubility of the amorphous state has proven to be very challenging.

  • 21

    1.4.2 Classical nucleation theory (solid state)

    Due to the non-equilibrium nature of the amorphous solids, the amorphous state is less

    thermodynamically stable than any crystalline form, leading to an inevitable tendency for the

    amorphous materials to transform to a crystalline phase. The higher free energy level of the

    amorphous form relative to the crystalline state provides the thermodynamic driving force for

    nucleation and crystallization, causing physical instability (i.e., recrystallization) in the dosage

    forms. Nonetheless, acceptable drug stability can still be achieved in amorphous formulations for

    oral dosage forms during pharmaceutical development if the kinetics of amorphous-to-crystalline

    phase transformation can be delayed to an adequate extent. The physical process of

    recrystallization from a high-energy amorphous form in the solid state is a complex phenomenon

    from a mechanistic viewpoint based on both nucleation and crystal growth. The Classical

    Nucleation Theory (CNT) first describes the kinetics of the homogenous nucleation process by

    considering the difference in the free energies of the crystalline and liquid phases (note: amorphous

    materials which lack crystalline lattice energy have a molecular structure similar to that of a liquid).

    The overall difference of Gibb’s free energy between the crystalline and amorphous phases is equal

    to the sum of two surface excess energies, Gs (the difference between the surface and the bulk of

    the crystalline phase) and GV (the difference between a very large crystalline particle (r = ∞) and

    the amorphous phase). Considering the growth of a spherical crystalline particle, the overall excess

    free energy is given:

    ∆𝐺 = ∆𝐺𝑆 + ∆𝐺𝑉 = 4𝜋𝑟2𝛾 +

    4

    3𝑟3∆𝐺𝜈 Equation 1.6

    where r is the radius of the crystalline particle, γ the interfacial tension and ∆𝐺𝜈 the free energy

    change of the transformation per unit volume. The overall ∆𝐺 as a sum of positive Gs and

    negative GV has a local maximum corresponding to the critical nucleus, rc, as illustrated in Figure

    1.8. This local maximum can be obtained by taking the first-derivative of Equation 1.6:

    𝑑∆𝐺

    𝑑𝑟= 8𝜋𝑟𝛾 + 4𝜋𝑟2∆𝐺𝜈 = 0 Equation 1.7

    Therefore, the critical radius of a nucleus, rc, and the free energy difference for such a nucleus,

    ∆𝐺𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙, are given by:

  • 22

    𝑟𝑐 =−2𝛾

    ∆𝐺𝜈 Equation 1.8

    ∆𝐺𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 =16𝜋𝛾3

    3(∆𝐺𝜈)2=

    4𝜋𝛾𝑟𝑐2

    3 Equation 1.9

    Figure 1.8: Free energy diagram for nucleation process. Figure adapted from Crystallization 3rd

    edition (Mullin, 2001) (reproduced with permission from the Crystallization 3rd edition, Copyright

    Butterworth-Heinemann, 1993).

    The nucleation rate J (i.e., the number of nuclei formed per unit volume per unit time) is expressed

    in the form of an Arrhenius equation, commonly used for the rate process of a thermally activated

    reaction:

    𝐽 = 𝐴𝑒(−∆𝐺

    𝑘𝑇) Equation 1.10

    where A is the pre-exponential frequency factor, k the Boltzmann constant and R the gas constant.

    For homogenous nucleation from the melt (i.e., liquid phase), the free energy change is estimated

    by:

    ∆𝐺𝜈 = Δ𝐻𝑓(𝑇𝑚−𝑇)

    𝑇𝑚 Equation 1.11

  • 23

    where Δ𝐻𝑓 is the latent heat of fusion and Tm the melting temperature (solid-liquid equilibrium

    temperature). Therefore, the critical radius of a nucleus and the critical free energy difference are

    given as:

    𝑟𝑐 =−2𝛾𝑇𝑚

    Δ𝐻𝑓(𝑇𝑚−𝑇) Equation 1.12

    ∆𝐺𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 =16𝜋𝛾3

    3(∆𝐺𝜈)2=

    16𝜋𝛾3𝑇𝑚2

    3Δ𝐻𝑓2(𝑇𝑚−𝑇)2

    Equation 1.13

    and the nucleation rate for the amorphous-to-crystalline transformation in the solid state may be

    expressed by substituting Equation 1.13 into Equation 1.10:

    𝐽 = 𝐴𝑒(

    −16𝜋𝛾3

    3𝑘𝑇Δ𝐻𝑓2(𝑇/𝑇𝑚)(1−𝑇/𝑇𝑚)

    2 )

    Equation 1.14

    The analytical solution describing the nucleation kinetics in the amorphous materials is a function

    of heat of fusion, melting temperature and interfacial tension. It is worth noting that the

    determination of specific interfacial tension between the developing crystalline surface and the

    amorphous bulk is not easily measured experimentally.

