Transcript
Page 1: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

IMPROVE THE MANUFACTURABILITY AND DISSOLUTION

PERFORMANCE OF DRUGS THROUGH SPHERICAL

CRYSTALLIZATION

A DISSERTATION

SUBMITTED TO THE FACULTY OF

UNIVERSITY OF MINNESOTA

BY

Hongbo Chen

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Changquan Calvin Sun, Advisor

July 2020

Page 2: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

© Hongbo Chen, July 2020

Page 3: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

i

Acknowledgements

I would like to thank my adviser, Dr. Changquan Calvin Sun, for his unlimited support,

careful guidance, unceasing encouragement, and lasting friendship during my study in the

Department of Pharmaceutics at UMN in the past five years. His passion in research

inspires me and sets a role model for me to learn and improve myself. I’m grateful to have

him to be my advisor.

I want to express my thanks to my committee members, Dr. Raj Suryanarayanan, Dr.

Timothy S. Wiedmann, and Dr. Aktham Aburub, for giving me constructive suggestions

on my research and reviewing my thesis critically. I want to thank Dr. Wiedmann for

reviewing my manuscript drafts critically and providing me with valuable comments. I am

grateful to Dr. Sury for granting me access to instruments in his lab to conduct part of my

research. I am especially thankful to Dr. Aburub for his constructive comments,

suggestions, guidance, and help on my research projects.

My thanks also go to Sun lab members, especially Shubhajit, Chenguang and Manish for

their help, suggestion and discussion on my research, Shao-yu, Wei-jhe, Jiangnan, Shenye,

Kunlin, Yiwang, Amy, Gerrit, Joan, Sibo, Zhongyang, Ling, Zhengxuan, Cosima, Xin,

Shuyu, Hiro, Mingjun, Shan, Hongliang, Yuebin, and Jun for their accompany. My

gratitude also extends to Katie, Amanda, Davin, Kelsey, Krutika, Navpreet, Jayesh, and

Yafan in our department. I also want to thank my collaborators, Dr. Mahanthappa, Dr.

Haynes, Hongyun, Hyunho, and Bo for their contributions on my research projects.

Page 4: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

ii

I am grateful to have my girlfriend Jiali, for her company, support, and love during my

study at UMN in the past five years. Because of you, my life is fruitful. Finally, I would

like to thank my family. Thanks for your financial and spiritual support throughout my life

and encouragement on my choice of career.

Page 5: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

iii

Dedicated to my family

Page 6: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

iv

Abstract

Spherical crystallization, a particle engineering technique, has been commonly

used to modify the size and shape of particles to enhance the micromeritics properties of a

powder and facilitate tablet formulation development. By identifying the key factors

affecting the generation of spherical agglomerates and the underlying mechanism,

spherical agglomerates with optimal size, shape and density can be obtained, presenting

excellent flowability, tabletability and dissolution. The goals of this thesis work include:

(i) prepare and evaluate the properties of the spherical agglomerates of drugs using the two

main methods including spherical agglomeration (SA) and quasi-emulsion solvent

diffusion (QESD) , (ii) explore high drug loading direct compression tablet formulations

enabled by the spherical crystallization technique, (iii) develop a spherical cocrystallization

technique to simultaneously improve the manufacturability and dissolution of drugs, and

(iv) reduce the punch sticking propensity via a polymer assisted QESD method.

To successfully prepare the spherical agglomerates, it is crucial to select a solvent

system-based phase diagram. Spherical agglomerates of a model API, ferulic acid (FA),

via a QESD process presented exceptional flowability and tabletability. Such spherical FA

crystals enabled the development of a direct compression tablet formulation containing 99%

of the API. Spherical griseofulvin agglomerates prepared by a SA process, exhibited not

only good flowability but also excellent tabletability. The unexpected profoundly

improved tabletability was attributed to a micro porous structure of primary crystals, due

to an in-situ solvation and desolvation during the SA process, which enhanced the plasticity

of the agglomerates. For low water-soluble drugs, spherical crystallization of more soluble

cocrystals, i.e., spherical cocrystallization, is an enabling technology to simultaneously

Page 7: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

v

improve the manufacturability and dissolution and enable the development of high drug

loading tablet formulation. Moreover, spherical agglomerates of celecoxib obtained via a

polymer assisted QESD process not only showed substantially improved tabletability and

flowability but also significantly reduced punch-sticking propensity. A thin layer of

polymer coating onto the agglomerates surface, confirmed by SEM and XPS analysis,

explains the reduced punch sticking propensity due to the physical barrier of polymer

between the drug and punch.

Page 8: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

vi

Table of Contents

Acknowledgements .............................................................................................................. i

Dedication .......................................................................................................................... iii

Abstract .............................................................................................................................. iv

Table of Contents ............................................................................................................... vi

List of Tables ................................................................................................................... xiv

List of Figures .................................................................................................................. xvi

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

1.1 Background ....................................................................................................................2

1.2 Literature review ............................................................................................................3

1.2.1 Spherical agglomeration (SA) ......................................................................4

1.2.1.1 Selection of a bridging liquid ...............................................................6

1.2.1.2 Phase diagram .......................................................................................7

1.2.2 Quasi-emulsion solvent diffusion (QESD) ................................................10

1.2.3 Powder property .........................................................................................11

1.2.3.1 Micromeritic property ........................................................................11

1.2.3.2 Flowability ..........................................................................................12

1.2.3.3 Tabletability ........................................................................................13

1.2.3.4 Punch sticking ....................................................................................14

1.2.3.5 Dissolution ..........................................................................................15

1.2.4 Continuous manufacturing .........................................................................16

Page 9: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

vii

1.2.5 Spherical cocrystallization .........................................................................17

1.3 Objective and hypothesis .............................................................................................18

1.4 Reference .....................................................................................................................24

Chapter 2: Direct compression tablet containing 99% active ingredient – a tale of

spherical crystallization ...................................................................................................31

2.1 Synopsis ......................................................................................................................32

2.2 Introduction .................................................................................................................32

2.3 Methods........................................................................................................................34

2.3.1 Preparation of the QESD FA .....................................................................34

2.3.2 Powder X-ray Diffractometry (PXRD) ......................................................35

2.3.3 Thermogravimetry Analysis (TGA) ..........................................................35

2.3.4 Scanning electron microscopy (SEM) .......................................................35

2.3.5 Particle size distribution (PSD) ..................................................................35

2.3.6 Flowability measurement ...........................................................................36

2.3.7 Direct Compression of Powders ................................................................36

2.3.8 Tablet disintegration ..................................................................................37

2.3.9 Assessment of Tablet Tensile Strength ......................................................37

2.3.10 Friability .....................................................................................................37

2.3.11 Determination of HPMC in QESD FA ......................................................38

2.3.12 Dissolution performance ............................................................................38

2.4 Results and discussion .................................................................................................38

2.4.1 Flowability of as-received FA ...................................................................38

2.4.2 QESD preparation and micromeritic property ...........................................39

Page 10: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

viii

2.4.3 Powder flowability and tabletability ..........................................................39

2.4.4 Formulation development ..........................................................................40

2.5 Conclusion ...................................................................................................................42

2.6 References ....................................................................................................................53

Chapter 3: Profoundly improved plasticity and tabletability of griseofulvin by in-

situ solvation and desolvation during spherical crystallization ...................................56

3.1 Synopsis .......................................................................................................................57

3.2 Introduction ..................................................................................................................57

3.3 Materials and methods .................................................................................................59

3.3.1 Materials .......................................................................................................59

3.3.2 Spherical crystallization ................................................................................59

3.3.2.1 Construction of ternary phase diagram ..........................................59

3.3.2.2 Preparation of spherical agglomerates of GSF ..............................60

3.3.3 Powder X-ray Diffractometry (PXRD) .........................................................60

3.3.4 Single Crystal X-ray Diffractometry ............................................................61

3.3.5 Thermal Analyses .........................................................................................61

3.3.6 Powder Flowability .......................................................................................62

3.3.7 Powder tabletability ......................................................................................62

3.3.8 Scanning electron microscopy (SEM) ..........................................................63

3.3.9 In-die Heckel analysis ...................................................................................63

3.3.10 Out-of-die Kuentz–Leuenberger (KL) equation .........................................64

3.3.11 Energy framework of crystal structures ......................................................65

3.4 Results and discussion .................................................................................................65

Page 11: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

ix

3.4.1 Ternary Phase Diagram.................................................................................65

3.4.2 Particulate properties ....................................................................................66

3.4.3 Powder flowability and tabletability .............................................................66

3.4.4 Role of plasticity in the improved tabletability of GSF agglomerates..........68

3.4.5 Origin of superior plasticity of GSF SA .......................................................69

3.5 Conclusion ...................................................................................................................71

3.6 References ....................................................................................................................84

Chapter 4: Spherical cocrystallization - an enabling technology for the development

of high dose direct compression tablets of poorly soluble drugs .................................88

4.1 Synopsis .......................................................................................................................89

4.2 Introduction ..................................................................................................................89

4.3 Materials and methods .................................................................................................91

4.3.1 Materials .......................................................................................................91

4.3.2 Preparation of spherical GSF-Acs cocrystal hydrate ....................................92

4.3.3 Powder X-ray Diffractometry (PXRD) .........................................................93

4.3.4 Thermal Analyses .........................................................................................93

4.3.5 Scanning electron microscopy (SEM) ..........................................................93

4.3.6 Contact angle measurement ..........................................................................94

4.3.7 Energy framework and topologic analysis of crystal structures ...................94

4.3.8 Formulation and Tablet Compression ...........................................................95

4.3.9 Powder Flowability .......................................................................................96

4.3.10 Expedited Friability ....................................................................................96

4.3.11 Artificial Stomach-Duodenum Dissolution (ASD) .....................................97

Page 12: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

x

4.4 Results and discussion .................................................................................................98

4.4.1 Characterization ............................................................................................98

4.4.2 Formulation development ...........................................................................101

4.5 Conclusions ................................................................................................................104

4.6 References ..................................................................................................................120

Chapter 5: Simultaneously improving the manufacturability and dissolution of

indomethacin by a novel quasi-emulsion solvent diffusion based spherical

cocrystallization strategy ...............................................................................................124

5.1 Synopsis .....................................................................................................................125

5.2 Introduction ................................................................................................................125

5.3 Materials and methods ...............................................................................................128

5.3.1 Materials .....................................................................................................128

5.3.2 Preparation of IMC-SAC cocrystal powders ..............................................129

5.3.2.1 Fine IMC-SAC cocrystal by anti-solvent (AS) method ...............129

5.3.2.2 IMC-SAC spherical cocrystal agglomerates by QESD-CC method

..................................................................................................................129

5.3.3 Particle size distribution (PSD) ...................................................................130

5.3.4 Specific surface area (SSA) measurement ..................................................130

5.3.5 HPMC content determination via size-exclusion chromatography (SEC) .130

5.3.6 Powder X-ray Diffractometry (PXRD) .......................................................132

5.3.7 Thermal Analyses .......................................................................................132

5.3.8 Scanning electron microscopy (SEM) ........................................................133

Page 13: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xi

5.3.9 Computational methods ..............................................................................133

5.3.9.1 Energy framework and topologic analysis of crystal structures ..133

5.3.9.2 Drug – Polymer Intermolecular Interaction .................................134

5.3.10 Preparation of formulated powders ...........................................................134

5.3.11 Powder Flowability ...................................................................................135

5.3.12 Powder Compaction ..................................................................................135

5.3.13 Expedited Friability ..................................................................................136

5.3.14 Dissolution performance ...........................................................................136

5.3.15 Solution 1H Nuclear Magnetic Resonance (NMR) Spectroscopy ............137

5.3.16 Statistical Analysis ....................................................................................137

5.4 Results and discussion ...............................................................................................137

5.4.1 Micromeritic properties ..............................................................................137

5.4.2 Phase purity .................................................................................................139

5.4.3 Flowability and tabletability of various IMC forms ...................................139

5.4.4 Formulation development ...........................................................................141

5.4.5 Dissolution ..................................................................................................143

5.4.6 Drug-polymer molecular interaction...........................................................143

5.4.7 Formation mechanism of spherical cocrystal agglomerates .......................145

5.5 Conclusion .................................................................................................................146

5.6 References ..................................................................................................................170

Chapter 6: Reduction of punch-sticking propensity of celecoxib by spherical

crystallization via polymer assisted quasi-emulsion solvent diffusion ......................174

6.1 Synopsis .....................................................................................................................175

Page 14: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xii

6.2 Introduction ................................................................................................................175

6.3 Materials and methods ...............................................................................................177

6.3.1 Materials .....................................................................................................177

6.3.2 Polymer Screening for Quasi-Emulsion Solvent Diffusion (QESD) ..........177

6.3.3 QESD Spherical Agglomerate Growth Kinetics.........................................178

6.3.4 Particle Size Distribution ............................................................................178

6.3.5 Powder X-ray Diffractometry (PXRD) .......................................................179

6.3.6 Thermal Analyses .......................................................................................179

6.3.7 True density measurement ..........................................................................180

6.3.8 Powder Flowability, Tabletability, and Punch Sticking Assessment .........180

6.3.9 Compressibility and Compactibility Analyses............................................182

6.3.10 Scanning Electron Microscopy (SEM) .....................................................182

6.3.11 X-ray Photoelectron Spectroscopy (XPS) ................................................183

6.3.12 HPMC Content Determination .................................................................183

6.3.13 1H NMR Spectroscopy ..............................................................................184

6.4 Results and discussion ...............................................................................................184

6.4.1 Polymer Screening and Granule Growth Process .......................................185

6.4.2 Sticking propensity reduction by CEL-QESD ............................................186

6.4.3 Solid form and residual solvent ..................................................................187

6.4.4 Manufacturability of CEL Powders ............................................................187

6.4.5 Polymer Coating on CEL-QESD ................................................................189

6.4.6 Interactions between CEL and polymers ....................................................190

6.5 Conclusions ................................................................................................................192

Page 15: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xiii

6.6 References ..................................................................................................................206

Chapter 7: Research summary and future work ........................................................209

Bibliography ...................................................................................................................216

Page 16: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xiv

List of Tables

2.1 Powder and tablet design criteria .................................................................................43

3.1 Plasticity parameters of as received GSF and GSF SA from in-die and out-of-die data

analyses. .............................................................................................................................76

3.2 Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p) dispersion

corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)), polarization

(E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the energy are

listed. R indicated the distance between centers of mass of the pair of molecules. .... 82-83

4.1 Formulation Composition of GSF-containing tablet using different powder forms. 105

4.2 Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p) dispersion

corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)), polarization

(E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the energy

are listed. R indicated the distance between centers of mass of the pair of molecules. . 113-

114

4.3 Simulated pharmacokinetic parameters of various GSF-containing tablets (n =3) ...118

4.4 Summary of various GSF tablet formulations in meeting key criteria (meeting– Y, not

meeting – N) ....................................................................................................................119

5.1 N2 physisorption isotherms of QESD-CC and AS ............................................ 147-150

5.2 Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p) dispersion

corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)), polarization

(E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the energy

are listed. R indicated the distance between centers of mass of the pair of molecules .. 152-

153

5.3 Formulation composition of IMC-containing tablets using different IMC forms .....155

Page 17: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xv

6.1 Elemental composition of sample surfaces ................................................................204

Page 18: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xvi

List of Figures

1.1 Schematic over the capillary conditions between two crystals ....................................21

1.2 Phase diagrams of solvent systems containing (a) three solvents and (b) two solvents

in SA method (M: miscible, I: immiscible) .......................................................................22

1.3 Processes for spherical crystallization via the quasi-emulsion solvent diffusion method

............................................................................................................................................23

2.1 Flowability of mixtures between lactose 316 and as-received FA powder. The

flowability of Avicel PH102 is used as the minimum flowability for high speed tableting.

............................................................................................................................................44

2.2 TGA profile of QESD FA. ...........................................................................................45

2.3 PXRD patterns of FA, pure QESD, as-received, and calculated. ................................46

2.4 Particle size distribution of QESD and as-received FA powders. ...............................47

2.5 SEM images of FA particles: (a) QESD powder (low magnification), (b) QESD powder

(high magnification), (c) as-received powder (low magnification), and (d) as-received

powder (high magnification). .............................................................................................48

2.6 Manufacturability of pure as-received and QESD FA powders (a) tabletability and (b)

flowability. .........................................................................................................................49

2.7 (a) Tablet disintegration time as a function of crospovidone concentration in mixtures

with QESD FA (at 60 MPa compaction pressure), (b) tablet ejection force profiles with and

without 0.25% magnesium stearate. ..................................................................................50

2.8 Properties of QESD based FA formulation (a) tabletability, (b) flowability, (c) friability,

and (d) tablet dissolution in 6.8 pH buffer (in comparison to a formulation using as-received

FA). ....................................................................................................................................51

2.9 Manufacturability of formulated as-received and QESD FA powders (99% loading) (a)

tabletability and (b) flowability. ........................................................................................52

Page 19: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xvii

3.1 Ternary phase diagram of GSF in water-DMF-DCM solvents system. ......................72

3.2 Microscopic images of GSF (a) as-received and (b) SA. ............................................73

3.3 TGA profiles of GSF SA and as-received GSF. ..........................................................74

3.4 (a) Flowability profiles of as-received GSF, GSF SA and Avicel PH102 (n=1); (b)

tabletability profiles of as-received GSF, GSF SA and GSF SA milled (n=3). ................75

3.5 (a) Tabletability, (b) compactibility, (c) compressibility, and (d) hardness profiles of

GSF (80%) and Avicel PH102 (20%) mixture. .................................................................77

3.6 (a) PXRD patterns of calculated GSF Form I, as-received GSF, GSF SA, calculated

GSF-DCM and GSF SA before drying; (b) DSC profiles of as-received GSF and GSF SA.

............................................................................................................................................78

3.7 SEM images at (x10,000 magnification) of GSF (a) as-received crystals, (b) SA, and

tablet fracture surface of (c) GSF-Avicel mixture, and (d) GSF SA-Avicel mixture. .......79

3.8 SEM image at (x20,000 magnification) of GSF SA. ...................................................80

3.9 Energy framework of (a) GSF Form I and (b) GSF-DCM solvate. The (c) desolvated

GSF-DCM with large void is also shown for comparison. The energy threshold for the

energy framework is set at −22 kJ/mol. .............................................................................81

4.1 SEM images of a) GSF, b) GSF-Acs, c) GSF-Acs/SA and d) GSF-Acs/HPC. .........106

4.2 Microscopy images of (a) GSF-Acs/HPC and (b) GSF-Acs/SA. ..............................107

4.3 Particle size distribution of GSF-Acs/SA and GSF-Acs/HPC powders. ...................108

4.4 PXRD patterns of Acs-H, GSF, GSF-Acs/HPC, GSF-Acs/SA, GSF-Acs, and calculated

GSF-Acs. ..........................................................................................................................109

4.5 Thermal behavior of GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC characterized by

differential scanning calorimetry. ....................................................................................110

4.6 Tabletability profiles of pure GSF, GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC.....111

Page 20: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xviii

4.7 Energy frameworks of a) GSF, and b) GSF-Acs, both viewed into a axis of

corresponding unit cell. The interaction energy threshold was set at −5 kJ/mol. ............112

4.8 Manufacturability of formulated GSF, GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC, (a)

tabletability and (b) friability. ..........................................................................................115

4.9 Flowability plots of formulated GSF, GSF-Acs, GSF-Acs/SA and GSF-Acs/HPC..116

4.10 (a) ASD duodenum concentration profiles for formulated GSF, GSF-Acs, GSF-

Acs/SA and GSF-Acs/HPC tablets prepared at 250 MPa, (b) wettability of GSF-Acs/HPC,

GSF-Acs/SA and GSF-Acs. .............................................................................................117

5.1 Correlation between the actual measured HPMC concentration by the SEC method.

..........................................................................................................................................151

5.2 Chemical structures of (a) IMC, (b) SAC and (c) HPMC. Hydrogen atoms that

participated in the intermolecular interactions responsible for QESD-CC to form are

marked..............................................................................................................................154

5.3 Optical microscopy images of particles prepared with aqueous solutions of HPMC in

different concentrations from 0% to 0.5% (scale bar = 500 µm) ....................................156

5.4 Particle size distribution of as-received IMC, AS and QESD-CC, indicating that the

QESD-CC yields highly spherical particles with a relatively narrower size distribution 157

5.5 Optical microscopy images of particles (a) with and (b) without 0.5% HPMC in the

crystallization medium (scale bar = 500 µm) ..................................................................158

5.6 PXRD patterns of calculated IMC-SAC, AS, QESD-CC), as-received IMC, and as-

received SAC ...................................................................................................................159

5.7 DSC profiles of as received IMC, AS, and QESD-CC. .............................................160

5.8 TGA profiles of as received IMC, AS and QESD-CC. .............................................161

5.9 SEM images of AS (a and b) and QESD-CC (c and d) powders at low (x 50) and high

(x 20,000) magnifications. Images e) and f) show the hollow structure of QESD-CC

particles. ...........................................................................................................................162

Page 21: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xix

5.10 Powder properties of as-received IMC, AS, and QESD-CC powders, a) flowability

and b) tabletability. ..........................................................................................................163

5.11 Energy frameworks of (a) IMC and (b) IMC-SAC with a likely slip layer shaded in

pink. The thickness of each cylinder (in blue) represents the relative strength of

intermolecular interaction. The energy threshold for the energy framework is set at −10

kJ/mol. ..............................................................................................................................164

5.12 Performance of tablet formulations based on as-received IMC, AS, and QESD-CC,

(a) flowability, (b) tabletability, c) tablet friability, and d) tablet dissolution. ................165

5.13 Partial 1H NMR spectra of (a) 2 mg/mL IMC-SAC, 2 mg/mL IMC-SAC with 0.5

mg/mL HPMC and 8 mg/mL IMC-SAC, (b) 0.5 mg/mL HPMC with and without 2 mg/mL

of IMC-SAC, and c) Intermolecular hydrogen bonding interactions in IMC-SAC cocrystal.