    1.4.3 Crystal growth (solid state)

    After the barrier to nucleation has been overcome, a number of newly formed stable nuclei will

    continue to grow into crystals of microscopic size. The process of crystal growth is generally

    described by models of (1) normal or continuous growth, (2) two-dimensional growth and (3)

    growth mediated by screw dislocation (Gutzow and Schmelzer, 2013). The crystal growth rate can

    be described by the following equation in a general form:

    𝑈 =𝐶𝑇𝜔

    𝜂(1 − 𝑒(

    −Δ𝐺𝑉𝑘𝑇

    )) Equation 1.15

    where C is a constant, ω the constant which depends on the mechanism of growth, η the viscosity

    of the system and Δ𝐺𝑉 the free energy difference between the amorphous and crystalline phases.

    The combination of equations calculating the nucleation rate (Equation 1.14) and the crystal

  • 24

    growth rate (Equation 1.15) can assess the overall crystallization kinetics of amorphous-to-

    crystalline phase transformation. However, it is worth highlighting that many of the parameters

    listed in the above mentioned equations are not easily accessible experimentally.

    1.4.4 Kolmogorov-Johnson-Mehl-Avrami (KJMA) Theory

    The kinetics of overall crystallization taking into account nucleation, crystallization and the lag

    time in both of these events can be described by the Kolmogorov-Johnson-Mehl-Avrami (KJMA)

    theory which reduces the complicated mechanism into a simple mathematical correlation

    describing the process of overall crystallization as a function of time (Avrami, 1939; Avrami 1940),

    which is convenient to use for the physical stability study of amorphous pharmaceuticals (Andronis

    and Zografi, 2000; Miyazaki 2004, Bhugar and Pikal, 2000; Gutzow Schmelzer 2013). The

    empirical equation is presented by an exponential term:

    𝑥(𝑡) = 1 − exp [−𝑘𝑡𝑛] Equation 1.16

    where x is the percentage of the crystalline phase, t the time, and k and n Avrami crystallization

    constants. One limitation of the KJMA theory is that time-dependent crystallization kinetics is

    only an approximation in which the constants of k or n may not have any physical significance.

    Despite the fact that the KJMA theory provides limited mechanistic information on crystallization

    kinetics, comparing the crystallization rate to the Avrami equation is very convenient. Therefore,

    the use of the KJMA theory offers an easy tool of direct comparison of crystallization rates between

    two systems.

  • 25

    1.5 Crystallization of supersaturated drug solutions

    1.5.1 Classical nucleation theory (solution state)

    Amorphous pharmaceuticals can generate a supersaturated concentration in aqueous solution

    significantly higher than the equilibrium saturation concentration of their crystalline counterparts.

    In a solute-solvent binary system (e.g., drug-water), solute concentration above the equilibrium

    solubility (i.e., supersaturation) provides the thermodynamic driving force for nucleation and

    crystallization. In a supersaturated drug solution, the free energy is higher in the solution phase

    than any newly formed crystalline phase. Classical nucleation theory based on the free energy

    difference can also be applied on the precipitation process of solute from a supersaturated solution.

    Equations 1.6 to 1.10 describe the nucleation rate for homogenous nucleation of the solute from a

    supersaturated solution. The free energy change ∆𝐺𝜈 in Equation 1.10 becomes

    ∆𝐺𝜈 =−2𝛾

    𝑟𝑐=

    −kTln(S)

    𝜐 Equation 1.17

    where υ is the molecular volume and s the supersaturation. This equation is derived from the basic

    Gibbs-Thomson relationship for a non-electrolyte:

    ln(𝑠) =2𝛾𝜐

    𝑘𝑇𝑟 Equation 1.18

    The substitution of Equation 1.17 into Equations 1.9 and 1.10 gives:

    ∆𝐺𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 =16𝜋𝛾3

    3(∆𝐺𝜈)2=

    16𝜋𝛾3𝜈2

    3(𝑘𝑇𝑙𝑛𝑆)2 Equation 1.19

    𝐽 = 𝐴𝑒(

    −16𝜋𝛾3𝜈2

    3𝑘3𝑇3(𝑙𝑛𝑆)2 )

    Equation 1.20

    where described here is the interfacial tension between the evolving crystalline surface and

    aqueous solution (note: different from described in Equations 1.12 to 1.14). Equations 1.19 and

    1.20 are for nucleation from a supersaturated solution which significantly depends on the degree

    of supersaturation are analogous to Equations 1.13 and 1.14 for nucleation from the melt.

  • 26

    1.5.2 Crystal growth (solution state)

    Once stable nuclei have formed (i.e., r > rc) in a supersaturated solution, these particles start to

    grow into crystals of visible size. Crystallization in a supersaturated solution is the reverse process

    of dissolution, and both phenomena are governed by diffusion in which the concentration gradient

    between the solid surface and the bulk of solution provide the driving force. During the

    precipitation process from a supersaturated solution, crystal growth requires both long-range

    transport of solute to the growing particles (i.e., diffusion) and local atomic rearrangement as the

    solutes near the particle surface (i.e., interface) are integrated into the crystal lattice. Depending

    on the rate-limiting step, crystal growth kinetics can be either diffusion-controlled or interface-

    controlled. In a diffusion-controlled crystallization process, the growth rate of a crystal can be

    predicted by the molecular flux which is related to the concentration gradient between the bulk