..........................................................................................................................................166

5.14 Partial 1H NMR spectra of (a) 2 mg/mL SAC, 8 mg/mL SAC and 2 mg/mL SAC with

0.5 mg/mL HPMC, and (b) 0.5 mg/mL HPMC and 2 mg/mL SAC with 0.5 mg/mL HPMC.

..........................................................................................................................................167

5.15 Optimized geometric complexes between IMC-SAC and three HPMC monomer units

with different substituents: a) -H, b) -CH3, c) -CH2CH(OH)CH3. .................................168

5.16 Formation process of QESD-CC particles ...............................................................169

6.1 Chemical structures of (a) the model API CEL, and commonly used polymeric

stabilizers (b) HPMC, (c) HPC, and (d) PVP. .................................................................193

6.2 Micrographs of agglomerates prepared by QESD method in the absence and presence

of various polymers..........................................................................................................194

6.3 Time-dependent growth of CEL spherical agglomerates in various aqueous polymer

solutions assessed by a polarized light microscopy: a) HPMC, b) HPC, and c) PVP. All

scale bars are 500 µm. ......................................................................................................195

6.4 Punch sticking propensity of CEL after compaction at 50 MPa: a) as-received CEL,

and CEL-QESD prepared from different concentrations of HPMC solutions b) 0.1%, c)

Page 22: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

xx

0.3%, d) 0.5%. (e) Quantitative sticking propensity assessment using formulated CEL

powders (79.5% (w/w) Avicel PH102, 20% CEL, and 0.5% MgSt). ..............................196

6.5 Thermal behavior of CEL as-received and CEL QESD (from a 0.5% HPMC solution)

a) differential scanning calorimetry, b) thermogravimetric analysis of as-received and

CEL-QESD samples, and c) PXRD patterns for calculated, as-received, and CEL-QESD

samples. ............................................................................................................................197

6.6 a) Flowability (at 1 kPa pre-shear normal stress) and b) Tabletability of CEL-QESD as

compared to as-received CEL. .........................................................................................198

6.7 a) particle size distributions of as-received CEL and CEL-QESD powders, b) HPMC

content in solid CEL-QESD plotted as a function of the HPMC solution concentration (n

= 3). ..................................................................................................................................199

6.8 Compaction properties of as-received CEL and CEL-QESD formulations consisting of

79.5 wt% Avicel PH102, 20 wt% CEL powder, and 0.5 wt% MgSt: a) tabletability and b)

compactibility. .................................................................................................................200

6.9 Compressibility plot of 20% (w/w) as-received CEL and CEL-QESD formulations in

79.5% Avicel PH102 and 0.5% Magnesium stearate. .....................................................201

6.10 SEM images of CEL-QESD (top row) and as-received CEL (bottom row) at low (left

column) and high (right column) magnifications. ...........................................................202

6.11 X-ray photoelectron spectral analysis of the particle surface elemental compositions

of as-received CEL and CEL-QESD, (a) powder, and (b) tablets. ..................................203

6.12 1D 1H NMR spectra of CEL (3 mg/mL) with different polymer additives (3 mg/mL)

in DMSO-d6 over the chemical shift ranges a) 7.80-7.90 ppm, b) 7.40-7.60 ppm, c)

7.12-7.28 ppm, and d) 2.20-2.40 ppm. The numbering of the resonances corresponds

to that given in Figure 6.1. ...............................................................................................205

Page 23: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

1

Chapter 1

Introduction

Page 24: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

2

1.1 Background

The tablet, one of the most preferred dosage form for the oral drug delivery, has

advantages over others dosage forms, including low manufacturing cost, long shelf life

because of both good physical and chemical stability, and high patient compliance.1, 2 A

successful tablet formulation should meet requirements on several properties, such as flow3,

4, tableting5, 6 and dissolution7, 8, in order to be manufactured in large scale and be

adequately absorbed in the gastrointestinal (GI) tract.

More than 70% active pharmaceutical ingredients (APIs) in the development

pipeline show poor solubility9, and many APIs frequently exhibit poor manufacturability

(e.g., flowability, tabletability)2, 10, posing major challenges in tablet formulation

development. Typically, functional excipients are blended with the API to achieve desired

manufacturability and dissolution properties.11 However, the incorporation of large

amounts of excipients results in large tablet size, which may cause poorer patient

compliance due to difficulty with swallowing, especially when the required drug dose is

high.2 Therefore, enhancing the properties of APIs to reduce the amount of excipients

required for making a high quality tablet is an attractive approach.

Spherical crystallization is a technique that can agglomerate small primary crystals

into large spherical agglomerates to improve flow, tableting and dissolution performance

of APIs.12-15 Based on the formation mechanism of spherical agglomerates, spherical

crystallization is mainly divided into two methods: spherical agglomeration (SA) and

quasi-emulsion solvent diffusion (QESD)16-18. Such spherical agglomerates usually

present excellent flowability owing to their more spherical shape and large particle size.19

In addition, the small primary crystals that form large spherical agglomerates may exhibit

Page 25: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

3

improved tabletability if large spherical agglomerates fracture into small crystals during

tablet compression, because of the larger bonding area of small primary crystals20.

Moreover, previous studies have shown that spherical agglomerates incorporated with

water-soluble polymers such as HPC13, 14, 21 and HPMC22 exhibit enhanced dissolution

because of the improved wettability of the API. However, spherical crystallization can

sometimes also reduce dissolution.16

The objective of this thesis is to improve the manufacturability and dissolution

performance of drugs through spherical crystallization technique. When the properties of

API are profoundly improved, amounts of excipients required for a successful tablet

formulation can be significantly reduced. Consequently, a high drug loading direct

compression tablet formulation can be developed. This thesis work focuses on the

following topics:

a) Developing high drug loading direct compression tablet formulation of soluble

APIs by spherical crystallization

b) High dose direct compression tablets of poorly soluble APIs by spherical

cocrystallization

c) Solid form transition during spherical crystallization process and its impact on

the mechanical property of APIs

d) Reducing punch sticking propensity of APIs by spherical crystallization based

on the QESD method

1.2 Literature review

Page 26: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

4

As a well-established particle engineering technique, spherical crystallization

modifies the size and shape of the API crystals to improve the micromeritic properties.12

Three common methods of spherical crystallization are SA, QESD and ammonia diffusion

(AD). Since the AD method is only applicable for the APIs that are soluble in ammonia

water,23 in this thesis work, we only focus on the two main methods SA and QESD. With

a clear understanding of both methods, we can develop suitable strategies for optimizing

properties of spherical agglomerates through effective process design and control.

1.2.1 Spherical agglomeration (SA)

The SA method was firstly developed by Kawashima et al., where needle-shape

salicylic acid was successfully transformed into large spherical agglomerates, presenting

better micromeritic property than the starting material.12 During the SA, a bridging liquid

is used to preferentially wet and agglomerate primary crystals into spherical agglomerates

under suitable hydrodynamics (agitation). This method consists of two steps:

crystallization and agglomeration, and depending on the order of introducing the bridging

liquid, these two processes can start simultaneously or separately.18, 24 Pena et al. studied

the impact of the order of introducing the liquids, including poor solvent, good solvent and

bridging liquid, into the system on the formation mechanisms of spherical agglomerates.25

In the first scenario, the API solution in a good solvent is first mixed with a bridging liquid,

which is then gradually fed into the poor solvent to form droplets containing bridging liquid

and good solvent. This is followed by API precipitation in the droplets due to the solvent

counter diffusion. Finally, the bridging liquid on the surface of crystals coalesces small

agglomerates into large and spherical agglomerates. In the second scenario, the bridging

Page 27: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

5

liquid is introduced after the generation of small primary crystals, which may be produced

by milling, cooling or anti-solvent, and the bridging liquid functions as a binder to

agglomerate the small crystals. The bridging liquid is crucial in the SA process in both

scenarios.

To be effective, a bridging liquid should have a high interfacial tension with the

crystallization medium in addition to good wettability with the solid.26 From a mechanistic

viewpoint, the crystals clustered together because of surface tension and capillary force

associated with liquid-liquid interface between the water-rich bulk solution (Liquid 2) and

the bridging liquid (Liquid 1) in the aggregates (Fig. 1).27 Adequate wettability of the

bridging liquid on the crystal surfaces is a prerequisite for developing a capillary force

between two particles, which pulls the two particles together. In such a situation, a

meniscus forms between the liquids, i.e., the contact angle becomes less than 90o (Figure

1.1), which leads to a lower pressure on the liquid 1 side than the liquid 2 side. This pressure

difference pull the two particles together and can be calculated using Eq. 1.128.

∆𝑃 = 𝐶1−𝑝

𝑝

𝛾

𝐷cos 𝜃 Eq. (1.1)

where C is a constant depending on the specific surface area of the particles, p is

the porosity of the agglomerate, D is the mean diameter of the agglomerate, γ is the

interfacial tension between the liquids (water and bridging liquid) and θ is the contact angle.

In addition, the amount of bridging liquid introduced to generate the spherical

agglomerates must be well controlled during spherical agglomeration process because the

volume ratio between the bridging liquid and solid particle, BSR, plays a key role in

determining the property of spherical agglomerates. Too low and too high values of BSR

Page 28: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

6

lead to insufficient agglomeration and past-like solids, respectively.29 Therefore,

preliminary study to determine the optimal BSR and accurate control of bridge liquid

amount during processing is required to robustly produce spherical agglomerates in large

scale.

1.2.1.1 Selection of a bridging liquid

A suitable bridging liquid is partially miscible or immiscible with the mixture of

good and poor solvent, since a separate phase of bridging liquid is required to wet the

primary crystals and drives the formation of spherical agglomerates. Additionally, the

bridging liquid must wet the solid. Hence, contact angle determination between liquid and

solid can guide the selection of a suitable bridging liquid. A liquid with a lower contact

angle indicates better wettability. Gonza et al.29 proposed and validated a method for

selecting the best bridging liquid based on a Washburn’s test where capillary rise of liquids

in a granular medium was used to determine contact angle. However, wettability is not the

only criterion for selecting a good bridging liquid. Other properties of the bridging liquid

such as viscosity, polarity and the interfacial tension between the bridging liquid rich phase

(liquid 1) and poor solvent rich phase (liquid 2) also influence the formation of spherical

agglomerates. 26

Another commonly used method to select the bridging liquid is the solubility

measurement. Jitkar et al.30 selected the bridging liquid for etodolac by determination the

solubility of API in various solvents. A moderate solubility of the API in the bridging

liquid is preferred. When the solubility of API in the bridging liquid is high, a small amount

of bridging liquid can dissolve the API crystals completely. Therefore, the rate of feeding

Page 29: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

7

bridging liquid into the suspension and efficiency of mixing influence agglomeration of

particles, which makes the process less reproducible and difficult to scale up. When the

solubility is too low, the bridging liquid usually does not wet the API particle readily. A

very low solubility also effectively eliminates the possibility of solid bridge formation upon

drying of the bridging liquid. Spherical agglomerates are usually too fragile to be used in

subsequent steps to manufacture tablets.

An efficient approach to screen for a suitable bridging liquid is to calculate the

adhesion free energy between the liquid and API crystal using a mathematic model or

computational tool 31. A negative adhesion value indicates the attraction between the

crystal and liquid so that wetting can occur. Chen et al.32 applied the Lifshitz-van der Waals

acid-base approach to calculate the adhesion free energy, through which a solvent system

for cefotaxime sodium and benzoic acid was successfully selected. Pagire et al.33 predicted

the dominant surface of the API crystal based on the calculated growth morphologies and

calculated the heats of adsorption of various solvents onto the dominant crystal faces. The

solvent with high heats of adsorption was selected and successfully used as the bridging

liquid to form spherical agglomerates.

1.2.1.2 Phase diagram

Ternary phase diagram of poor solvent, good solvent, and bridging liquid can be

used to guide the spherical agglomeration process development (Figure 1.2a). Spherical

agglomeration can only occur in the immiscible region, where free bridging liquid is

available. If the crystallization of API is induced by mixing the good solvent and the poor

solvent before a bridging liquid is added, the starting point of solvent system during the

Page 30: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

8

SA process lies somewhere on the line between the good and poor solvents in the ternary

phase diagram. As the bridging liquid is gradually introduced into the solvent system, the

composition of solvent system initially enters the miscible region and then enters the

immiscible region, where free bridging liquid is available. The boundary between the

miscible and immiscible zone is used to calculate the BSRmin. When the amount of

bridging liquid exceeds a certain amount a paste-like product form. This composition

marks the maximum BSR for a given SA process (BSRmax). Therefore, spherical

agglomerates can be produced only between BSRmin and BSRmax (the shaded region in

Figure 1.2). Understandably, a larger difference between BSRmin and BSRmax, i.e., wider

shaded area in Figure 1.2, corresponds to a more robust SA process. This is another

criterion that can be used to select a suitable bridging liquid. To construct the ternary phase

diagram, poor and good solvents at different volume ratios are mixed and maintained at a

preset temperature, followed by introducing the bridging liquid into the solvent system

carefully with a pipette or a microsyringe and the mixture was shaken vigorously by hand.

When the phase separation occurs, a small droplet of bridging liquid can be observed, and

the volume of bridging liquid is noted. A die indicator, such as methylene blue in the

solvent system may help to mark the formation of a separate phase of bridging liquid.

Finally, the phase separation points of each mixture are connected, where the upper and

lower side of the phase separation line implies miscible and immiscible regions,

respectively.34

In a two solvents system, consisting a poor solvent and a bridging liquid, the binary

phase diagram is relevant (Figure 1.2b). In this process, the primary crystals are produced

by milling or cooling instead of an anti-solvent method. Therefore, a good solvent is not

Page 31: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

9

required to produce fine API crystals. Similar to that in the ternary phase diagram, when

the bridging liquid is gradually added into the poor solvent, the solvent composition moves

into the miscible region initially, and then the immiscible region when the amount of

bridging liquid exceeds its solubility in the poor solvent. Again, an excessive amount of

bridging liquid causes paste-like product. Spherical agglomerates can be generated only

in the shaded zone in Figure 1.2b.

It should be mentioned that not the entire shaded area (between BSRmin and BSRmax)

is suitable for producing high quality spherical agglomerates. In the shaded area near the

miscible region, most of solid suspended in the medium are small primary crystals because

of the insufficient amount of bridging liquid to enable complete agglomeration. In the

shaded area on near the immiscible region, particles may be too large to be used for

manufacturing pharmaceutical products. An optimal amount of bridging liquid for SA,

corresponding to BSRopt, depends on the API property and processing parameters, such as

solvent addition rate and mixing intensity.

Key processing parameters that influence the quality of spherical agglomerates

include agitation speed, agitation time, BSR value and temperature. Previous studies have

shown that increasing the agitation speed could reduce the particle size significantly due to

enhanced shear force of agitated liquid on the particle, breaking down the large particles

into smaller ones.18 Increasing the value of BSR could enlarge the particles since more

bridging liquid is available.35, 36 Agitation time may or may not influence the particle size

depending on the amount of bridging liquid introduced during the SA process. Morishima

et al.18 showed that, in the SA method, the size of spherical agglomerates was independent

of the agglomeration time when BSR value is low, while strong correlation between the

Page 32: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

10

agitation time and agglomerate size was observed when BSR exceeds a critical value. In

addition, the longer agitation time can lower the porosity of the particles since the particles

have more collisions with other particles and the wall of vessel, which strengths the

particles.35 The influence of processing temperature on the spherical agglomerates is

complex. Kawashima et al.37 found that the size of salicylic acid spherical agglomerates

decreased with increasing temperature below a critical temperature of 10 oC, but the size

increased with increasing temperature above 10 oC. The different trends were attributed to

the competition between the amount of crystals precipitated out, induced by the anti-

solvent process, and the solubility of the bridging liquid in the existing medium. This is

because that the solubility of bridging liquid and salicylic acid in the crystallization

medium increased with temperature at different rates.

1.2.2 Quasi-emulsion solvent diffusion (QESD)

QESD is another agglomeration method of spherical crystallization technique

developed by Kawashima et al.38 Different from the SA method, QESD normally involves

two miscible solvents, good solvent and poor solvent (Figure 1.3), which represent

dispersed and continuous phase, respectively. When a solution of an API in a good solvent

is mixed with a poor solvent under agitation, instead of immediate crystallization in SA

method, transient emulsions form. In some cases, emulsifiers are required to stabilize the

emulsions.39, 40 Crystallization of API crystals occurs when the solvent composition

changes due to the good solvent diffusing out the droplets while the poor solvent diffusing

in.41

Page 33: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

11

Since the supersaturation level of API at the droplet surface is the highest,

crystallization initiates from the surface of emulsion and then the crystals gradually grow

inward until all the API inside of the emulsion is consumed,18 which may lead to the

spherical agglomerates with a hollow structure.39 Therefore, the size of the resulting

spherical agglomerates is defined by the size of emulsion, which can be influenced by the

agitation speed and concentration of emulsifier.18, 34 Residual good solvent in the droplets

acted as a bridging liquid to agglomerate the precipitated crystals.42

Agitation speed and polymer concentration are the two most important process

parameters that affect the quality of spherical agglomerates by QESD. As shown by

Morishima et al.18, the diameter of the microspheres produced by QESD is mainly

determined by the size of droplet in the initial state, where a higher agitation speed led to

smaller microspheres. The concentration of polymer can influence the size18, 39 and

smoothness38 of microspheres. When other parameters were the same, a higher

concentration of HPMC led to smaller microspheres with higher sphericity because HPMC

on the emulsion surface during QESD prevents the coalescence of droplets. The agitation

time has little impact on the property of microspheres. The size of droplet increases slightly

at the beginning when the two phases are mixed together due to the coalescence of small

droplets. The size of droplet remains constant regardless of the duration of agitation.18

1.2.3 Powder property

1.2.3.1 Micromeritic property

Particle size and shape influence not only the downstream processes, such as

filtration and drying, but also bulk powder properties, such as bulk and tapped density,

Page 34: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

12

flowability, and tabletability.43, 44 Usually large and spherical particles with a high bulk

density are preferred. However, very large spherical particle can lead to segregation

problems when mixed with other excipients. Hence, the size of spherical agglomerates

needs to be controlled within a suitable range for pharmaceutical manufacturing.

1.2.3.2 Flowability

During high speed tablet manufacturing, adequate flow property of a formulation

is required for consistent powder filling, which is critical to reduce the weight variation of

tablets and maintain content uniformity.45 The flowability of a powder is affected by

particle size distribution, particle shape, chemical composition of particles, moisture, and

temperature.19, 46, 47 In general, large and more spherical particles exhibit better flowability

than small and irregular particles. The spherical agglomerates of APIs such as salicylic

acid48, ibuprofen38, acebutolol hydrochloride42, bucillamine49, clopidogrel hydrogen

sulfate50, etodolac30, albendazole13 and lactose51, generated by SA or QESD processes,

showed improved flowability over the starting material as indicated by lower angle of

repose, determined by a fixed funnel method. However, an objective criterion for adequate

flow property is needed to evaluate whether or not the improvement is sufficient. A grade

of microcrystalline cellulose, Avicel PH102, was shown to lie near the borderline between

acceptable and unacceptable flowability regions during high speed tableting. 4 Therefore,

one task required to carry out in the development of a spherical crystallization process is

to compare the flowability of spherical agglomerates to Avicel PH102. Spherical

agglomerates with flow property poorer than Avicel PH102 likely exhibit flow problems

Page 35: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

13

and further improvement in the flowability is recommended, especially if a direct

compression formulation is sought.

1.2.3.3 Tabletability

Tabletability, the ability of a powder to form a tablet upon compression, is another

important property of a powder in the pharmaceutical research.20, 52 It is represented by a

plot of the tablet tensile strength as a function of compaction pressure. Adequate tablet

tensile strength is required for a tablet to withstand the stresses during shipping and

handling to remain intact. Based on the bonding area (BA)-bonding strength (BS) model,20

the total interparticular area and strength of bonding over a unit area determine the strength

of a tablet. BA and BS can be assessed using compressibility and compactibility plots,

respectively.

Because BS is mainly determined by the functional groups on the surface of crystals,

the particle size and shape modifications through spherical agglomeration is expected to

modify BS but not BA, similar to wet or dry granulation.53-55 In absence of extensive

fragmentation, larger particles reduce the surface area available for bonding among

particles. Therefore, the observation of improved tableting performance analysis of

spherical agglomerates generated by SA method 56, 57 suggested unique mechanism(s) that

leads to stronger tablets. Martino et al.58 observed the vertical cross section of a tablet under

SEM and found that the large spherical agglomerates had fragmented into small particles

during tablet compaction. This corresponds to larger SA than that predicted from the size

of agglomerates. Similar observations were made on naproxen21 and ascorbic acid.59 The

Page 36: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

14

agglomerates of ibuprofen38 and acebutolol hydrochloride42 generated by the QESD

method also showed improved tableting performance. Different from the SA method, in a

QESD process, a suitable polymer is introduced to stabilize the emulsion, and this polymer

may coat onto the surface of the spherical agglomerates after the QESD process, which can

potentially influence the tableting performance of a powder. Polymer coating on the sand

to form particles with a core/shell structure could dramatically improve the tabletability of

sand due to the formation of an extensive strong 3D bonding network in the tablet by

compaction.6

1.2.3.4 Punch sticking

The sticking of API onto punches during compression is one of the common

problems that must be overcome for successful tablet manufacturing.60 Punch sticking is

detrimental to not only the aesthetics but also the quality of a tablet due to poor content

uniformity. 61 Powder may accumulate on the punch surface as the compaction process

proceeds, which eventually weakens the tablets and leads to a failed batch.62, 63

Previous approaches to mitigate the punch sticking propensity have been

extensively explored by applying both crystal and particle engineering techniques,

including salt formation61, mixing with excipients64, dry granulation65. Similar to the dry

granulation process,66 which reduces the punch sticking propensity due to the enlarged

particle size, spherical crystallization has the potential to reduce the punch sticking.

However, this has not been investigated to address the punch sticking problem.

Page 37: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

15

1.2.3.5 Dissolution

Besides the manufacturability (e.g., flowability and tabletability) discussed above,

dissolution is another property that is crucial for a successful formulation since it

determines the bioavailability of a drug.67, 68 This is particularly important when

considering that over 70% of APIs under development have low water solubility,69, 70

which implies slow dissolution of an API from a tablet in the GI tract. If not properly

addressed, such slow dissolution can significantly influence the absorption, and

consequently bioavailability, of the API.

Factors that can affect the dissolution include solubility, particle size, solid form

stability, use of precipitation inhibitor (polymer/surfactant).71 For low solubility APIs, i.e.,

BCS II and IV drugs, 72 more soluble solid forms, such as cocrystals, salts and amorphous

solids are commonly explored to improve the dissolution.73-75 Particle size influences

dissolution because it affects the area of an API in contact with the medium.76 However,

smaller particles tend to exhibit poor flowability. In practice, an API is milled or

micronized to increase area for faster dissolution but micronized API is then granulated

with excipients to ensure adequate flow property for successful tablet manufacturing. A

disintegrant is also usually incorporated in a tablet formulation to promote the dispersion

of API in the medium.77 If a more soluble solid form is selected to develop a tablet

formulation, the stability of the soluble solid form both in solid state and during dissolution

should be maintained, by preventing the crystallization of the less soluble crystal form, to

maintain the desired fast dissolution.

Page 38: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

16

Spherical crystallization can lead to variations in solid form of the API, depending

on the solvents and rate of crystallization during this process. Kawashima et al.78 found

that when indomethacin and epirizole were agglomerated together, different solvent

systems could produce different polymorphs combinations, which further influenced the

dissolution of the API. Similarly, Varshosaz et al.79 observed enhanced dissolution of

spherical agglomerates than untreated powder because of particle solid form change of

simvastatin from crystalline into amorphous in the spherical agglomerates. The

incorporation of a water-soluble polymer or surfactant in the spherical agglomerates

improved the dissolution of tolbutamide22 and etodolac30 , due to enhanced wettability.

1.2.4 Continuous manufacturing

Scaling up a spherical crystallization process, from the lab scale to the commercial

scale, requires full understanding of the key process parameters.80 The batch manufacturing

is the main method in the productions of pharmaceuticals for a long time. However, batch

to batch variation is a concern. Continuous manufacturing has received much attention in

the pharmaceutical industry because of its efficiency, economy, better quality control, and

easiness to scale-up.81, 82, 83

For successful spherical crystallization, several parameters, such as agitation speed,

agitation time, temperature, and feeding speed of bridging liquid, must be well controlled.

These parameters invariably change upon scaling up during the batch manufacturing.

Therefore, scale up of spherical crystallization is complex, which presents a major

challenge in the product development process. In this regard, continuous a spherical

Page 39: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

17

crystallization process is favored since the agglomerates are generated and monitored in

real-time and process parameters can be adjusted on the fly as needed. Additionally, a

continuous manufacturing process only needs a small volume of material to run, which

makes optimization and control an much easier and robust effort.84

Because of these benefits, continuous spherical crystallization has been explored in

pharmaceutical field. Pena et al.80 developed a two-stage continuous mixed suspension

mixed product removal (MSMPR) system, where the nucleation/growth and agglomeration

of crystals are decoupled, for the continuous manufacturing of the spherical agglomerates

with desired properties. This continuous MSMPR process allows to tailor both the

properties of primary crystals and generated agglomerates in a SA process to achieve good

biopharmaceutical (e.g., dissolution) and processing (e.g., filtering, drying) performance.

Similarly, Tahara et al.85 applied this continuous MSMPR system to manufacture the

spherical agglomerates of albuterol sulfate. Different from Pena’s work, Tahara et al.

applied the QESD method during the continuous manufacturing. One benefit of using the

QESD method for continuous manufacturing is the easy solvent recycling, where 90% of

the mother liquor was reused during the continuous spherical crystallization,85 making it to

a greener manufacturing process. Moreover, the continuous QESD process, together with

an effective purification process by agglomerating with a complexing agent, can be used

for the impurity separation. This was demonstrated using a model drug, fenofibrate.86

1.2.5 Spherical cocrystallization

Page 40: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

18

Spherical cocrystallization, which combines spherical crystallization and

cocrystallization, was firstly applied on the stabilization and spheroidization of ammonium

nitrate to produce smokeless and easy packing cocrystal agglomerates.87 Pagire et al.

applied this technique to prepare spherical carbamazepine-saccharin cocrystals.33 Spherical

cocrystallization was further carried out to control the cocrystal stoichiometry and particle

size.88 Despite the development of the spherical cocrystallization process, properties of the

spherical cocrystal agglomerates and their applications in tablet formulation are very

limited.

1.3 Objective and hypothesis

The overarching goal of this thesis work is to enable the development of high drug

loading direct compression tablet formulations by engineering APIs to improve

manufacturability (e.g., flowability, tabletabillity) and dissolution of drugs, through

applying the spherical crystallization technique.

Chapter 2 aims to develop an extremely high drug loading formulation using the

spherical crystallization of a water-soluble API. The main challenges with developing a

high drug loading tablet formulation, i.e., the manufacturability and dissolution deficiency

of the API, can be overcome by API engineering.

Hypothesis: API agglomerates produced through a spherical crystallization

process show significantly improved manufacturability (e.g., flowability, tabletability),

which enables the development of an extremely high drug loading direct compression

formulation for a water soluble API.

Page 41: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

19

Chapter 3 reports an in-situ solid form transition during the SA process, through

which the free-flowing spherical agglomerates exhibits a micro porous structure. This

unique structure leads to an improved plasticity of API bulk solid and, consequently, better

tabletability.

In Chapter 4 and 5, we aim to simultaneously improve the manufacturability and

dissolution of BCS Class II drugs by the spherical cocrystallization technique. Previous

efforts that showed improved dissolution performance of spherical agglomerates were

attributed to the transformation into a more soluble solid form or the incorporation of

polymer that increases the wettability of API. However, the stability of the more soluble

solid form is a concern, which requires extensive investigations before they can be

developed into a formulation. On the other hand, the low degree of enhancement in

dissolution by the improved wettability makes it not very effective to address dissolution

problems of poorly soluble drugs. Both issues can be addressed using the spherical

cocrystallization technique.

Hypothesis: spherical cocrystallization technique can take the advantages of both

spherical crystallization and cocrystallization, and therefore, the generated spherical

cocrystal agglomerates that possess appropriate pharmaceutical properties to enable the

development of high drug loading direct compression tablet formulations.

The spherical cocrystallization technique in Chpater 4 and 5 is based on SA and

QESD method, respectively. Spherical cocrystallization based on SA method can be

developed only if the cocrystal is stable in the complex solvent system and can be bound

by a bridging liquid to form the spherical agglomerates. On the other hand, to develop the

Page 42: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

20

spherical cocrystallization based on QESD method, the cocrystal must form through an

anti-solvent process and a suitable polymer may be required to stabilize the emulsion.

In Chapter 6, we investigate the mitigation of API punch sticking by a QESD

method. Based on the mechanism of a QESD process, we examine the effectiveness of a

polymer assisted QESD method to reduce the punch sticking of a highly sticky API.

Hypothesis: While functioning as a stabilizer of the emulsions during the QESD

process, a polymer can also coat the surface of solid spherical agglomerates. Such polymer

coating is an effective physical barrier to prevent the direct contact between the API and

punch during tablet compaction, which consequently mitigates the punch sticking problem.

Page 43: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

21

Figure 1.1. Schematic over the capillary conditions between two crystals

Page 44: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

22

Figure 1.2. Phase diagrams of solvent systems containing (a) three solvents and (b) two

solvents in SA method (M: miscible, I: immiscible)

a) b)

Page 45: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

23

Figure 1.3. Processes for spherical crystallization via the quasi-emulsion solvent diffusion

method.

Page 46: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

24

1.4 Reference

1. Gupta, H.; Bhandari, D.; Sharma, A. Recent trends in oral drug delivery: A review.

Recent Pat. Drug Deliv. Formul. 2009, 3, 162-73.

2. Roopwani, R.; Buckner, I. S. Co-processed particles: An approach to transform

poor tableting properties. J. Pharm. Sci. 2019, 108, 3209-3217.

3. Taylor, M. K.; Ginsburg, J.; Hickey, A. J.; Gheyas, F. Composite method to

quantify powder flow as a screening method in early tablet or capsule formulation

development. AAPS PharmSciTech 2000, 1, 20-30.

4. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

5. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine

active pharmaceutical ingredient powder blends and tablets at high drug loading via dry

particle coating. Int. J. Pharm. 2018, 543, 288-299.

6. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

7. Braig, V.; Konnerth, C.; Peukert, W.; Lee, G. Enhanced dissolution of naproxen

from pure-drug, crystalline nanoparticles: A case study formulated into spray-dried

granules and compressed tablets. Int. J. Pharm. 2019, 554, 54-60.

8. Ullah, M.; Hussain, I.; Sun, C. C. The development of carbamazepine-succinic

acid cocrystal tablet formulations with improved in vitro and in vivo performance. Drug

Dev. Ind. Pharm. 2016, 42, 969-976.

9. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

10. Chen, H.; Wang, C.; Sun, C. C. Profoundly improved plasticity and tabletability

of griseofulvin by in situ solvation and desolvation during spherical crystallization. Cryst.

Growth Des. 2019, 19, 2350-2357.

11. Chen, L.; Ding, X.; He, Z.; Huang, Z.; Kunnath, K. T.; Zheng, K.; Davé, R. N.

Surface engineered excipients: I. Improved functional properties of fine grade

microcrystalline cellulose. Int. J. Pharm. 2018, 536, 127-137.

12. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

13. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

14. Garala, K. C.; Patel, J. M.; Dhingani, A. P.; Dharamsi, A. T. Preparation and

evaluation of agglomerated crystals by crystallo-co-agglomeration: An integrated approach

of principal component analysis and box-behnken experimental design. Int. J. Pharm. 2013,

452, 135-56.

Page 47: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

25

15. Garala, K. C.; Patel, J. M.; Dhingani, A. P.; Dharamsi, A. T. Quality by design

(qbd) approach for developing agglomerates containing racecadotril and loperamide

hydrochloride by crystallo-co-agglomeration. Powder Technol. 2013, 247, 128-146.

16. Ribardière, A.; Tchoreloff, P.; Couarraze, G.; Puisieux, F. Modification of

ketoprofen bead structure produced by the spherical crystallization technique with a two-

solvent system. Int. J. Pharm. 1996, 144, 195-207.

17. Subero-Couroyer, C.; Mangin, D.; Rivoire, A.; Blandin, A. F.; Klein, J. P.

Agglomeration in suspension of salicylic acid fine particles: Analysis of the wetting period

and effect of the binder injection mode on the final agglomerate size. Powder Technol.

2006, 161, 98-109.

18. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

19. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

20. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

21. Maghsoodi, M.; Taghizadeh, O.; Martin, G. P.; Nokhodchi, A. Particle design of

naproxen-disintegrant agglomerates for direct compression by a crystallo-co-

agglomeration technique. Int. J. Pharm. 2008, 351, 45-54.

22. Sano, A.; Kuriki, T.; Handa, T.; Takeuchi, H.; Kawashima, Y. Particle design of

tolbutamide in the presence of soluble polymer or surfactant by the spherical crystallization

technique: Improvement of dissolution rate. J. Pharm. Sci. 1987, 76, 471-474.

23. Masumi Ueda, Y. N. a. Y. K. Particle design of enoxacin by spherical

crystallization technique. I principle of ammonia diffusion system (ads). Chem. Pharm.

Bull. 1990, 38, 2537-2541.

24. Chow, A. H. L.; Leung, M. W. M. A study of the mechanisms of wet spherical

agglomeration of pharmaceutical powders. Drug Dev. Ind. Pharm. 1996, 22, 357-371.

25. Peña, R.; Jarmer, D. J.; Burcham, C. L.; Nagy, Z. K. Further understanding of

agglomeration mechanisms in spherical crystallization systems: Benzoic acid case study.

Crys. Growth Des. 2019, 19, 1668-1679.

26. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

27. Butt, H.-J.; Kappl, M. Normal capillary forces. Adv. Colloid Interface Sci. 2009,

146, 48-60.

28. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

29. Amaro-Gonza, D.; Biscans, B. a. Spherical agglomeration during crystallization of

an active pharmaceutical ingredient. Powder Technol. 2002, 128, 188-194.

Page 48: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

26

30. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

31. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Cryst. Growth Des. 2013, 13, 4162-4167.

32. Chen, M.; Liu, X.; Yu, C.; Yao, M.; Xu, S.; Tang, W.; Song, X.; Dong, W.; Wang,

G.; Gong, J. Strategy of selecting solvent systems for spherical agglomeration by the

lifshitz-van der waals acid-base approach. Chem. Eng. Sci. 2020, 220, 115613.

33. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

34. Yoshiaki Kawashima, F. C. Parameters determining the agglomeration behaviour

and the micromeritic properties of spherically agglomerated crystals prepared by the

spherical crystallization. Int. J. Pharm. 1995, 119, 139-147.

35. Blandin, A. F.; Mangin, D.; Rivoire, A.; Klein, J. P.; Bossoutrot, J. M.

Agglomeration in suspension of salicylic acid fine particles: Influence of some process

parameters on kinetics and agglomerate final size. Powder Technol. 2003, 130, 316-323.

36. Thati, J.; Rasmuson, A. C. On the mechanisms of formation of spherical

agglomerates. Eur. J. Pharm. Sci. 2011, 42, 365-79.

37. Kawashima, Y.; Okumura, M.; Takenaka, H. The effects of temperature on the

spherical crystallization of salicylic acid. Powder Technol. 1984, 39, 41-47.

38. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

39. Nocent, M.; Bertocchi, L.; Espitalier, F.; Baron, M.; Couarraze, G. Definition of a

solvent system for spherical crystallization of salbutamol sulfate by quasi‐ emulsion

solvent diffusion (qesd) method. J. Pharm. Sci. 2001, 90, 1620-1627.

40. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion method. Chem. Pharm. Bull. 1990, 38, 733-739.

41. Zhou, X.; Zhang, Q.; Xu, R.; Chen, D.; Hao, S.; Nie, F.; Li, H. A novel spherulitic

self-assembly strategy for organic explosives: Modifying the hydrogen bonds by polymeric

additives in emulsion crystallization. Cryst. Growth Des. 2018, 18, 2417–2423.

42. Kawashima, Y.; Cui, F.; Takeuchi, H.; Niwa, T.; Hino, T.; Kiuchi, K.

Improvements in flowability and compressibility of pharmaceutical crystals for direct

tabletting by spherical crystallization with a two-solvent system. Powder Technol. 1994,

78, 151-157.

43. Kedia, K.; Wairkar, S. Improved micromeritics, packing properties and

compressibility of high dose drug, cycloserine, by spherical crystallization. Powder

Technol. 2019, 344, 665-672.

Page 49: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

27

44. Orlewski, P. M.; Ahn, B.; Mazzotti, M. Tuning the particle sizes in spherical

agglomeration. Cryst. Growth Des. 2018, 10, 6257-6265.

45. Sun, W.; Chen, H.; Aburub, A.; Sun, C. C. A platformdirect compression

formulation for low dose sustained-release tablets enabled by a dual particle engineering

approach. Powder Technol. 2019, 342, 856-863.

46. Teunou, E.; Fitzpatrick, J. J. Effect of relative humidity and temperature on food

powder flowability. J. Food Eng. 1999, 42, 109-116.

47. Li, Q.; Rudolph, V.; Weigl, B.; Earl, A. Interparticle van der waals force in powder

flowability and compactibility. Int. J. Pharm. 2004, 280, 77-93.

48. Kawashima, Y.; Okumura, M.; Takenaka, H.; Kojima, A. Direct preparation of

spherically agglomerated salicylic acid crystals during crystallization. J. Pharm. Sci. 1984,

73, 1535-1538.

49. Kenji, M.; Yoshiaki, K.; Hirofumi, T.; Toshiyuki, N.; Tomoaki, H.; Yoichi, K.

Tabletting properties of bucillamine agglomerates prepared by the spherical crystallization

technique. Int. J. Pharm. 1994, 105, 11-18.

50. Chen, M.; Du, S.; Zhang, T.; Wang, Q.; Wu, S.; Gong, J. Spherical crystallization

and the mechanism of clopidogrel hydrogen sulfate. Chem. Eng. Technol. 2018, 41, 1259-

1265.

51. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

52. Joiris, E.; Martino, P. D.; Berneron, C.; Guyot-Hermann, A.-M.; Guyot, J.-C.

Compression behavior of orthorhombic paracetamol. Pharm. Res. 1998, 15, 1122-1130.

53. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

54. Malkowska, S.; Khan, K. A. Effect of re-conpression on the properties of tablets

prepared by dry granulation. Drug Dev. Ind. Pharm. 1983, 9, 331-347.

55. Arndt, O. R.; Baggio, R.; Adam, A. K.; Harting, J.; Franceschinis, E.; Kleinebudde,

P. Impact of different dry and wet granulation techniques on granule and tablet properties:

A comparative study. J. Pharm. Sci. 2018, 107, 3143-3152.

56. Toshiyuki Niwa, H. T., Tomoaki Hino, Akira Itoh, Yoshiaki Kawashima, Katsumi

Kiuchi. Preparation of agglomerated crystals for direct tabletting and microencapsulation

by the spherical crystallization technique with a continuous system. Pharm. Res. 1994, 11,

478-484.

57. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. V. Improvement of

dissolution and bioavailability of direct compressed tablets prepared using tolbutamide

agglomerated crystals. Chem. Pharm. Bull. 1992, 40, 3030-3035.

58. Martino, P. D.; Cristofaro, R. D.; Barthelemy, C.; Joiris, E.; Filippo, G. P.; Sante,

M. Improved compression properties of propyphenazone spherical crystals. Int. J. Pharm.

2000, 197, 95-106.

Page 50: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

28

59. Kawashima, Y.; Imai, M.; Takeuchi, H.; Yamamoto, H.; Kamiya, K.; Hino, T.

Improved flowability and compactibility of spherically agglomerated crystals of ascorbic

acid for direct tableting designed by spherical crystallization process. Powder Technol.

2003, 130, 283-289.

60. Hutchins A, M. B., Mullerney MP. Assessing tablet sticking propensity. Pharm.

Tech. 2012, 36, 31-34.

61. Paul, S.; Wang, C.; Wang, K.; Sun, C. C. Reduced punch sticking propensity of

acesulfame by salt formation: Role of crystal mechanical property and surface chemistry.

Mol. Pharm. 2019, 16, 2700-2707.

62. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.;

Meenan, P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J.

Pharm. Sci. 2017, 106, 151-158.

63. Samiei, L.; Kelly, K.; Taylor, L.; Forbes, B.; Collins, E.; Rowland, M. The

influence of electrostatic properties on the punch sticking propensity of pharmaceutical

blends. Powder Technol. 2017, 305, 509-517.

64. Al-Karawi, C.; Leopold, C. S. A comparative study on the sticking tendency of

ibuprofen and ibuprofen sodium dihydrate to differently coated tablet punches. Eur. J.

Pharm. Biopharm. 2018, 128, 107-118.

65. Simmons, D. M.; Gierer, D. S. A material sparing test to predict punch sticking

during formulation development. Drug Dev. Ind. Pharm. 2012, 38, 1054-1060.

66. Bejugam, N. K.; Mutyam, S. K.; Shankar, G. N. Tablet formulation of an active

pharmaceutical ingredient with a sticking and filming problem: Direct compression and

dry granulation evaluations. Drug Dev. Ind. Pharm. 2015, 41, 333-341.

67. Zhang, Y.; Maier, W. J.; Miller, R. M. Effect of rhamnolipids on the dissolution,

bioavailability, and biodegradation of phenanthrene. Environ. Sci. Technol. 1997, 31,

2211-2217.

68. Amidon, G. L.; Lennernäs, H.; Shah, V. P.; Crison, J. R. A theoretical basis for a

biopharmaceutic drug classification: The correlation of in vitro drug product dissolution

and in vivo bioavailability. Pharm. Res. 1995, 12, 413-420.

69. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

70. O'Driscoll, C. M.; Griffin, B. T. Biopharmaceutical challenges associated with

drugs with low aqueous solubility—the potential impact of lipid-based formulations. Adv.

Drug Deliv. Rev. 2008, 60, 617-624.

71. Hörter, D.; Dressman, J. B. Influence of physicochemical properties on dissolution

of drugs in the gastrointestinal tract. Adv. Drug Deliv. Rev. 2001, 46, 75-87.

72. Savjani, K. T.; Gajjar, A. K.; Savjani, J. K. Drug solubility: Importance and

enhancement techniques. ISRN pharmaceutics 2012, 2012, 195727-195727.

Page 51: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

29

73. Indulkar, A. S.; Lou, X.; Zhang, G. G. Z.; Taylor, L. S. Insights into the dissolution

mechanism of ritonavir–copovidone amorphous solid dispersions: Importance of

congruent release for enhanced performance. Mol. Pharm. 2019, 16, 1327-1339.

74. Cao, F.; Amidon, G. L.; Rodríguez-Hornedo, N.; Amidon, G. E. Mechanistic basis

of cocrystal dissolution advantage. J. Pharm. Sci. 2018, 107, 380-389.

75. Li, S.; Wong, S.; Sethia, S.; Almoazen, H.; Joshi, Y. M.; Serajuddin, A. T. M.

Investigation of solubility and dissolution of a free base and two different salt forms as a

function of ph. Pharm. Res. 2005, 22, 628-635.

76. Sun, J.; Wang, F.; Sui, Y.; She, Z.; Zhai, W.; Wang, C.; Deng, Y. Effect of particle

size on solubility, dissolution rate, and oral bioavailability: Evaluation using coenzyme q₁₀

as naked nanocrystals. Int. J. Nanomed. 2012, 7, 5733-5744.

77. Bolhuis, G. K.; Zuurman, K.; te Wierik, G. H. P. Improvement of dissolution of

poorly soluble drugs by solid deposition on a super disintegrant. Ii. The choice of super

disintegrants and effect of granulation. Eur. J. Pharm. Sci. 1997, 5, 63-69.

78. Kawashima, Y.; Lin, S. Y.; Ogawa, M.; Handa, T.; Takenaka, H. Preparations of

agglomerated crystals of polymorphic mixtures and a new complex of indomethacin—

epirizole by the spherical crystallization technique. J. Pharm. Sci. 1985, 74, 1152-1156.

79. Varshosaz, J.; Tavakoli, N.; Salamat, F. A. Enhanced dissolution rate of

simvastatin using spherical crystallization technique. Pharm. Dev. Technol. 2011, 16, 529-

535.

80. Peña, R.; Nagy, Z. K. Process intensification through continuous spherical

crystallization using a two-stage mixed suspension mixed product removal (msmpr) system.

Cryst. Growth Des. 2015, 15, 4225-4236.

81. Konstantinov, K. B.; Cooney, C. L. White paper on continuous bioprocessing. May

20–21, 2014 continuous manufacturing symposium. J. Pharm. Sci. 2015, 104, 813-820.

82. Mascia, S.; Heider, P. L.; Zhang, H.; Lakerveld, R.; Benyahia, B.; Barton, P. I.;

Braatz, R. D.; Cooney, C. L.; Evans, J. M. B.; Jamison, T. F.; Jensen, K. F.; Myerson, A.

S.; Trout, B. L. End-to-end continuous manufacturing of pharmaceuticals: Integrated

synthesis, purification, and final dosage formation. Angew. Chem. 2013, 52, 12359-12363.

83. Lee, S. In Present and future for continuous manufacturing: Fda perspective,

Proceedings of the 3rd FDA/PQRI Conference on Advancing Product Quality, March,

2017; pp 22-24.

84. Lee, S. L.; O’Connor, T. F.; Yang, X.; Cruz, C. N.; Chatterjee, S.; Madurawe, R.

D.; Moore, C. M. V.; Yu, L. X.; Woodcock, J. Modernizing pharmaceutical manufacturing:

From batch to continuous production. J. Pharm. Innov. 2015, 10, 191-199.

85. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

86. Tahara, K.; Kono, Y.; Myerson, A. S.; Takeuchi, H. Development of continuous

spherical crystallization to prepare fenofibrate agglomerates with impurity complexation

using mixed-suspension, mixed-product removal crystallizer. Cryst. Growth Des. 2018, 18,

6448-6454.

Page 52: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

30

87. Lee, T.; Chen, J. W.; Lee, H. L.; Lin, T. Y.; Tsai, Y. C.; Cheng, S.-L.; Lee, S.-W.;

Hu, J.-C.; Chen, L.-T. Stabilization and spheroidization of ammonium nitrate: Co-

crystallization with crown ethers and spherical crystallization by solvent screening. Chem.

Eng. J. 2013, 225, 809-817.

88. Wang, J. R.; Bao, J.; Fan, X.; Dai, W.; Mei, X. Ph-switchable vitamin b9 gels for

stoichiometry-controlled spherical co-crystallization. Chem. Commun. 2016, 52, 13452-

13455.

Page 53: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

31

Chapter 2

Direct compression tablet containing 99% active ingredient – a

tale of spherical crystallization

Page 54: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

32

2.1 Synopsis

Direct compression (DC) is the easiest and most cost-effective process for tablet

manufacturing, since it only involves blending and compression. However, active

pharmaceutical ingredients (API) generally exhibit poor mechanical and micromeritic

properties, which necessitate dilution and the use of high percentage of excipients to enable

a robust DC manufacturing process. Consequently, drug loading in DC tablets is usually

low (typically < 30%, w/w). In this study, spherical crystallization by the quasi-emulsion

solvent diffusion method was used to engineer a poorly flowing model compound, ferulic

acid (FA), to attain superior mechanical properties, particle size distribution, and

morphology. The engineered FA particles enabled the successful development of DC

tablets containing 99% FA, which is in sharp contrast to the maximum 10% FA loading

using as-received FA. The record high API loading in this work illustrates the potential for

spherical crystallization to overcome low drug loading when developing a tablet product

using the DC manufacturing process.

2.2 Introduction

Tablets account for nearly 80% of all marketed dosage forms ,1 due to economical

and stability advantages than other dosage forms. Among the tablet manufacturing

processes, direct compression (DC) is the simplest, and hence, most desirable due to many

advantages, including significantly higher manufacturing efficiency and physical and

chemical stability. 2-4 3, 5

Page 55: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

33

Other than very potent drugs, a high drug loading in a tablet formulation can

improve patient compliance by reducing tablet size. Furthermore, it also reduces the risk

of an inhomogeneous blend or non-uniform active pharmaceutical ingredient (API) content

in finished tablets. However, API loading in most DC tablet formulations does not exceed

30 % due to the poor processability.1, 2, 5, 6 Both powder flowability and tabletability are

critical to ensure the manufacturability of a DC tablet formulation.1, 7, 8 Powders consisting

of large particles and spherical shape are expected to exhibit good flowability and packing

properties.3, 9 However, large API particle size can potentially lead to slow drug release,

and a spherical shape is rarely achieved in traditional crystallization. A small particle size

of the API maybe required to provide adequate dissolution, but this can lead to a poor

flowability.10 Poor flowability of API, in turn, limits the maximum achievable API loading

in a DC formulation.

Spherical crystallization is a promising method to produce good flowability and

tableting performance,11-13 where fine crystals are generated and then agglomerated to form

spherical structures.14 Recent papers on lactose,15 albendazole,12 etodolac16 etc have

showed the improved mechanical properties of spherical crystals with larger particle size,

and the improved fragmentation and plasticity are the main mechanism. Spherical

crystallization can be achieved by quasi-emulsion solvent diffusion (QESD), spherical

agglomeration, and ammonia diffusion methods as well as crystallo-co-agglomeration.17-19

In this work, the QESD method was applied to engineer a model API, ferulic acid (FA), to

attain improved bulk powder properties for developing a high API loading tablet. FA is a

natural phenolic phytochemical proposed for use as a medicament for treating disorders,

including Alzheimer’s disease,20 cancer,21 cardiovascular diseases,22 diabetes mellitus23

Page 56: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

34

and skin disease.24 The aim of this work was to develop a DC tablet containing a maximum

API loading by forming free-flowing spherical particles consisting of fine FA crystals.13

A successful formulation must ensure that the tablets meet requirements for

acceptable manufacturability, quality, and biopharmaceutical performance. As such, tablet

formulations are typically evaluated against established criteria for key critical

manufacturing parameter and tablet properties, including tabletability, flowability,

friability, disintegration time, tablet ejection force, and dissolution performance (Table 2.1).

If any is not met, the formulation is excluded, and a new formulation is designed and

evaluated to address identified deficiencies. This process is repeated until an optimized

formulation is identified. However, in contrast to empirical screening, we adopted a much

more efficient and material sparing development process, guided by the materials science

tetrahedron.25 In addition, we used predictive tools to characterize powder

manufacturability, e.g., compaction simulator for tabletability and shear cell for flowability

(see Supporting Information for details).

2.3 Methods

2.3.1 Preparation of the QESD FA

A quantity of 25 g ferulic acid was dissolved in 250 ml acetone at room temperature. The

FA solution was directly poured into 2 L 0.1% (w/v) HPMC (K15M) aqueous solution

stirred at 300 rpm by an overhead digital stirrer. After 1 h agitation, the QESD FA was

collected and dried at 60 °C in an oven.

Page 57: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

35

2.3.2 Powder X-ray Diffractometry (PXRD)

A powder X-ray diffractometer with Cu Kα radiation (1.54059 Å) was used to characterize

powders. Samples were scanned between 5 to 35° 2θ with a step size of 0.02° and a dwell

time of 1 s/step. The tube voltage and amperage were set at 40 kV and 40 mA, respectively.

The diffraction pattern of the QESD powder was compared to the original powder or the

calculated pattern from the single crystal structure (CCDC refcode: GASVOL02).

2.3.3 Thermogravimetry Analysis (TGA)

Approximately 2 mg of QESD FA was heated in an open aluminum pan from room

temperature to 300oC at 10oC/min under dry nitrogen purge (40 mL/min) on a

thermogravimetric analyzer (Model Q50; TA Instruments, New Castle, DE).

2.3.4 Scanning electron microscopy (SEM)

Particles were sprinkled onto carbon tapes adhered to a SEM stub. The samples were

sputter-coated with a thin layer of platinum (~ 75 Å thickness), using an ion-beams sputter

(IBS/TM200S; VCR Group Inc., San Clemente, CA). The particle morphology and surface

features were evaluated under a scanning electron microscope (JEOL 6500F; JEOL Ltd.,

Tokyo, Japan).

2.3.5 Particle size distribution (PSD)

Page 58: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

36

The particle size distribution of as-received and QESD FA powders were analyzed by sieve

analysis. Nine standard sieves (U.S.A. Standard Test Sieve) were used. The weight

retained on different sieves were measured after 10 min sieving interval. Sieve fractions

were calculated accordingly.

2.3.6 Flowability measurement

Powder flowability was measured using a ring shear cell tester (RST-XS; Dietmar Schulze,

Wolfenbüttel, Germany), with a 10 mL cell at 1 kPa preshear normal stresses. The normal

stresses for shear testing were 0.252, 0.4, 0.55, 0.7 and 0.85 kPa. The measurements were

made in triplicate, and the unconfined yield strength (fc) and major principal stress (σn)

were obtained from each yield locus by drawing Mohr’s circles. Flowability index (ffc) was

calculated using Eq. (1):

𝑓𝑓c =𝜎𝑛

𝑓c (1)

2.3.7 Direct Compression of Powders

A compaction simulator (Presster; Metropolitan Computing Corporation, NJ) simulating a

Korsch XL100 (10 stations) press at a speed corresponding to 25 ms dwell time (49,300

tablets/h) using 10.0 mm round flat-faced punch-die sets were used to prepare tablets with

compaction pressures ranging from 25 to 300 MPa. Tablet weight was maintained at 300

mg throughout this study.

Page 59: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

37

2.3.8 Tablet disintegration

A disintegration tester (DJ-1; Tianjin Guoming Medical Equipment CO., LTD) was used

to measure the disintegration time (DT) of the tablets compressed at 60 MPa for different

formulations. Tablets were immersed into a beaker containing 900 mL phosphate buffer,

and the temperature was kept at 37°C. DT was recorded for three tablets.

2.3.9 Assessment of Tablet Tensile Strength

Tablets were broken diametrically on a texture analyzer (Texture Technologies Corp.,

Surrey, UK) at a test speed of 0.01 mm/s. Tablet tensile strength, σ, was determined from

equation 2.26

𝜎 =2𝐹

𝜋𝑑ℎ (2)

Where F, d, and h are the breaking force, diameter, and thickness of the tablet, respectively.

2.3.10 Friability

An expedited friability method27 was applied to determine the friability performance of the

formulated powders. Coded tablets prepared under different compaction pressures were

weighted and loaded into a friabilator (Model F2, Pharma Alliance Group Inc., Santa

Clarita, CA) at 25 rpm for 4 min. The final weight of each tablet was noted, and percentage

weight loss of each tablet was plotted against compaction pressure. The compaction

pressure corresponding to 0.8% tablet weight loss was identified from the friability plot.

Page 60: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

38

2.3.11 Determination of HPMC in QESD FA

About 500 mg of QESD FA was dissolved in 10 ml DMF. The concentrated FA solution

was diluted with distilled water. The UV absorbance was measured at 315 nm, and the

amount of FA was determined from an appropriate standard curve. The amount of HPMC

was calculated from mass balance.

2.3.12 Dissolution performance

The dissolution performance of the tablets was studied with 500 mL of a pH 6.8 potassium

phosphate buffer. The concentration of dissolved FA was monitored in real-time by a UV

dip probe at 360 nm.

2.4 Results and discussion

2.4.1 Flowability of as-received FA

The as-received FA exhibits poor flowability (Figure 2.1). Therefore, a free-

flowing grade of lactose, Lactose 316,28 was used as a filler to address this deficiency.

Flowability decreases rapidly with increasing percent of FA in the mixture (Figure 2.1).

The maximum achievable loading of as-received FA was only 10% in order to meet the

minimum requirement for powder flowability (Figure 2.1, Table 2.1). To develop a high

loading FA tablet, the QESD method was used to overcome poor flowability.

Page 61: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

39

2.4.2 QESD preparation and micromeritic property

During the QESD process, one part of a solution of FA in acetone (100 mg/mL)

was introduced to eight parts of an aqueous solution of HPMC (1 mg/mL) under agitation.

HPMC was used to stabilize the transient emulsion.29, 30 Counter-diffusion between acetone

and water phases leads to the crystallization of FA within the emulsion to eventually obtain

relatively spherical agglomerates.19 The amount of HPMC in the dried QESD powder was

0.75 ± 0.64% (n = 3). Powder X-ray diffraction patterns of the as-received and the QESD

powders were identical and conformed to that calculated from FA single crystal structure

(Figure 2.2).31 The moisture and residual solvent were negligible (Figure 2.3).

The QESD powder had a narrow particle size distribution (most between 125-250

μm, Figure 2.4) and an approximately spherical morphology (Figure 2.5a). These micro-

porous particles consisted of bundles of small needle-like primary crystals (Figure 2.5b),

which likely arose from the high degree of supersaturation and fast precipitation of FA

during the QESD process. This hierarch particle structure of QESD particles is very distinct

compared to the as-received FA crystals, which are needles with a broad size distribution

(most higher than 500 μm, Figure 2.5c, Figure 2.4) and rough surfaces (Figure 2.5d). The

larger and more round QESD particles are expected to exhibit improved flowability over

the as-received FA.9

2.4.3 Powder flowability and tabletability

Tabletability and flowability of the QESD powder of FA are both significantly

better than the as-received FA (Figures 2.6a and 2.6b). Tablet tensile strength of ~6 MPa

Page 62: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

40

can be reached with the QESD powder, which is approximately twice that of the as-

received FA (~3 MPa). The enhanced tabletability is attributed to the smaller primary

crystals in QESD powder that favors stronger tablets by affording larger interparticulate

bonding area.32 The flowability of QESD powder is also significantly better as measured

by flowability index (ffc), as expected from its spherical morphology and large particle

size.9 In fact, compared to the expected minimum flowability for high speed tableting

exhibited by Avicel PH102, the flowability of the as-received FA powder is much worse,

while that of the QESD powder is significantly better. The fact that the QESD powder

already meets the minimum requirements for tabletability and flowability (Table 2.1)33, 34

suggests suitability for developing a higher loading tablet formulation that will still meet

the design criteria.

2.4.4 Formulation development

Compacting the QESD powder alone at 60 MPa resulted in tablets that meet the

tensile strength requirement of 2 MPa (Figure 2.6a). However, the disintegration time (>30

min) is longer than the requirement of < 15 min for an immediate release tablet (Table

2.1).35 To determine the amount of a disintegrant, crospovidone, that sufficiently shortens

disintegration time to 15 min, different QESD - crospovidone mixtures were compressed

at 60 MPa to obtain tablets for disintegration time determination. Addition of 0.75 %

crospovidone was sufficient to lower the disintegartion time to about 10 min (Figure 2.7a).

However, when neat QESD powder was compressed, the tablet ejection force was too high

(Figure 2.7b). Therefore, 0.25 % magnesium stearate was added to reduce tablet ejection

force. Consequtnly, ejection forces below 400 N over the entire range of pressure were

Page 63: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

41

observed (Figure 2.7b). Thus, 0.25% magnesium stearate was adequate for lubricating the

formulation (Table 2.1). At this point, a formulation containing 99% of FA satisfies the

requirements of short disintegration time and low ejection force as well.

The suitability of the 99% FA loading formulation for manufacturing under realistic

conditions was further examined by assessing tabletability, flowability, friability, and

dissolution. Tablets with tensile strength of 2 and 7 MPa could be made at 60 and 250 MPa,

respectively (Figure 2.8a), which confirms the excellent tabletability of this formulation.

The much higher ffc compared to Avicel PH102 indicates excellent flowability (Figure

2.8b). At a speed corresponding to 49,300 tablets/h (25 ms dwell time), the relative

standard deviation of tablet weight was 2.1% (average tablet weight was 301.8 mg) with

no tablet outside 7.5% of the average weight. This meets the requirement for acceptable

tablet weight variation (Table 2.1).36 Using an expedited method,37 the friability profile of

formulated QESD tablets revealed that a compaction pressure of 30 MPa was sufficient to

guarantee less than 0.8 % weight loss (Figure 2.8c). Tablet friability was about 0.4% at 60

MPa, which is well within the requirement for friability (Table 2.1).38 Finally, the

dissolution performance of FA QESD formulations in a pH 6.8 phosphate buffer is

essentially the same as a control tablet containing 99% as-received FA.More than 80% of

FA was released within 1 h (Figure 2.8d), which meets the USP dissolution requirement

for botanical dosage forms (75% release in 60 min).39 Thus, the improved

manufacturability afforded by QESD process does not sacrifice dissolution performance of

FA. Not surprisingly, this formulated QESD FA exhibited both superior tabletability and

flowability over a formulation containing 99% as-received FA (Figure 2.9a and 2.9b). In

the event that challenges are encountered during manufacturing at the commercial scale,

Page 64: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

42

such as punch sticking,40, 41 there is plenty of formulation space to appropriately address

them while still keeping a very high loading of FA.

2.5 Conclusion

In summary, by using spherical crystallization based on the quasi-emulsion solvent

diffusion method, a DC tablet formulation containing 99% of the API, ferulic acid was

developed. The resulting tablet met all critical quality attributes. This work highlights the

potential for spherical crystallization to significantly increase drug loading in DC tablet

formulations, when needed.

Page 65: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

43

Table 2.1. Powder and tablet design criteria.

Properties Requirement reference

Tablet weight

variation

Out of 20 tablets, ≤ 2 tablets differ from the average weight

by more than 7.5%, and no tablet differs in weight by more

than 15%

36

Tablet friability For a batch of tablets (unit weight less than 650 mg)

weighing approximately 6.5 g, mean weight loss ≤ 1.0%

38

Tensile strength ≥ 2 MPa 33

Flowability Better than Avicel PH102 34

Tablet disintegration 15 min, unless otherwise specified in the individual

monograph a

35

Dissolution Varies

Ejection force ≤ 600 N b

a For uncoated tablets except soluble tablets, dispersible tablets, effervescent tablets and tablets for

use in the mouth) b Based on experience

Page 66: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

44

Figure 2.1. Flowability of mixtures between lactose 316 and as-received FA powder. The

flowability of Avicel PH102 is used as the minimum flowability for high speed tableting.

Page 67: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

45

Figure 2.2. TGA profile of QESD FA.

Page 68: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

46

Figure 2.3. PXRD patterns of FA, pure QESD, as-received, and calculated

Page 69: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

47

Figure 2.4. Particle size distribution of QESD and as-received FA powders.

Page 70: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

48

Figure 2.5. SEM images of FA particles: (a) QESD powder (low magnification), (b) QESD

powder (high magnification), (c) as-received powder (low magnification), and (d) as-

received powder (high magnification).

Page 71: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

49

Figure 2.6. Manufacturability of pure as-received and QESD FA powders (a) tabletability

and (b) flowability.

a) b)

Page 72: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

50

Figure 2.7. (a) Tablet disintegration time as a function of crospovidone concentration in

mixtures with QESD FA (at 60 MPa compaction pressure), (b) tablet ejection force profiles

with and without 0.25% magnesium stearate.

a) b)

Page 73: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

51

Figure 2.8. Properties of QESD based FA formulation (a) tabletability, (b) flowability, (c)

friability, and (d) tablet dissolution in 6.8 pH buffer (in comparison to a formulation using

as-received FA).

a) b)

c) d)

Page 74: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

52

Figure 2.9. Manufacturability of formulated as-received and QESD FA powders (99%

loading) (a) tabletability and (b) flowability.

a) b)

Page 75: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

53

2.6 References

1. Jivraj, M.; Martini, L. G.; Thomson, C. M. An overview of the different excipients useful

for the direct compression of tablets. Pharm Sci Technolo Today 2000, 3, 58-63.

2. Hirschberg, C.; Sun, C. C.; Rantanen, J. Analytical method development for powder

characterization: Visualization of the critical drug loading affecting the processability of a

formulation for direct compression. J. Pharm. Biomed. Anal. 2016, 128, 462-468.

3. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine active

pharmaceutical ingredient powder blends and tablets at high drug loading via dry particle coating.

Int. J. Pharm. 2018, 543, 288-299.

4. Capece, M.; Huang, Z.; Davé, R. Insight into a novel strategy for the design of tablet

formulations intended for direct compression. J. Pharm. Sci. 2017, 106, 1608-1617.

5. Martinello, T.; Kaneko, T. M.; Velasco, M. V. R.; Taqueda, M. E. S.; Consiglieri, V. O.

Optimization of poorly compactable drug tablets manufactured by direct compression using the

mixture experimental design. Int. J. Pharm. 2006, 322, 87-95.

6. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high drug

loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

7. Thoorens, G.; Krier, F.; Leclercq, B.; Carlin, B.; Evrard, B. Microcrystalline cellulose, a

direct compression binder in a quality by design environment—a review. Int. J. Pharm. 2014, 473,

64-72.

8. Chattoraj, S.; Sun, C. C. Crystal and particle engineering strategies for improving powder

compression and flow properties to enable continuous tablet manufacturing by direct compression.

J. Pharm. Sci. 2018, 107, 968-974.

9. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow properties

using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

10. Geldart, D. Types of gas fluidization. Powder Technol. 1973, 7, 285-292.

11. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization of

aceclofenac and paracetamol spherical crystals and their tableting properties. Powder Technol.

2015, 274, 446-454.

12. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm 2016,

18, 1489-1494.

13. Fadke, J.; Desai, J.; Thakkar, H. Formulation development of spherical crystal

agglomerates of itraconazole for preparation of directly compressible tablets with enhanced

bioavailability. AAPS PharmSciTech 2015, 16, 1434-1444.

14. Yoshiaki Kawashima, M. O. Spherical crystallization direct spherical agglomeration of

salicylic acid crystals during crystallization. Science 1982, 216, 1127-1128.

15. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose with

enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

16. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration of

platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

17. Kovačič, B.; Vrečer, F.; Planinšek, O. Spherical crystallization of drugs. Acta Pharm 2012,

62, 1-14.

Page 76: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

54

18. Leon, R. A. L.; Wan, W. Y.; Badruddoza, A. Z. M.; Hatton, T. A.; Khan, S. A.

Simultaneous spherical crystallization and co-formulation of drug(s) and excipient from

microfluidic double emulsions. Cryst. Growth Des. 2014, 14, 140-146.

19. Yoshiaki Kawashima, T. N. Preparation of controlled-release microspheres of ibuprofen

with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J. Pharm. Sci. 1988,

78, 68-72.

20. Sgarbossa, A.; Giacomazza, D.; Di Carlo, M. Ferulic acid: A hope for alzheimer’s disease

therapy from plants. Nutrients 2015, 7, 5764-5782.

21. Janicke, B.; Hegardt, C.; Krogh, M.; Onning, G.; Akesson, B.; Cirenajwis, H. M.; Oredsson,

S. M. The antiproliferative effect of dietary fiber phenolic compounds ferulic acid and p-coumaric

acid on the cell cycle of caco-2 cells. Nutr. Cancer 2011, 63, 611-22.

22. Pagidipati, N. J.; Gaziano, T. A. Estimating deaths from cardiovascular disease: A review

of global methodologies of mortality measurement. Circulation 2013, 127, 749-56.

23. Prabhakar, P. K.; Prasad, R.; Ali, S.; Doble, M. Synergistic interaction of ferulic acid with

commercial hypoglycemic drugs in streptozotocin induced diabetic rats. Phytomedicine 2013, 20,

488-494.

24. Saija, A.; Tomaino, A.; Trombetta, D.; De Pasquale, A.; Uccella, N.; Barbuzzi, T.; Paolino,

D.; Bonina, F. In vitro and in vivo evaluation of caffeic and ferulic acids as topical photoprotective

agents. Int. J. Pharm. 2000, 199, 39-47.

25. Sun, C. C. Materials science tetrahedron--a useful tool for pharmaceutical research and

development. J. Pharm. Sci. 2009, 98, 1671-1687.

26. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-compression

test. J. Pharm. Sci. 1970, 59, 688-691.

27. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet friability

method. Int J Pharm 2015, 484, 146-55.

28. Paul, S.; Chang, S.-Y.; Dun, J.; Sun, W.-J.; Wang, K.; Tajarobi, P.; Boissier, C.; Sun, C.

C. Comparative analyses of flow and compaction properties of diverse mannitol and lactose grades.

Int. J. Pharm. 2018, 546, 39-49.

29. Kawashima, K. M. a. Y. Micromeritic characteristics and agglomeration mechanisms in

the spherical crystallization of bucillamine by the spherical agglomeration and the emulsion solvent

diffusion methods. Powder Technol. 1993, 76, 57-64.

30. Cui, F.; Yang, M.; Jiang, Y.; Cun, D.; Lin, W.; Fan, Y.; Kawashima, Y. Design of

sustained-release nitrendipine microspheres having solid dispersion structure by quasi-emulsion

solvent diffusion method. J. Control. Release 2003, 91, 375-384.

31. Kumar, N.; Pruthi, V. Structural elucidation and molecular docking of ferulic acid from

parthenium hysterophorus possessing cox-2 inhibition activity. 3 Biotech 2015, 5, 541-551.

32. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity. J

Adhes Sci Technol. 2011, 25, 483-499.

33. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

34. Sun, C. C. Setting the bar for powder flow properties in successful high speed tableting.

Powder Technol. 2010, 201, 106-108.

Page 77: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

55

35. WHO. Revision of monograph of tablets. 2011, Document QAS/09.324/Final.

36. USP. Usp41(online), general chapters. Weight variation of dietary supplements-tablet.

Rockville: United states pharmacopoeial convention. 2018.

37. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet friability

method. Int. J. Pharm. 2015, 484, 146-155.

38. USP. Usp41(online), general chapters. Tablet friability. Rockville: United states

pharmacopoeial convention. 2018.

39. USP. Usp34 (online), general chapter. Disintegration and dissolution of dietary

supplements. Rockville: United states pharmacopoeial convention. 2011.

40. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.; Meenan,

P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J. Pharm. Sci. 2017,

106, 151-158.

41. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through excipient

selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

Page 78: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

56

Chapter 3

Profoundly improved plasticity and tabletability of griseofulvin

by in-situ solvation and desolvation during spherical

crystallization

Page 79: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

57

3.1 Synopsis

Griseofulvin (GSF) is a high dose drug exhibiting poor flowability and tabletability,

which makes formulating a high drug loading tablet challenging. In using spherical

crystallization to improve flowability, the spherical agglomerates (SA) of GSF were found

to have profoundly improved tabletability over the as-received GSF. An improved

plasticity of GSF SA was observed despite its identical crystal structure and larger particle

size when compared to the as-received GSF. Tracking solid forms during the process

revealed the formation of a GSF solvate with dichloromethane, the bridging liquid for

spherical agglomeration. With subsequent desolvation during drying, nanoporous GSF

crystals were obtained, which could be plastically deformed more easily and exhibited

much improved tabletability.

3.2 Introduction

Successful tablet manufacturing requires the powder blends to meet several criteria,

including adequate tabletability, flowability, uniformity, and free from sticking to

punches.1, 2 Particle engineering and crystal engineering are two effective approaches to

overcome inherent problematic physical or mechanical properties and, thereby, enable

successful tablet development.3-5 Common methods to improve powder flowability and

tabletability of active pharmaceutical ingredients (API) have recently been summarized.6

Methods to improve flowability include making spherical granules,7, 8 increasing size,3

reducing surface roughness,9 reducing cohesion by nano-coating,10 and changing surface

chemistry.11 Methods to improve tabletability, such as surface polymer coating,12 size

reduction,13 increasing roughness,14 cocrystallization,15 and crystal hydration,16 mainly rely

Page 80: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

58

on increasing bonding area among particles.17 Typically, flowability enhancement by size

enlargement and surface smoothing by granulation also leads to deteriorated tabletability.18

In comparison to common granulation methods, spherical crystallization (SC)

stands out as an efficient way to simultaneously improve flowability and tabletability.19-21

SC is a method that transforms fine crystals into compact spherical agglomerates (SA).22

Such spherical crystal agglomerates exhibit good flowability because of the large size and

spherical shape.8, 23, 24 While a systematic investigation remains incomplete, the improved

tabletability by SC, is commonly attributed to the enhanced interparticle contacts due to

fragmentation of agglomerates,21, 25 smaller primary crystals19 and easier plastic

deformation.26 Spherical agglomerates are commonly produced by either spherical

agglomeration22 or quasi-emulsion solvent diffusion (QESD)27. In the spherical

agglomeration, three solvents, good solvent, poor solvent, and bridging liquid, are used.

During this process, the API is dissolved in a good solvent, followed by adding the drug

solution to a miscible poor solvent to induce immediate precipitation of fine API crystals.

Then, a bridging liquid is added to facilitate agglomeration of primary fine API crystals

into large spherical agglomerates through the surface tension effect.28 The polarity of

solvents is important in the SA process. The good and poor solvents should have similar

polarity and they are miscible. When they mix through molecular diffusion, API

precipitates out because of the reduced interactions with the good solvent due to the affinity

between the poor and good solvent molecules. The faster the mixing between good and

poor solvents, the faster the crystallization and finer primary crystals, which will

subsequently form agglomerates by the bridging liquid. In general, the polarity of the

bridging liquid should be different from the poor solvent, and hence they are immiscible.

Page 81: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

59

This ensures that the bridging liquid is available for collecting small API crystals into

agglomerates.29

Griseofulvin (GSF), an antifungal drug with dose as large as 1 g, is marketed as

high drug loading tablets (up to 500 mg per tablet). Pure GSF shows poor powder

flowability and tabletability. In an effort to improve flowability of GSF by spherical

agglomeration to enable the development of a high dose tablet,1 we surprisingly found

simultaneously and profoundly improved tabletability; the mechanism of which was

investigated.

3.3 Materials and methods

3.3.1 Materials

Griseofulvin was obtained from GSK Glaxo laboratories, London, UK.

Microcrystalline cellulose (Avicel PH102; FMC Biopolymers, Newark, DE),

dimethylformamide (DMF, Sigma-Aldrich, St. Louis, MO), dichloromethane (DCM,

Fisher Scientific, Fair Lawn, NJ), ultrapure deionized water (0.066 μS/cm, Thermo

Scientific, USA) were used as received. All crystallization experiments were performed

using glass beakers.

3.3.2 Spherical crystallization

3.3.2.1 Construction of ternary phase diagram

DMF, water, and DCM were used as the good solvent, poor solvent and bridging

liquid for preparing spherical GSF crystals using the spherical agglomeration method from

Page 82: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

60

a preliminary screening study. A ternary phase diagram was constructed to identify the

agglomeration zone by the following process.30 DMF and DCM were cooled to 0 oC in an

ice-bath and then combined in volume ratios of 1:9, 2:8, 3:7, 4:6, 5:5, 4:6, 7:3, 8:2, and 9:1.

A 5 mL aliquot of each was transferred to a screw capped vial, and water was added

dropwise with intermittent mixing. When the clear mixed solvents became turbid, the

volume of water was noted. To obtain the agglomeration zone on the phase diagram, 1 g

of GSF was dissolved in 7 ml of DMF. The DMF solution was poured into water with

DMF:water volume ratios of 7:17.5, 7:30, 7:40, 7:60 and 7:80, where the API precipitated.

Then, DCM was added dropwise until a paste-like product was observed, and the volume

of DCM was noted. A ternary phase diagram, including agglomeration zone, miscible and

immiscible regions, was then constructed.

3.3.2.2 Preparation of spherical agglomerates of GSF

To prepare spherical agglomerates of GSF, 9 g of GSF was dissolved in 60 mL of

DMF at 120 oC to obtain a clear solution. The solution was poured into 270 mL of water

in a 500 mL glass beaker, which was immersed in an ice bath and agitated with an overhead

stirrer at 600 rpm. After 5 min agitation, DCM was added dropwise until fine crystals were

transformed into large agglomerates. The system was agitated for another hour, and the

agglomerates were collected and dried in a 60 oC oven for 4 hours.

3.3.3 Powder X-ray Diffractometry (PXRD)

A powder X-ray diffractometer (PANalytical X'pert pro, Westborough, MA) with

Cu Kα radiation (1.54059 Å) was used to characterize powders. Samples were scanned

Page 83: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

61

between 5 to 35° 2θ with a step size of 0.02° and a dwell time of 1 s/step. The tube voltage

and amperage were set at 40 kV and 40 mA, respectively. The diffraction patterns of the

powders were compared to the original powder and the calculated PXRD pattern from the

single crystal structure of GSF (CCDC refcode: GRISFL06).31

3.3.4 Single Crystal X-ray Diffractometry

Single crystal X-ray diffraction experiment was carried out on a Bruker-AXS

Venture Photon-II diffractometer (Bruker AXS Inc., Madison, Wisconsin) equipped with

a Photon-II (CMOS) detector. The data collection was performed using Mo Kα radiation

at 100 K. A suite of software from Bruker, including APEX3, SADABS, SAINT and

XPREP was used to analyze the collected data. The crystal structure was solved and refined

using Bruker ShelXle. Direct-methods solution was applied to place non-hydrogen atoms

from the E-map. All non-hydrogen atoms were refined with anisotropic displacement

parameters. Hydrogen atoms that involved in hydrogen bonding were located from the

residual peaks in the Fourier map, otherwise were generated geometrically.

3.3.5 Thermal Analyses

Powder samples (3−5 mg) were loaded into hermetic aluminum pans and heated

from 30 to 240 °C at a heating rate of 10 °C/min on a differential scanning calorimeter

(Q1000, TA Instruments, New Castle, DE) under a continuous nitrogen purge at a flow

rate of 50 mL/min. DSC cell parameter was calibrated with indium for heat flow and

indium and cyclohexane for temperature. To measure any volatile content in a solid, a

Page 84: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

62

thermogravimetry analyzer (Q500, TA Instruments, New Castle, DE) was applied to

analyze the samples (3∼10 mg) placed in an open aluminum pan. The samples were heated

from room temperature to 300 °C at 10 °C/min under 25 mL/ min nitrogen purge.

3.3.6 Powder Flowability

Powder flowability was measured using a ring shear cell tester (RST-XS; Dietmar

Schulze, Wolfenbüttel, Germany), with a 10 mL cell at 1 and 3 kPa pre-shear normal

stresses. The unconfined yield strength (fc) and major principal stress (σn) were obtained

from each yield locus by drawing Mohr’s circles. Flowability index (ffc) was calculated

using Eq. (1):

𝑓𝑓c =𝜎𝑛

𝑓c (1)

3.3.7 Powder tabletability

A universal material test machine (model 1485, Zwick/Roell, Germany) was used

to prepare tablets with compaction pressure ranging from 25 to 350 MPa at speed of 4

mm/min. The punch tip and die wall was lubricated using magnesium stearate. Tablets

were relaxed under ambient environment for at least 24 h before being broken diametrically

using a texture analyzer (TA-XT2i; Texture Technologies Corporation, Scarsdale, NY).

Tablet tensile strength was calculated from the breaking force and tablet dimensions

following the standard procedure.32

True density of powders was measured using a helium pycnometer (Quantachrome

Instruments, Ultrapycnometer 1000e, Byonton Beach, FL). The sample cell was filled with

accurately weighed powders (1-2g). The measurement was stopped when the coefficient

Page 85: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

63

of volume variation of the last five consecutive measurements was below 0.005%.

Otherwise, the experiment terminates automatically at a maximum of 100 measurements.

The mean and standard deviation of the last five measurements were reported.

3.3.8 Scanning electron microscopy (SEM)

An ion-beams sputter (IBS/TM200S; VCR Group Inc., San Clemente, CA) was

used to was sputter-coated a thin layer platinum (thickness ~ 75 Å) onto the samples.

Scanning electron microscopy (JEOL 6500F; JEOL Ltd., Tokyo, Japan) operated at SEI

mode with an acceleration voltage of 5kV was used to evaluate particle morphology and

surface feathers and a high vacuum (10-4-10-5 Pa) was maintained during the imaging

process.

3.3.9 In-die Heckel analysis

A universal material test machine (model 1485, Zwick/Roell, Germany) was used

to prepare tablets (about 200mg) with compaction pressure 300 MPa at speed of 4 mm/min

(n=3). The force and normal strain during the compression are exported. Both GSF SA and

as-received were analyzed using the Heckel equation,33 Eq. (2):

− ln(1 − 𝐷) = 𝐾𝑃 + 𝐴 (2)

where P is the compaction pressure, K is the slope of the linear portion of the Heckel plot,

and A is the y-axis intercept of the linear portion. D is the relative density calculated using

Eq. (3)

Page 86: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

64

D =tablet density

true density (3)

The reciprocal of K, which is termed mean yield pressure (Py), is a parameter that can be

used to quantify plasticity of the powder. A lower Py indicates higher plasticity of a powder.

The measurement was triplicated.

3.3.10 Out-of-die Kuentz–Leuenberger (KL) equation

The Kuentz–Leuenberger (KL) equation, Eq. (4)34 was applied to obtain the

plasticity parameter, 1/C, by fitting tablet porosity – pressure data. A lower 1/C value

indicates higher plasticity of the powder. The parameter, εc, describes the critical powder

porosity at which the powder starts to gain rigidity or strength.

𝑃 =1

𝐶[𝜀 − 𝜀𝑐 − 𝜀𝑐𝑙𝑛(

𝜀

𝜀𝑐) (4)

Out-of-die Hardness determination by macroindentation

A spherical stainless indenter (3.18 mm in diameter), attached to a texture analyzer

(TA-XT2i, Texture Technologies Corp., NY), was used to indent the cylindrical tablet at

the center of a flat face. During loading, the indenter was moved downward at a speed of

0.01 mm/min, and the indenting force was set around 60% of the breaking force of the

tablet prepared at the same compaction pressure and maintained for 3 min. The indented

area was measured using a calibrated digital microscope (×200 magnification, Dino-Lite

Pro AM413MT; AnMo Electronics Corporation, Taiwan). A three-point method was used

to identify the circular indented area and the projected area, A, was calculated using

DinoCapture software (V2.0 AnMo Electronics Corporation, Taiwan). To more accurately

Page 87: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

65

determine the indented area, the contrast between the indent and surrounding flat tablet

surface was enhanced by gently rubbing the tablet face against a piece of graphite-coated

paper. Tablet hardness, H, was calculated using Eq. (5):

𝐻 =𝐹

𝐴 (5)

H0, hardness at zero porosity, was obtained by extrapolating the function to zero porosity.

3.3.11 Energy framework of crystal structures

The pairwise intermolecular interaction energy was estimated using CrystalExplorer and

Gaussian09W with experimental geometry of GSF Form I.35-38 The hydrogen positions

were normalized to standard neutron diffraction values before calculation using the

B3LYP-D2/6-31G(d,p) electron densities model. The total intermolecular interaction

energy for a given molecule is the sum of the electrostatic, polarization, dispersion, and

exchange-repulsion components with scale factors of 1.057, 0.740, 0.871, and 0.618,

respectively. Intermolecular interaction between two molecules was ignored, when the

closest atom - atom distance was more than 3.8 Å. The interaction energies were

graphically presented as a framework by connecting centers of mass of molecules with

cylinder thickness proportional to the total intermolecular interaction energies, where

interaction energies below -5 kJ/mol were omitted for clarity.

3.4 Results and discussion

3.4.1 Ternary Phase Diagram

Page 88: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

66

The ternary phase diagram (Figure 3.1) pertaining to SA of GSF in this work

revealed a distinct agglomeration zone for producing spherical agglomerates (shaded

region). This zone is in the immiscible region neighboring the miscible zone, where free

bridging liquid (DCM) was available to collect small primary GSF crystals into large SA.

No agglomerates were obtained in the miscible zone. Outside the shaded area in the

immiscible zone, a paste-like product was formed because of excess amount of bridging

liquid.

3.4.2 Particulate properties

The microscopic images and particle size analysis revealed the particle shape and

size after the spherical agglomeration process. The as-received GSF consisted of irregular

crystals with size ranging 1 - 10 µm, while the agglomerates were round with size ranging

300 - 600 µm (Figure 3.2). Moreover, no weight loss was detected for SA before 153 oC

(boiling point of DMF) by thermal gravimetric analysis (Figure 3.3), indicating negligible

residual solvent.

3.4.3 Powder flowability and tabletability

The flowability of the powder is an important powder property in tablet

manufacturing. At normal stresses of 1 and 3 kPa, the ffC values of GSF SA were similar

to those of Avicel PH102 (Figure 3.4a), which suggests their adequate flowability for high

speed tableting.39 The flow property of the as-received GSF, on the other hand, was much

Page 89: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

67

poorer than that of Avicel PH102 (Figure 3.4a). The much improved flowability of the GSF

SA is a consequence of the large size and round shape.23

Similar to a previous study,40 intact tablets of as-received GSF could not be

obtained, due to lamination, over the entire pressure range between 25 and 350 MPa. The

problematic tabletability of GSF is attributed to minimal plasticity due to a lack of active

slip planes in its crystal structure.40 Surprisingly, the GSF SA exhibited excellent

tabletability with the highest tensile strength around 5 MPa at ~300 MPa (Figure 3.4b).

The profoundly improved tabletability could not be explained by any changes in solid state

form, because the GSF SA were phase pure GSF Form I, the same as the as-received GSF.

Another commonly cited mechanism to explain improved tabletability of spherical

agglomerates is extensive fracture of agglomerates during compaction,21, 25 which exposes

fresh surfaces for developing a network of interparticle bonding. This mechanism is

reasonable for lubricated powders where coating of particles by poorly bonding lubricant

weakens bonding strength. In that case, generation of particle surfaces free from such

deleterious effect by fragmentation improves tablet tensile strength through higher bonding

strength according to the bonding area – bonding strength model.17, 18, 41 However, this

explanation is not appropriate, because GSF powders were not mixed with any lubricant.

The external lubrication of tooling could not have affected interparticulate bonding strength.

In addition, fragmentation of agglomerates during compaction unlikely reduced GSF to the

size of as-received GSF, which is very small (1 - 10 µm). Thus, larger bonding area due

to smaller particle also does not explain the improved tabletability. To verify this point,

GSF SA were milled and then compressed. The milled GSF SA exhibited similar

tabletability profile as the un-milled agglomerates (Figure 3.4b). Hence, in-situ particle

Page 90: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

68

fragmentation can be excluded as a main mechanism to the profoundly increased

tabletability. Therefore, a different mechanism must have caused significantly larger

interparticulate bonding area, which led to the striking improvement in tabletability.

3.4.4 Role of plasticity in the improved tabletability of GSF agglomerates

To seek an explanation for the improved tabletability, the plasticity of the different

GSF powders was assessed by Heckel analysis of in-die compressibility data. A lower

value of the plasticity parameter, Py, from such an analysis indicates higher plasticity of

the material. Results show that the Py followed the order of: GSF SA ≈ GSF SA milled <

as-received GSF (Table 3.1), which corresponded well with the order of tabletability: GSF

SA ≈ GSF SA milled > as-received GSF. This data supports the premise that the improved

tabletability of GSF SA resulted from increased plasticity. The Py values of all three

materials are higher than that of Avicel PH102, a commonly used plastic excipient with a

Py value of ~ 62 MPa, 42 and exhibiting excellent tabletability.18 However, the higher

plasticity of agglomerated GSF cannot be attributed to a change in particle size, since the

Py of the milled GSF agglomerates is the same as that of un-milled agglomerates.

To further verify the different plasticity between the two powders using out-of-die

data, physical mixtures of GSF (80%) and Avicel PH102 (20%) were studied. Avicel P102

was used, because pure as-received GSF could not be compressed into intact tablets. The

tabletability of the mixture containing GSF SA outperformed that containing as-received

GSF (Figure 3.5a). Although the relative order in tabletability of the mixture and pure GSF

did not change, the tabletability difference between the two mixtures was much smaller

than that between pure GSF powders. The tensile strength extrapolated to zero porosity,

Page 91: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

69

which is a parameter to infer apparent bonding strength, was lower for the mixture

containing GSF SA. Thus, the better tabletability of the GSF SA mixture was not due to

higher bonding strength between the particles (Figure 3.5b). However, at the same

compaction pressure, tablets of the mixture containing GSF SA exhibited significantly

lower porosity (Figure 3.5c), indicating higher plasticity. In fact, the plasticity parameter,

1/C, from the KL analysis of the mixture containing GSF SA was lower than the mixture

containing the as-received GSF (Table 3.1), confirming the higher plasticity of SA.

The hardness at zero porosity, H0, is also a measure of plasticity of the powder,

where a lower H0 value relates to a higher plasticity.43 By this measure, the mixture

containing GSF SA was also more plastic because of its lower H0 (Figure 3.5d, Table 3.1).

Therefore, results from all complementary methods confirm the higher plasticity of the

GSF SA. The higher plasticity, in turn, leads to larger bonding area and higher tabletability

for the mixture containing GSF SA.

3.4.5 Origin of superior plasticity of GSF SA

At this point, the underlying reason for the superior plasticity remained elusive.

Because GSF SA had a crystal structure identical to the as-received GSF, as shown by the

nearly identical PXRD patterns (Figure 3.6a) and DSC thermograms (Figure 3.6b).

However, scanning electronic microscope (SEM) images revealed intriguing

difference in the structures of the two powders, where many nanopores with opening size

of ~500 nm were clearly observed in GSF SA but absent in the as-received GSF (Figures

Page 92: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

70

3.7 and 3.8). These nanopores can collapse during compression to make particles more

deformable, which leads to larger interparticulate contact area in the tablet.

The collapse of nanopores after compaction was verified by observing GSF

agglomerate crystals exposed to the fracture surface of a tablet prepared at 250 MPa

(mixture containing 20% Avicel PH102) after breaking diametrically (Figure 3.7). It is

interesting to note that small nanosized particles also appeared on the tablet fracture surface

of the GSF SA (Figure 3.7). These nanosized particles indicate possible fragmentation of

the porous GSF crystals, either during compaction or during tablet breaking. In contrast,

no obvious changes to the as-received GSF crystals were observed (Figure 3.7). Thus, the

as-received GSF crystals undergo largely reversible, elastic deformation during

compaction, which correlates to its extremely poor tabletability (Figure 3.4b).

The origin of the nanopores in GSF SA is attributed to the solvation and desolvation

process during the spherical agglomeration process. The formation of a DCM solvate is

confirmed by the PXRD patterns of the fresh prepared GSF SA, which matched the

calculated pattern of GSF-DCM solvate (Figure 3.6a).44 While the GSF molecules are

closely packed in GSF Form I (Figure 3.9a), they form a hexagonal reticular structure

fortified through a series of C-H...O (3.347, 3.348, 3.484, 3.476, 3.478 Å) and C-H...π (3.664

Å) interactions. The voids in the solvate are filled by DCM molecules, which interact with

GSF through C-H...O (3.284, 3.361, 3.364 Å) weak hydrogen bonds (Figure 3.9b). The

weak interaction strength between DCM and GSF is consistent with the observed relatively

facile desolvation process, which goes to completion at 60 oC in 4 h (Table 3.2). The open

structure of the desolvated GSF-DCM crystal, with 18.6% of void by volume, is unstable

(Figure 3.9c), as suggested by the calculated large lattice energy difference (~20 kcal/mol)

Page 93: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

71

between GSF Form I (-94.24 kcal/mol) and GSF anhydrous (-73.98 kcal/mol). Driven by

such a large energy difference, fast spontaneous collapse of the desolvated structure to

Form I is expected, which induced cracks and generated nanopores in GSF crystals. A

similar mechanism was previously reported for acetaminophen, which also led to

significantly improved tabletability.45-47 Although we only studied DCM solvate of GSF in

this work, it is likely that the solvation -desolvation of other GSF solvates, such as GSF-

chloroform, GSF-benzene, GSF dioxane,44, 48-50 can achieve a similar effect. This is being

investigated in our laboratory.

3.5 Conclusion

The spherical GSF Form I agglomerates, prepared by the spherical crystallization

process, exhibited not only good flowability as expected but also surprisingly excellent

tabletability without a change in the solid form. The latter is attributed to the in-situ

formation of a DCM solvate and subsequent desolvation during drying, which generates

nanopores and makes the GSF crystals more deformable by collapse of pores during

compaction. Such superior deformability leads to large interparticular bonding area and

stronger tablets. This mechanism for enhancing crystal plasticity may be useful to improve

tabletability of other poorly compressible APIs that can form a solvate or hydrate.

Page 94: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

72

Figure 3.1. Ternary phase diagram of GSF in water-DMF-DCM solvents system.

Page 95: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

73

Figure 3.2. Microscopic images of GSF (a) as-received and (b) SA.

(b) (a)

Page 96: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

74

Figure 3.3. TGA profiles of GSF SA and as-received GSF.

Page 97: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

75

Figure 3.4. (a) Flowability profiles of as-received GSF, GSF SA and Avicel PH102 (n=1);

(b) tabletability profiles of as-received GSF, GSF SA and GSF SA milled (n=3).

a) b)

Page 98: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

76

Table 3.1. Plasticity parameters of as received GSF and GSF SA from in-die and out-of-

die data analyses.

Sample name As-received GSF GSF SA GSF SA milled

Py (MPa), n=3 143.0 ±5.3 107.9 ±1.3 106.4 ±2.3

1/C (MPa) a 450.0 ± 208.5 (R2=0.98) 143.4 ± 38.4 (R2=0.99) --

H0 (MPa) a 435 ± 21 (R2=0.98) 333 ± 18 (R2=0.98) --

a Parameters were derived for mixtures containing 20% Avicel PH102

Page 99: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

77

Figure 3.5. (a) Tabletability, (b) compactibility, (c) compressibility, and (d) hardness

profiles of GSF (80%) and Avicel PH102 (20%) mixture.

c) d)

a) b)

Page 100: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

78

Figure 3.6. (a) PXRD patterns of calculated GSF Form I, as-received GSF, GSF SA,

calculated GSF-DCM and GSF SA before drying; (b) DSC profiles of as-received GSF and

GSF SA.

b

) a)

Page 101: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

79

Figure 3.7. SEM images at (x10,000 magnification) of GSF (a) as-received crystals, (b)

SA, and tablet fracture surface of (c) GSF-Avicel mixture, and (d) GSF SA-Avicel mixture.

(a) (b)

(c) (d)

Page 102: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

80

Figure 3.8. SEM image at (x20,000 magnification) of GSF SA.

Page 103: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

81

Figure 3.9. Energy framework of (a) GSF Form I and (b) GSF-DCM solvate. The (c)

desolvated GSF-DCM with large void is also shown for comparison. The energy threshold

for the energy framework is set at −22 kJ/mol.

a) b)

c)

Page 104: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

82

Table 3.2. Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p)

dispersion corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)),

polarization (E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the

energy are listed. R indicated the distance between centers of mass of the pair of molecules.

Symop R E_ele E_pol E_dis E_rep E_tot

- 6.41 -9.4 -1.3 -42.9 28.8 -30.5

x, y, z 8.50 -1.7 -4.2 -16.3 19.5 -7.0

x, y, z 13.93 -6.6 -1.8 -3.1 1.5 -10.0

- 8.02 -18.0 -4.7 -46.9 37.4 -40.3

- 7.08 -1.4 -0.9 -10.2 5.7 -7.5

- 4.73 -6.9 -1.3 -19.7 21.3 -12.3

- 8.08 -16.9 -4.6 -30.1 18.3 -36.2

- 7.37 -1.4 -0.1 -6.4 4.7 -4.2

x, y, z 11.67 -11.9 -4.4 -10.6 8.8 -19.5

- 8.53 -15.4 -4.0 -34.0 20.1 -36.4

- 8.93 0.7 -0.4 -3.8 2.7 -1.2

- 11.96 -0.9 -2.3 -6.1 2.3 -6.5

- 8.34 -12.4 -4.0 -8.2 12.2 -15.7

- 5.89 -8.4 -2.9 -14.4 11.8 -16.2

- 7.70 -20.0 -6.2 -37.8 30.3 -40.0

- 6.93 -2.5 -0.8 -7.4 7.9 -4.8

- 7.37 -1.4 -0.1 -6.4 4.7 -4.2

- 6.93 -2.5 -0.8 -7.4 7.9 -4.8

- 7.79 -8.7 -1.2 -7.7 4.9 -13.9

- 6.14 -2.0 -0.6 -10.0 4.5 -8.6

- 9.19 3.4 -0.6 -3.0 0.8 1.0

- 8.72 -2.9 -0.4 -4.6 3.1 -5.5

- 7.08 -1.4 -0.9 -10.2 5.7 -7.5

- 8.93 0.7 -0.4 -3.8 2.7 -1.2

- 8.34 -12.4 -4.0 -8.2 12.2 -15.7

Page 105: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

83

- 6.86 0.9 -0.6 -4.6 1.8 -2.4

- 6.00 -3.4 -0.1 -3.1 11.1 0.5

- 5.72 -2.5 -0.2 -4.7 6.7 -2.8

x, y, z 8.50 1.0 -3.8 -15.9 16.4 -5.5

x, y, z 11.67 -18.0 -6.7 -16.7 17.7 -27.6

- 6.86 0.9 -0.6 -4.6 1.8 -2.4

- 8.80 -1.8 -0.3 -3.6 1.9 -4.1

- 6.14 -2.0 -0.6 -10.0 4.5 -8.6

- 5.20 -6.2 -0.9 -16.6 19.0 -10.0

- 8.72 -2.9 -0.4 -4.6 3.1 -5.5

- 7.79 -8.7 -1.2 -7.7 4.9 -13.9

- 5.72 -13.0 -4.0 -17.6 15.2 -22.6

- 9.19 3.4 -0.6 -3.0 0.8 1.0

- 5.72 -13.0 -4.0 -17.6 15.2 -22.6

- 5.20 -6.2 -0.9 -16.6 19.0 -10.0

- 5.89 -8.4 -2.9 -14.4 11.8 -16.2

- 4.73 -6.9 -1.3 -19.7 21.3 -12.3

- 8.80 -1.8 -0.3 -3.6 1.9 -4.1

Page 106: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

84

3.6 References

1. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

2. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through

excipient selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

3. Kristensen, H. G.; Schaefer, T. Granulation: A review on pharmaceutical wet-

granulation. Drug Dev. Ind. Pharm. 1987, 13, 803-872.

4. Rouan, E.; Rhodes, C. T. The spray drying of pharmaceuticals. Drug Dev. Ind.

Pharm. 1992, 18, 1169-1206.

5. Remenar, J. F.; Morissette, S. L.; Peterson, M. L.; Moulton, B.; MacPhee, J. M.;

Guzmán, H. R.; Almarsson, Ö. Crystal engineering of novel cocrystals of a triazole drug

with 1,4-dicarboxylic acids. J. Am. Chem. Soc. 2003, 125, 8456–8457.

6. Chattoraj, S.; Sun, C. C. Crystal and particle engineering strategies for improving

powder compression and flow properties to enable continuous tablet manufacturing by

direct compression. J. Pharm. Sci. 2017, 107, 968-974.

7. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

8. Maghsoodi, M.; Taghizadeh, O.; Martin, G. P.; Nokhodchi, A. Particle design of

naproxen-disintegrant agglomerates for direct compression by a crystallo-co-

agglomeration technique. Int. J. Pharm. 2008, 351, 45-54.

9. Geldart, D. Types of gas fluidization. Powder Technol. 1973, 7, 285-292.

10. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

11. Kim, E. H.; Chen, X. D.; Pearce, D. Effect of surface composition on the

flowability of industrial spray-dried dairy powders. Colloids Surf. B 2005, 46, 182-187.

12. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

13. Vromans, H.; Bolhuis, G. K.; Lerk, C. F.; van de Biggelaar, H.; Bosch, H. Studies

on tableting properties of lactose. Vii. The effect of variations in primary particle size and

percentage of amorphous lactose in spray dried lactose products. Int. J. Pharm. 1987, 35,

29-37.

14. Jain, S. Mechanical properties of powders for compaction and tableting: An

overview. Pharm. Sci. Technol. Today 1999, 2, 20-31.

15. Jones, W.; Motherwell, W. D. S.; Trask, A. V. Pharmaceutical cocrystals: An

emerging approach to physical property enhancement. MRS Bulletin 2011, 31, 875-879.

Page 107: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

85

16. Chang, S. Y.; Sun, C. C. Superior plasticity and tabletability of theophylline

monohydrate. Mol. Pharm. 2017, 14, 2047-2055.

17. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

18. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

19. Martino, P. D.; Cristofaro, R. D.; Barthelemy, C.; Joiris, E.; Filippo, G. P.; Sante,

M. Improved compression properties of propyphenazone spherical crystals. Int. J. Pharm.

2000, 197, 95-106.

20. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

21. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

22. Kawashima, Y.; Okumura, M.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

23. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

24. Nokhodchi, A.; Maghsoodi, M. Preparation of spherical crystal agglomerates of

naproxen containing disintegrant for direct tablet making by spherical crystallization

technique. AAPS PharmSciTech 2008, 9, 54-59.

25. Kawashima, Y.; Imai, M.; Takeuchi, H.; Yamamoto, H.; Kamiya, K.; Hino, T.

Improved flowability and compactibility of spherically agglomerated crystals of ascorbic

acid for direct tableting designed by spherical crystallization process. Powder Technol.

2003, 130, 283-289.

26. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization

of aceclofenac and paracetamol spherical crystals and their tableting properties. Powder

Technol. 2015, 274, 446-454.

27. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

28. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

29. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

30. Zhang, H.; Chen, Y.; Wang, J.; Gong, J. Investigation on the spherical

crystallization process of cefotaxime sodium. Ind. Eng. Chem. Res. 2010, 49, 1402-1411.

Page 108: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

86

31. Su, Y.; Xu, J.; Shi, Q.; Yu, L.; Cai, T. Polymorphism of griseofulvin: Concomitant

crystallization from the melt and a single crystal structure of a metastable polymorph with

anomalously large thermal expansion. Chem. Commun. 2018, 54, 358-361.

32. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

33. Heckel, W. An analysis of powder compaction phenomena. Trans. Metall. Soc.

AIME 1961, 221, 671-675.

34. Kuentz, M.; Leuenberger, H. Pressure susceptibility of polymer tablets as a critical

property: A modified heckel equation. J. Pharm. Sci. 1999, 88, 174-179.

35. Chen, Y.; Liu, C.; Chen, Z.; Su, C.; Hageman, M.; Hussain, M.; Haskell, R.;

Stefanski, K.; Qian, F. Drug–polymer–water interaction and its implication for the

dissolution performance of amorphous solid dispersions. Mol. Pharm. 2015, 12, 576-589.

36. Tao, J.; Sun, Y.; Zhang, G. G. Z.; Yu, L. J. P. R. Solubility of small-molecule

crystals in polymers: D-mannitol in pvp, indomethacin in pvp/va, and nifedipine in pvp/va.

Pharm. Res. 2009, 26, 855-864.

37. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

38. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

39. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

40. Chen, H.; Guo, Y.; Wang, C.; Dun, J.; Sun, C. C. Spherical cocrystallization - an

enabling technology for the development of high dose direct compression tablets of poorly

soluble drugs. Crys. Growth Des. 2019, 19, 2503-2510.

41. Wu, S. J.; Sun, C. Insensitivity of compaction properties of brittle granules to size

enlargement by roller compaction. J. Pharm. Sci. 2007, 96, 1445-50.

42. Paul, S.; Sun, C. C. The suitability of common compressibility equations for

characterizing plasticity of diverse powders. Int. J. Pharm. 2017, 532, 124-130.

43. Patel, S.; Sun, C. C. Macroindentation hardness measurement—modernization and

applications. Int. J. Pharm. 2016, 506, 262-267.

44. Shirotani, K.-i.; Suzuki, E.; Morita, Y.; Sekiguchi, K. Solvate formation of

griseofulvin with alkyl halide and alkyl dihalides. Chem. Pharm. Bull. 1988, 36, 4045-

4054.

45. Fachaux, J. M.; Guyot Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Huvenne, J. P.; Bouche, R. Compression ability improvement by solvation/desolvation

process: Application to paracetamol for direct compression. Int. J. Pharm. 1993, 99, 99-

107.

Page 109: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

87

46. Fachaux, J. M.; Guyot-Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Veesler, S.; Boistelle, R. Pure paracetamol for direct compression part i. Development of

sintered-like crystals of paracetamol. Powder Technol. 1995, 82, 123-128.

47. Fachaux, J. M.; Guyot-Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Veesler, S.; Boistelle, R. Pure paracetamol for direct compression part ii. Study of the

physicochemical and mechanical properties of sintered-like crystals of paracetamol.

Powder Technol. 1995, 82, 129-133.

48. Sekiguchi, K.; Suzuki, E.; Tsuda, Y.; Morita, Y. Thermal analysis of griseofulvin-

chloroform system by high pressure differential scanning calorimetry. Chem. Pharm. Bull.

1983, 31, 2139-2141.

49. Sekiguchi, K.; Horikoshi, I.; Himuro, I. Studies on the method of size reduction of

medicinal compounds. Iii. Size reduction of griseofulvin by solvation and desolvation

method using chloroform. (3). Chem. Pharm. Bull. 1968, 16, 2495-2502.

50. Sekiguchi, K.; Ito, K.; Owada, E.; Ueno, K. Studies on the method of size reduction

of medicinal compounds. Ii. Size reduction of griseofulvin by solvation and desolvation

method using chloroform (2). Chem. Pharm. Bull. 1964, 12, 1192-1197.

Page 110: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

88

Chapter 4

Spherical cocrystallization - an enabling technology for the

development of high dose direct compression tablets of poorly

soluble drugs

Page 111: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

89

4.1 Synopsis

Spherical cocrystallization (SCC), an integrated crystal and particle engineering

approach, was successfully applied to generate spherical cocrystal agglomerates (SCA)

between a poorly soluble drug, griseofulvin (GSF) and acesulfame (Acs). The SCC process

consists of two steps: 1) GSF-Acs cocrystal formation by reaction crystallization in a slurry,

2) agglomeration by adding a bridging liquid. The obtained SCA exhibited superior

mechanical and powder properties over GSF. Inclusion of a small amount (<1%) HPC in

the agglomerates improved both the tabletability and dissolution performance. Using the

SCA, a high GSF loading (55.7%) direct compression tablet formulation was successfully

developed, which exhibited much improved manufacturability and drug release property

over GSF based tablet formulation.

4.2 Introduction

The successful development of pharmaceutical tablets using the much-preferred

direct compression (DC) process needs to overcome several challenges, depending on dose

and physicochemical properties of the active pharmaceutical ingredient (API). When the

API dose is low (< 5 mg), content uniformity (CU) is a key challenge.1 When the API dose

is high (> 100 mg), manufacturability is a key challenge.2, 3 The CU challenge for low dose

drugs can be very effectively addressed using a versatile platform DC tablet formulation

employing particle engineering that involves forming a drug – carrier composite.4 However,

an effective strategy for solving the manufacturability challenge facing the development of

high dose APIs by the DC process is lacking. Additionally, APIs exhibiting solubility

Page 112: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

90

limiting bioavailability is approximately 80-90% of the new chemical entities in

pharmaceutical pipeline.5 For such APIs, solubility in gastric and intestinal fluids must be

sufficiently high to achieve adequate bioavailability if permeability does not limit API

absorption. For a given solid form of an API, a common practice to alleviate solubility

limiting bioavailability is API size reduction, which can enhance dissolution rate by the

virtue of larger API surface area.6, 7 However, smaller API particles correspond to poorer

flowability,8 which deteriorates flowability of the blend if the same API loading is used in

the tablet formulation. Hence, manufacturability problems may be encountered depending

on the dose and flowability of the size-reduced API powder. One way to address such flow

related manufacturability problems is to use lower API loading by using a larger tablet.

However, larger tablets are generally undesirable because of the lower patient compliance.9

An alternative approach to improve API bioavailability is the use of soluble solid

form, such as amorphous solid, cocrystal, and salts.10 Among these solid forms, cocrystal

is attractive because it exhibits better physical stability than amorphous solids during

storage and it is applicable to non-ionizable APIs that are not amenable to salt formation.11

However, the presence of coformer in cocrystal necessarily increases the loading in the

formulation for the same API amount. This presents an additional challenge for high dose

APIs if the more soluble cocrystal does not exhibit desired flow or tableting properties,

which likely dominate manufacturability of the blend.2, 12

Spherical crystallization is a process that directly transforms fine crystals into

compact spherical agglomerates during the crystallization process.13-16 For a given crystal

form, spherical crystallization has been found effective in improving powder flowability,

packability, compressibility, and tabletability on study such as lactose,17 albendazole,14

Page 113: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

91

tolbutamide,18 acebutolol hydrochloride19 and etodolac20. In favorable cases, DC tablet

formulation containing as high as 99% API could be attained by using spherical

agglomerates.21

It would be reasonable to expect that spherical cocrystallization (SCC), which

combines spherical crystallization and cocrystallization, can simultaneously improve

manufacturability and dissolution performance of drugs to enable the development of DC

tablet formulations of high dose APIs. In this sense, SCC is a true integrated API crystal

and particle engineering strategy that can be potentially very useful to pharmaceutical

tablet development and manufacturing. SCC was firstly reported on the energetic material

ammonium nitrate to produce smokeless and easy packing cocrystal agglomerates.22

Similarly, it was applied to generate carbamazepine/saccharin cocrystal agglomerates.23

Moreover, it can be used to control the cocrystal stoichiometry and particle size.24

Surprisingly, SCC has not been explored to improve the manufacturing and dissolution

performance of poorly soluble APIs in high dose tablets. The goal of this work was to fill

this knowledge gap using a high dose model drug, griseofulvin (GSF), which is challenging

to formulate by DC process because of its poor powder flowability and tabletability in

addition to low solubility (10 μg·ml-1 at 37 oC).25

4.3 Materials and methods

4.3.1 Materials

Griseofulvin (GSF; Glaxo Laboratories, London, UK), Acesulfame potassium

(Acs-K, Tokyo Chemical Industry Co., Ltd., Japan), fumed silica (CAB-O-Sil, M5-P,

Page 114: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

92

Cabot Corporation, Boston, MA), microcrystalline cellulose (Avicel PH105; FMC

Biopolymers, Newark, DE), croscarmellose sodium (CCM-Na, SD-711, FMC Biopolymer,

Philadelphia, PA), magnesium stearate (MgSt, Mallinckrodt Inc., St. Louis, MO),

Hydroxypropylcellulose (HPC, EF Pharm, Ashland., Wilmington, DE), and

dimethylformamide (DMF, Sigma-Aldrich, St. Louis, MO)were received from respective

manufacturers and were used as received.

4.3.2 Preparation of spherical GSF-Acs cocrystal hydrate

A mixture of 0.01 mol of GSF and 0.01 mol of Acs-K was suspended in 70 mL of

methanol and water (2:8, v/v) mixture. Concentrated hydrochloric acid (1 mL) was added

to convert the Acs-K into acesulfame free acid (Acs-H) completely. The suspension was

agitated overnight. This reaction crystallization led to phase pure fine GSF-Acs crystals,

which was a monohydrate.25

The suspension was cooled to 0 oC in an ice-bath and stirred at 500 rpm using an

overhead mixer. Under stirring, DCM was added dropwise into the suspension until the

initially turbid medium became clear, which corresponded to the elimination of fine

crystals from the suspension. Then, the agitation was continued for 1h before filtration and

drying at 60 oC for 4 hr. In some batches, HPC (35 mg) was added to the vessel before

DCM was added to improve wettability of agglomerates.26 Subsequent steps were the same

as the process without HPC. DCM was a bridging liquid that facilitated the agglomeration

of fine GSF crystals.

Page 115: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

93

4.3.3 Powder X-ray Diffractometry (PXRD)

A powder X-ray diffractometer with Cu Kα radiation (1.54059 Å) was used to

characterize crystallographic properties of samples. Samples were scanned between 5 to

35° 2θ with a step size of 0.02° and a dwell time of 1 s/step. The X-ray tube voltage and

amperage were 40 kV and 40 mA, respectively. A theoretical PXRD pattern of GSF-Acs

was calculated from the single crystal structure (CCDC refcode: GEPJIW).25

4.3.4 Thermal Analyses

Powder samples (3−5 mg) were loaded into hermetically sealed aluminum pans

(Tzero) and heated from 30 to 200 °C at a rate of 10 °C/min using a differential scanning

calorimeter (Q2000, TA Instruments, New Castle, DE). The sample was purged with

nitrogen at a flow rate of 50 mL/min. The instrument was equipped with a refrigerated

cooling system. The temperature and cell constant were calibrated using high purity indium.

4.3.5 Scanning electron microscopy (SEM)

A thin layer platinum (thickness ~ 75 Å) was sputter-coated onto the samples,

which were mounted onto carbon tapes, using an ion-beam sputter (IBS/TM200S; VCR

Group Inc., San Clemente, CA). The particle morphology and surface features were

evaluated by scanning electron microscopy (JEOL 6500F; JEOL Ltd., Tokyo, Japan),

operated at SEI mode with an acceleration voltage of 5kV. A high vacuum (10-4-10-5 Pa)

was maintained during the imaging process.

Page 116: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

94

4.3.6 Contact angle measurement

Tablets, prepared by compressing powders at 300 MPa, were used for contact angle

measurements using the sessile drop method on a goniometer (MAC-3, Kyowa Interface

Science Co. Ltd., Japan). A drop of distilled water (∼2 μL) was gently placed on the surface

of the tablet using a syringe dispenser. The image of the water drop was recorded from the

side every 67 ms for 60 s using a high-speed camera. The angle between the sample surface

and the tangent line at the edge of the drop was determined using image analysis software,

FAMAS3.72 (Kyowa Interface Science Co. Ltd., Japan). Three measurements were made

at different locations on each tablet, and the mean and standard deviations were calculated.

4.3.7 Energy framework and topologic analysis of crystal structures

The pairwise intermolecular interaction energy was estimated using

CrystalExplorer and Gaussian09W with experimental geometry of GSF-Acs monohydrate

and GSF (Form I, CSD refcode: GRISFL06).27-29 The hydrogen positions were normalized

to standard neutron diffraction values before calculation using the B3LYP-D2/6-31G(d,p)

electron densities model. The total intermolecular interaction energy for a given molecule

is the sum of the electrostatic, polarization, dispersion, and exchange-repulsion

components with scale factors of 1.057, 0.740, 0.871, and 0.618, respectively.

Intermolecular interaction between two molecules is ignored when the closest atom - atom

distance was more than 3.8 Å. The interaction energies were graphically presented as a

framework by connecting centers of mass of molecules with cylinder thickness

Page 117: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

95

proportional to the total intermolecular interaction energies, where interaction energies less

than -5 kJ/mol were omitted for clarity. The crystal packing topology was analyzed using

a python program.30

4.3.8 Formulation and Tablet Compression

MCC (Avicel PH105) was used as a tablet binder because of its excellent

tabletability.31 However, the flowability of Avicel PH105 is poor. Thus, it was coated with

nano silica particles using a dry comilling process to improve its flowability before mixing

with API.32 Briefly, 50 g of Avicel PH105 and colloidal silica (1.0% loading) were

geometrically mixed and passed through a sieve (150 µm opening). The powder mixture

was then placed in the Comil chamber and passed through a stainless-steel screen (round

holes of 0.9905 mm, or “0.039”, diameter; part number: 7B039R03125) with an impeller

speed of 2200 rpm. The process was repeated 40 times to ensure uniform distribution of

silica particles onto Avicel PH105 particles. API and all other excipients (Table 4.1),

totaling 5 g, were added to a 30 mL plastic bottle, which was manually rotated diagonally

to allow particles slide and mix for 15 min (approximately 900 turns).

A universal material test machine (model 1485, Zwick/Roell, Germany) was used

to prepare tablets with compaction pressure ranging from 25 to 350 MPa at a speed of 4

mm/min. The punch tip and die wall were coated with a 5% (w/v) magnesium stearate

suspension in ethanol and then dried with a fan. Tablets were relaxed under ambient

environment for at least 24 h before being broken diametrically using a texture analyzer

(TA-XT2i; Texture Technologies Corporation, Scarsdale, NY). Tablet tensile strength was

Page 118: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

96

calculated from the breaking force (F) and tablet thickness (h) and diameter (d) using

equation (1).33 Tablet flashing was carefully removed to improve accuracy of measured

tablet thickness.34

𝜎 =2𝐹

𝜋𝑑ℎ (1)

4.3.9 Powder Flowability

Powder flowability was measured using a ring shear cell tester (RST-XS; Dietmar

Schulze, Wolfenbüttel, Germany), with a 10 mL cell at 1 and 3 kPa pre-shear normal

stresses. A yield locus was obtained by conducting shear test under different normal

stresses, from which the unconfined yield strength (fc) and major principal stress (σn) were

obtained by drawing Mohr’s circles. The flowability index (ffc) was calculated using Eq.

(2):

𝑓𝑓c =𝜎𝑛

𝑓c (2)

4.3.10 Expedited Friability

An expedited friability method was applied to determine the ability of compressed

tablets to resist wearing and abrasion.35 Coded tablets prepared under different pressures

were weighted, coded, and loaded into a friabilator (Model F2, Pharma Alliance Group

Inc., Santa Clarita, CA) at 25 rpm for 4 min. The final weight of each tablet was used to

calculate percent weight loss, which was plotted against compaction pressure to construct

Page 119: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

97

a friability plot. The compaction pressure corresponding to 0.8% tablet weight loss was

identified from the friability plot.35

4.3.11 Artificial Stomach-Duodenum Dissolution (ASD)

Although the griseofulvin is non-ionizable, the coformer of acesulfame (pKa=2.0)

shows strong acidity nature. Therefore, the solubility of griseofulvin - acesufulame

cocrystal is expected to be pH-dependent. Dissolution results in the single chamber

apparatus, such as USP method, may not reflect the real condition and result in wrong

conclusion. Thus, using ASD apparatus with two chambers (pH 1.2 and 6.8) is a more

reliable approach.

The ASD apparatus consisted of two jacketed beakers that were connected using a

short tubing, mimicking stomach chamber and duodenum chamber, respectively. The

temperature was kept at 37 °C by circulating water. Fluid flow from stomach chamber to

duodenum chamber was regulated by a programmed peristaltic pump (Masterflex L/S

Easy-Load II pump, Cole-Parmer, Vernon Hills, IL).36

To simulate human physiological conditions in the fasted state, 0.1 N HCl (pH =

1.2) was prepared for gastric liquid in stomach and 0.1 M sodium phosphate buffer (pH =

6.8) for the duodenum. The initial volume of the stomach chamber was 250 mL, which was

emptied to 50 mL, following a first-order kinetics with a half-life of 15 min to simulate

stomach empty process. Throughout the entire experiment, the fluid volume in the

duodenum chamber was maintained at 30 mL, achieved by setting a vacuum line at a

calibrated height. In addition, the two chambers were infused with fresh gastric or duodenal

Page 120: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

98

secretion liquid at 2 mL/min to mimic in vivo secretion processes. A fiber optic UV/Vis

probe was used to monitor the drug concentration. Mixing was achieved by an overhead

paddle stirrer in the stomach chamber and a magnetic stirrer in the duodenum chamber.

Prior to each experiment, calibration of all pumps and spectrometers were performed. All

fluids used in the experiment were degassed to avoid the generation of bubbles that may

interfere the data collection using the UV dip probe.

The drug concentration−time profile in the duodenum was analyzed, using the non-

compartmental method (PKsolver 2.0) to determine the peak plasma concentration (Cmax),

time to reach Cmax (Tmax), and area under the plasma concentration−time curve (AUC).37

4.4 Results and discussion

4.4.1 Characterization

The as-received GSF crystals were 1-5 µm in size and with irregular shape (Figure

4.1a). The GSF-Acs cocrystals (GSF-Acs) from reaction crystallization were prism shape

with size ranging 1 - 10 µm. The primary crystals in the cocrystal agglomerates without

HPC (GSF-Acs/SA) and with HPC (GSF-Acs/HPC) were comparable to those in the GSF-

Acs, indicating that the addition of DCM and HPC did not significantly influence the

property of the primary GSF crystals (Figure 4.1). This is understandable since the

cocrystal forming process was identical in all cases. The subsequent introduction of

bridging liquid and HPC affects, at most, attrition and ripening of these primary crystals.

Page 121: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

99

The size of the GSF-Acs/HPC was much smaller compared to that of the GSF-

Acs/SA (Figure 4.2 & Figure 4.3). The smaller size of GSF-Acs/HPC is attributed to the

impact on the agglomeration process by changes in surface tension between bridging liquid

and bulk medium or medium viscosity by the introduction of HPC. Although smaller, the

GSF-Acs/HPC agglomerates were more spherical than GSF-Acs/SA (Figure 4.2a & 4.2b).

Similar effect of HPC on size and shape of spherical agglomeration was reported for

etodolac.38

PXRD patterns of the GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC all matched well

with the calculated pattern from the GSF-Acs crystal structure (Figure 4.4). The slight

shift of some peaks to higher 2θ values in the calculated PXRD pattern is attributed to the

thermal expansion of crystal lattice since the crystal structure used for calculating PXRD

pattern was solved at 100 K while the experimental PXRD patterns were obtained at ~298

K. The largest 2θ shift (~0.5o) observed in 27.2o and 28.3o is attribute the ease of crystal

thermal expansion along the (0 6 0) and (0 6 1) planes. This is consistent with the fact that

absence of strong interlayer hydrogen bonds interaction between the (0 2 0). Moreover,

characteristic peaks of the starting materials were absent in the experimental PXRD

patterns, confirming their phase purity. DSC thermograms of the three different GSF-Acs

powders (Figure 4.5) were all similar to the reported data with two endotherms between

130-160 oC and 170-190 oC, respectively.25 Collectively, the phase purity of all three GSF-

Acs powders was confirmed.

The tabletability, i.e., tensile strength vs. compaction pressure, of the as-received

GSF was very poor as no intact tablets could be prepared due to lamination (Figure 4.6).

The fine GSF-Acs showed slightly improved tabletability with a maximum tensile strength

Page 122: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

100

of ~ 0.4 MPa at 100 MPa compaction pressure (Figure 4.6). Intact tablets were so weak

that very gentle handling was required when measuring their breaking strength. Extensive

tablet lamination still occurred when the compaction pressure was higher than 150 MPa.

The tableting performance of the GSF-Acs/SA was similar to GSF-Acs, except lamination

was observed earlier at 100 MPa compaction pressure. The GSF-Acs/HPC exhibited much

better tabletability compared to GSF-Acs and GSF-Acs/SA. In fact, the highest tensile

strength of GSF-Acs/HPC was 2.5 MPa at a compaction pressure of 150 MPa. The similar

tableting performance of GSF-Acs and GSF-Acs/SA indicated that extensive fracture of

the agglomerates had occurred, which eliminated the potential effect on tabletability by the

larger agglomerate size and microstructure of the SA. HPC is a good tablet binder.39, 40 The

improved tabletability of GSF-Acs/HPC at such a low level of HPC (< 1%) indicated

uniform coating of GSF-Acs crystals by HPC.38 Such polymer distribution is effective in

attaining an extensive three dimensional bonding network in the tablet with increased inter-

particular bonding area.41, 42

It has been demonstrated the solid forms with excellent compaction properties

usually display active slip/cleavage planes with lower interlayer interaction energy and

smooth plane topology, which facilitate plastic deformation.43 However, clear

identification of the active slip planes in both GSF and GSF-Acs is challenging by

visualization of crystal structures.44 While GSF showed no strong hydrogen bonds in its

structure and the GSF-Acs exhibited three-dimensional hydrogen bonds network, neither

shows visually clear molecular layers that can serve as active slip planes. The energy

framework of both crystal structures suggested an absence of crystallographic planes that

can serve as active slip planes with much weaker interlayer interactions (Figure 4.7, also

Page 123: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

101

see Table 4.2 for detailed intermolecular interaction energies). Additionally, all the

interlayer distances computed by topology analysis are negative, implying interlocking

between possible molecular layers, which hindered the slip among them. For example, the

most likely slip planes based on topology analysis, (0 1 1) for GSF and (1 -1 1) for GSF-

Acs, had interplanar overlapping of-3.80 Å and -1.72 Å, respectively. Thus, neither GSF

(Form I) nor GSF-Acs exhibits crystal structure that promotes plasticity, which is essential

for forming strong tablets. This is consistent with the very poor tabletability of both

crystals (Figure 4.6).

4.4.2 Formulation development

A successful formulation for tablet product development needs to meet several

criteria: (1) adequate blend flowability for smooth and reproducible hopper emptying and

die filling, which are critical for sustaining high speed tableting, (2) adequate mechanical

strength of the tablet (tabletability), (3) low friability, and (4) adequate dissolution

performance. For APIs exhibiting poor tabletability that still need to be delivered in high

doses, such as GSF, the use of tablet binders with excellent tabletability is necessary to

assure adequate tabletability of the final formulation. Avicel PH105 was chosen in this

work because of its excellent tabletability.31 The 70% loading of GSF-Acs (corresponding

to 55.7% GSF) was used (Table 4.1), since lamination was observed when cocrystal

loading was 80%.

All formulations exhibited good tabletability, where tablet tensile strength could

reach 2 MPa 45 at a compaction pressure above 120 MPa. The tabletability of formulations

Page 124: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

102

containing GSF-Acs, GSF-Acs/SA, or GSF-Acs/HPC was similar, which was lower than

the formulation containing GSF (Figure 4.8a). Based on the friability plots (Figure 4.8b),

all formulations met the minimum friability requirement of < 0.8% when the compaction

pressure was higher than 250 MPa.35 The GSF formulation exhibited the best friability,

which required no more than 150 MPa in order to attain < 0.8% friability (Figure 4.8b).

The superior tabletability and friability of the GSF formulation is reasonable because it

contained about 60% higher amount of highly compressible Avicel PH105 than other three

formulations (Table 4.1). Therefore, it is expected that both tabletability and friability of

this formulation is better.46 It is interesting to note that the tabletability of the formulation

containing GSF-Acs/HPC was the same as other two GSF-Acs containing formulations,

despite the significantly better tabletability of GSF-Acs/HPC (Figure 4.6). This highlights

the difficulty in accurately predicting tabletability of a mixture from constitutive

components and the need for mechanistic investigation into compaction behavior of

mixtures.47, 48

Another important property of a successful tablet formulation is the flowability.

Only GSF-Acs/HPC and GSF-Acs/SA formulations exhibited flowability better than that

desired for high speed tableting, marked by Avicel PH102 (Figure 4.9).49 The differences

in flowability are attributed to the different particle sizes and shapes of the API powders

used in these formulations. Both GSF-Acs and GSF consisted of small elongated crystals

(Figure 4.1), while both GSF-Acs/SA and GSF-Acs/HPC consisted of large and more

round agglomerates (Figure 4.2).8

Having verified the acceptable manufacturability of both formulations containing

GSF-Acs agglomerates, we then assessed the dissolution performance of formulated tablets

Page 125: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

103

using the ASD apparatus. All three cocrystal formulations released GSF more rapidly than

the formulation of GSF. Both the AUC and Cmax followed the order of GSF-Acs/HPC >

GSF-Acs >> GSF-Acs/SA > GSF (Figure 4.10a, Table 4.3). The enhanced dissolution

performance of the cocrystal formulations compared to that of GSF may be attributed to

the approximately 3 fold higher solubility of GSF-Acs than GSF.25 The time to reach Cmax,

i.e., Tmax, is longer for formulation that exhibited higher Cmax, which is reasonable.

To further understand the different dissolution performance of the three cocrystal

formulations, wettability of the pure APIs was assessed using contact angle measurement.

Contact angle of GSF could not be determined using this method because it did not form

intact tablets. However, it is expected to be high given its significantly lower aqueous

solubility. Contact angle of pure cocrystal samples follow the order of GSF-Acs/SA> GSF-

Acs > GSF-Acs/HPC (Figure 4.10b). Thus, GSF-Acs/HPC exhibited better wettability than

GSF-Acs/SA and GSF-Acs. The better wettability of GSF-Acs/HPC is consistent with the

better dissolution of the GSF-Acs/HPC tablet, despite its much larger size of GSF-

Acs/HPC than GSF-Acs. This suggests that wetting is a key step in the dissolution of GSF-

Acs from tablet. The effect of surface area of API on dissolution is shown by the

significantly faster dissolution from tablet containing GSF-Acs than that containing GSF-

Acs/SA (Figure 4.10a), although wettability of pure API powders was only slightly

different (Figure 4.10b). Thus, the overall effect of both faster wetting and larger API size

of GSF-Acs/HPC led to dissolution performance of its tablet comparable to that of GSF-

Acs tablet (larger surface area and intermediate wetting), both were significantly better

than GSF (poor wetting and large area) and GSF-Acs/SA (small surface area and

intermediate wetting) formulations.

Page 126: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

104

The overall performance of each formulation considering manufacturability and

dissolution is summarized in Table 4.4. Only the GSF-Acs/HPC formulation exhibited

acceptable manufacturability and significantly improved dissolution performance. Both the

GSF and GSF-Acs formulation suffered poor flowability due to the small size and poor

flowability of API. Thus, they are unfit for commercial tablet manufacturing. The GSF-

Acs/SA formulation exhibited good manufacturability but only marginal improvement in

dissolution. Therefore, only the use of GSF-Acs/HPC successfully enabled a high dose DC

tablet formulation of GSF, which was both manufacturable and fast releasing.

4.5 Conclusions

In this work, an integrated crystal and particle engineering approach, spherical

cocrystallization (SCC), was used to prepare GSF-Acs spherical agglomerates exhibiting

both improved manufacturability and dissolution performance. Such spherical cocrystal

agglomerates enabled the development of a high GSF loading (55.7%) direct compression

tablet formulation suitable for manufacturing of GSF tablets with enhanced dissolution

performance. The simultaneous improvement of tablet manufacturability and dissolution

by SCC is a powerful approach that may have broad applicability to enable the successful

development of high dose tablets of other poorly soluble drugs.

Page 127: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

105

Table 4.1. Formulation Composition of GSF-containing tablet using different powder

forms

Components (%) Function GSF GSF-Acs GSF-Acs/SA GSF-Acs/HPC

GSF or GSF-Acs Active 55.7 70 70 70

Silica coated MCC

PH105 Dry binder 38.8 24.5 24.5 24.5

CCS-Na Disintegrate 5 5 5 5

Magnesium stearate Lubricant 0.5 0.5 0.5 0.5

Total 100 100 100 100

Page 128: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

106

Figure 4.1. SEM images of a) GSF, b) GSF-Acs, c) GSF-Acs/SA and d) GSF-Acs/HPC.

a) b)

c)

Page 129: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

107

Figure 4.2. Microscopy images of (a) GSF-Acs/HPC and (b) GSF-Acs/SA.

(a) (b)

Page 130: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

108

Figure 4.3. Particle size distribution of GSF-Acs/SA and GSF-Acs/HPC powders.

Page 131: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

109

Figure 4.4. PXRD patterns of Acs-H, GSF, GSF-Acs/HPC, GSF-Acs/SA, GSF-Acs, and

calculated GSF-Acs.

Page 132: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

110

Figure 4.5. Thermal behavior of GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC

characterized by differential scanning calorimetry.

Page 133: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

111

Figure 4.6. Tabletability profiles of pure GSF, GSF-Acs, GSF-Acs/SA, and GSF-Acs/HPC.

Page 134: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

112

Figure 4.7. Energy frameworks of a) GSF, and b) GSF-Acs, both viewed into a axis of

corresponding unit cell. The interaction energy threshold was set at −5 kJ/mol.

a) b)

Page 135: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

113

Table 4.2. Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p)

dispersion corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)),

polarization (E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the

energy are listed. R indicated the distance between centers of mass of the pair of molecules.

Sym. Op. R Eele Epol Edis Erep Etot

GSF (form I)

x, y, z 8.90 -3.1 -1.6 -19.8 8.2 -16.7

-y, x, z+1/4 6.75 -9.2 -2.3 -32.1 23.6 -24.8

x, y, z 8.90 -2.2 -2.3 -11.0 11.6 -6.5

-y, x, z+1/4 7.00 -30.3 -12.0 -59.1 47.8 -62.9

y, -x, z+3/4 9.59 1.8 -2.7 -20.3 11.8 -10.6

-y, x, z+1/4 9.77 -11.6 -7.0 -23.7 18.0 -27.0

-x, -y, z+1/2 10.47 -11.2 -4.4 -19.3 13.6 -23.5

GSF-Acs cocrystal

- 7.76 -38.7 -9.6 -5.6 47.1 -23.8

- 6.00 2.2 -1.1 -4.1 2.1 -0.7

- 6.71 0.1 -0.2 -0.8 0.0 -0.7

- 7.50 -41.0 -9.8 -6.9 46.2 -28.1

- 7.36 2.4 -0.7 -3.0 1.3 0.3

- 7.35 4.2 -0.5 -2.1 0.2 2.4

- 4.37 -10.6 -2.2 -4.9 4.1 -14.6

- 4.45 -96.9 -24.3 -7.6 107.9 -60.5

- 7.97 -8.3 -3.4 -9.3 2.8 -17.6

- 6.07 -17.8 -4.0 -39.2 26.1 -39.8

- 10.37 0.2 -0.4 -1.5 0.0 -1.5

x, y, z 8.71 4.4 -0.7 -2.3 0.4 2.4

- 4.45 -96.9 -24.3 -7.6 107.9 -60.5

- 7.66 -9.3 -5.4 -18.1 15.1 -20.2

- 7.25 -2.1 -1.6 -4.7 1.1 -6.7

- 7.61 -19.6 -4.8 -20.7 19.4 -30.3

Page 136: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

114

- 5.55 -25.7 -8.7 -58.2 47.3 -55.0

- 4.37 -10.6 -2.2 -4.9 4.1 -14.6

x, y, z 13.62 -8.5 -2.9 -4.6 5.6 -11.7

x, y, z 8.71 -1.1 -3.7 -14.6 13.6 -8.2

- 8.29 2.9 -1.6 -17.5 7.5 -8.7

- 7.72 -20.1 -5.9 -34.7 26.9 -39.2

- 8.14 -16.2 -3.9 -26.5 13.2 -35.0

x, y, z 11.53 -12.2 -4.6 -10.9 11.3 -18.8

-x, y+1/2, -z 11.13 -0.1 -0.5 -5.4 2.0 -4.0

- 8.85 -16.1 -4.0 -29.3 16.8 -35.1

- 7.84 -20.6 -5.3 -54.7 43.7 -46.3

- 7.25 -2.1 -1.6 -4.7 1.1 -6.7

- 7.66 -9.3 -5.4 -18.1 15.1 -20.2

- 7.50 -41.0 -9.8 -6.9 46.2 -28.1

- 6.84 -1.2 -1.1 -25.7 13.8 -16.0

- 5.55 -25.7 -8.7 -58.2 47.3 -55.0

- 10.37 0.2 -0.4 -1.5 0.0 -1.5

- 7.36 2.4 -0.7 -3.0 1.3 0.3

- 7.35 4.2 -0.5 -2.1 0.2 2.4

x, y, z 8.71 0.6 -3.8 -14.5 15.2 -5.3

x, y, z 11.53 -16.4 -6.6 -18.4 17.9 -27.2

-x, y+1/2, -z 11.52 -0.1 -0.4 -3.8 0.5 -3.3

- 7.61 -19.6 -4.8 -20.7 19.4 -30.3

- 6.07 -17.8 -4.0 -39.2 26.1 -39.8

- 6.00 2.2 -1.1 -4.1 2.1 -0.7

- 7.97 -8.3 -3.4 -9.3 2.8 -17.6

- 7.76 -38.7 -9.6 -5.6 47.1 -23.8

- 6.71 0.1 -0.2 -0.8 0.0 -0.7

Page 137: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

115

Figure 4.8. Manufacturability of formulated GSF, GSF-Acs, GSF-Acs/SA, and GSF-

Acs/HPC, (a) tabletability and (b) friability.

a) b)

Page 138: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

116

Figure 4.9. Flowability plots of formulated GSF, GSF-Acs, GSF-Acs/SA and GSF-

Acs/HPC.

Page 139: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

117

Figure 4.10. (a) ASD duodenum concentration profiles for formulated GSF, GSF-Acs,

GSF-Acs/SA and GSF-Acs/HPC tablets prepared at 250 MPa, (b) wettability of GSF-

Acs/HPC, GSF-Acs/SA and GSF-Acs.

b) a)

Page 140: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

118

Table 4.3. Simulated pharmacokinetic parameters of various GSF-containing tablets (n =3)

GSF GSF-Acs GSF-Acs/SA GSF-Acs/HPC

Tmax (min) 20 ± 0.0 20 ± 0.0 25 ± 7.1 23 ± 4.2

Cmax (μg/ml) 15.3 ± 0.2 47.9 ± 1.6 21.1 ± 0.8 51.5 ± 3.4

AUC0-t

(μg/ml min) 1394 ± 55 3255 ± 152 1910 ±264 3697 ± 297

Page 141: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

119

Table 4.4. Summary of various GSF tablet formulations in meeting key criteria (meeting–

Y, not meeting – N)

Formulation Tabletability

(≥ 2 MPa)

Friability

(≤ 0.8%)

Flowability

(better than

Avicel PH102)

Dissolution Overall

GSF Y Y N N N

GSF-Acs Y Y N Y N

GSF-Acs/SA Y Y Y N N

GSF-Acs/HPC Y Y Y Y Y

Page 142: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

120

4.6 References

1. Sun, W.-J.; Aburub, A.; Sun, C. C. Particle engineering for enabling a formulation

platform suitable for manufacturing low-dose tablets by direct compression. J. Pharm. Sci.

2017, 106, 1772-1777.

2. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high

drug loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

3. Lakshman, J. P.; Kowalski, J.; Vasanthavada, M.; Tong, W. Q.; Joshi, Y. M.;

Serajuddin, A. T. Application of melt granulation technology to enhance tabletting

properties of poorly compactible high-dose drugs. J. Pharm. Sci. 2011, 100, 1553-1565.

4. Sun, W.-J.; Chen, H.; Aburub, A.; Sun, C. C. A platformdirect compression

formulation for low dose sustained-release tablets enabled by a dual particle engineering

approach. Powder Technol. 2019, 342, 856-863.

5. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

6. Rabinow, B.; Kipp, J.; Papadopoulos, P.; Wong, J.; Glosson, J.; Gass, J.; Sun, C.

S.; Wielgos, T.; White, R.; Cook, C.; Barker, K.; Wood, K. Itraconazole iv nanosuspension

enhances efficacy through altered pharmacokinetics in the rat. Int. J. Pharm. 2007, 339,

251-60.

7. Keck, C. M.; Muller, R. H. Drug nanocrystals of poorly soluble drugs produced by

high pressure homogenisation. Eur. J. Pharm. Biopharm. 2006, 62, 3-16.

8. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

9. Sastry, S. V.; Nyshadham, J. R.; Fix, J. A. Recent technological advances in oral

drug delivery – a review. Pharm. Sci. Technolo. Today 2000, 3, 138-145.

10. Williams, H. D.; Trevaskis, N. L.; Charman, S. A.; Shanker, R. M.; Charman, W.

N.; Pouton, C. W.; Porter, C. J. H. Strategies to address low drug solubility in discovery

and development. Pharmacol. Rev. 2013, 65, 315-499.

11. Hickey, M. B.; Peterson, M. L.; Scoppettuolo, L. A.; Morrisette, S. L.; Vetter, A.;

Guzman, H.; Remenar, J. F.; Zhang, Z.; Tawa, M. D.; Haley, S.; Zaworotko, M. J.;

Almarsson, O. Performance comparison of a co-crystal of carbamazepine with marketed

product. Eur. J. Pharm. Biopharm. 2007, 67, 112-119.

12. Martinello, T.; Kaneko, T. M.; Velasco, M. V. R.; Taqueda, M. E. S.; Consiglieri,

V. O. Optimization of poorly compactable drug tablets manufactured by direct

compression using the mixture experimental design. Int. J. Pharm. 2006, 322, 87-95.

13. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization

of aceclofenac and paracetamol spherical crystals and their tableting properties. Powder

Technol. 2015, 274, 446-454.

14. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

Page 143: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

121

15. Fadke, J.; Desai, J.; Thakkar, H. Formulation development of spherical crystal

agglomerates of itraconazole for preparation of directly compressible tablets with enhanced

bioavailability. AAPS PharmSciTech 2015, 16, 1434-1444.

16. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

17. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

18. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion metho. Chem. Pharm. Bull. 1990, 38, 733-739.

19. Kawashima, Y.; Cui, F.; Takeuchi, H.; Niwa, T.; Hino, T.; Kiuchi, K.

Improvements in flowability and compressibility of pharmaceutical crystals for direct

tabletting by spherical crystallization with a two-solvent system. Powder Technol. 1994,

78, 151-157.

20. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

21. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient - a tale of spherical crystallization. J. Pharm. Sci. 2018, 108, 1396-1400.

22. Lee, T.; Chen, J. W.; Lee, H. L.; Lin, T. Y.; Tsai, Y. C.; Cheng, S.-L.; Lee, S.-W.;

Hu, J.-C.; Chen, L.-T. Stabilization and spheroidization of ammonium nitrate: Co-

crystallization with crown ethers and spherical crystallization by solvent screening. Chem.

Eng. J. 2013, 225, 809-817.

23. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

24. Wang, J. R.; Bao, J.; Fan, X.; Dai, W.; Mei, X. Ph-switchable vitamin b9 gels for

stoichiometry-controlled spherical co-crystallization. Chem. Commun. 2016, 52, 13452-

13455.

25. Aitipamula, S.; Vangala, V. R.; Chow, P. S.; Tan, R. B. H. Cocrystal hydrate of an

antifungal drug, griseofulvin, with promising physicochemical properties. Cryst Growth

Des. 2012, 12, 5858-5863.

26. Talukder, R.; Reed, C.; Durig, T.; Hussain, M. Dissolution and solid-state

characterization of poorly water-soluble drugs in the presence of a hydrophilic carrier.

AAPS PharmSciTech 2011, 12, 1227-1233.

27. Chen, Y.; Liu, C.; Chen, Z.; Su, C.; Hageman, M.; Hussain, M.; Haskell, R.;

Stefanski, K.; Qian, F. Drug–polymer–water interaction and its implication for the

dissolution performance of amorphous solid dispersions. Mol. Pharm. 2015, 12, 576-589.

Page 144: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

122

28. Tao, J.; Sun, Y.; Zhang, G. G. Z.; Yu, L. J. P. R. Solubility of small-molecule

crystals in polymers: D-mannitol in pvp, indomethacin in pvp/va, and nifedipine in pvp/va.

Pharm. Res. 2009, 26, 855-864.

29. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

30. Bryant, M. J.; Maloney, A. G. P.; Sykes, R. A. Predicting mechanical properties

of crystalline materials through topological analysis. CrystEngComm 2018, 20, 2698-2704.

31. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

32. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

33. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

34. Paul, S.; Chang, S.-Y.; Sun, C. C. The phenomenon of tablet flashing — its impact

on tableting data analysis and a method to eliminate it. Powder Technol. 2017, 305, 117-

124.

35. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int. J. Pharm. 2015, 484, 146-155.

36. Wang, C.; Chopade, S. A.; Guo, Y.; Early, J. T.; Tang, B.; Wang, E.; Hillmyer, M.

A.; Lodge, T. P.; Sun, C. C. Preparation, characterization, and formulation development

of drug-drug protic ionic liquids of diphenhydramine with ibuprofen and naproxen. Mol.

Pharm. 2018, 15, 4190-4201.

37. Zhang, Y.; Huo, M.; Zhou, J.; Xie, S. Pksolver: An add-in program for

pharmacokinetic and pharmacodynamic data analysis in microsoft excel. Comput. Methods

Programs Biomed. 2010, 99, 306-314.

38. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

39. Rajniak, P.; Mancinelli, C.; Chern, R. T.; Stepanek, F.; Farber, L.; Hill, B. T.

Experimental study of wet granulation in fluidized bed: Impact of the binder properties on

the granule morphology. Int. J. Pharm. 2007, 334, 92-102.

40. Skinner, G. W.; Harcum, W. W.; Barnum, P. E.; Guo, J.-H. The evaluation of fine-

particle hydroxypropylcellulose as a roller compaction binder in pharmaceutical

applications. Drug Dev. Ind. Pharm. 1999, 25, 1121-1128.

41. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

42. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

Page 145: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

123

43. Sun, C.; Grant, D. J. W. Influence of crystal structure on the tableting properties

of sulfamerazine polymorphs. Pharm. Res. 2001, 18, 274-280.

44. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

45. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

46. Paul, S.; Sun, C. C. Dependence of friability on tablet mechanical properties and a

predictive approach for binary mixtures. Pharm. Res. 2017, 34, 2901-2909.

47. Sun, C. C. A classification system for tableting behaviors of binary powder

mixtures. Asian J. Pharm. 2016, 11, 486-491.

48. Wu, C. Y.; Best, S. M.; Bentham, A. C.; Hancock, B. C.; Bonfield, W. Predicting

the tensile strength of compacted multi-component mixtures of pharmaceutical powders.

Pharm. Res. 2006, 23, 1898-1905.

49. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

Page 146: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

124

Chapter 5

A Novel Quasi-Emulsion Solvent Diffusion-based Spherical

Cocrystallization Strategy for Simultaneously Improving the

Manufacturability and Dissolution of Indomethacin

Page 147: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

125

5.1 Synopsis

Successful development of tablet formulations of many active pharmaceutical

ingredients (APIs) is challenged by their poor manufacturability (e.g., flowability,

tabletability) and dissolution characteristics. Here, we report a novel quasi-emulsion

solvent diffusion cocrystallization (QESD-CC) method as an integrated crystal and particle

engineering approach for successful generation of spherical cocrystal agglomerates of the

poorly soluble drug indomethacin (IMC) and the sweet co-crystallization agent saccharin

(SAC). The QESD-CC process consists of two distinct steps: 1) formation of a transient

emulsion containing solvated IMC and SAC, and 2) subsequent precipitation of IMC-SAC

cocrystals from the emulsion as hollow spherical particles. Solution 1H NMR analyses and

computational modeling studies indicated that hydroxypropyl methylcellulose (HPMC)

preferentially interacts with the SAC molecules through hydrogen bonds, thus driving the

polymer to form a shell that stabilizes the transient emulsion droplets and coats the

resulting particles. Spherical QESD-CC particles exhibited excellent flowability, while the

HPMC coating and micro-size primary crystals led to excellent tabletability. Outstanding

manufacturability and high cocrystal solubility thus enabled the successful development of

a high drug loading tablet formulation comprising 46.3 wt% IMC.

5.2 Introduction

Tablet manufacturing by direct compression (DC) processes is attractive to the

pharmaceutical industry because of its cost effectiveness. DC tableting is particularly

suitable for continuous manufacturing due to its procedural simplicity.1 For typical active

pharmaceutical ingredients (APIs) with doses ≥ 100 mg, high API-loaded tablet

formulations are preferred to reduce tablet size, making them easier to swallow and thus

Page 148: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

126

enhancing patient compliance. However, a DC tablet formulation typically does not contain

more than 30 wt% API due to its common and undesirable physical properties, including

poor powder flowability, poor tabletability, and poor water solubility, which requires the

use of large amounts of solubilizing excipients to enable adequate therapeutic absorption.2-

4 APIs exhibiting both poor powder properties and low solubility present a major challenge

for developing a DC tablet formulation.

Spherical crystallization is an established technique for improving micromeritic

properties (e.g., particle size, shape, bulk density) of drugs by agglomerating the small

primary crystals into large and spherical agglomerates, mainly by spherical agglomeration

(SA)5-7 and quasi-emulsion solvent diffusion (QESD) techniques.8, 9 In fact, spherical

crystallization of water-soluble ferulic acid enabled the development of a DC tablet

formulation containing a record high 99% API.3 However, spherical crystallization often

detrimentally impacts the rate of drug release of poorly soluble APIs because this process

reduces the exposed surface area available for dissolution.10

The problem of slow dissolution of poorly soluble APIs may be addressed by the

use of solubilizing excipient(s) in tablet formulations. However, this approach invariably

leads to large tablets due to the significant amounts of solubilizing excipient required.

Another effective approach is to use a soluble solid form of the poorly soluble API, such

as an amorphous solid dispersion (ASD),11 a salt,12 or a cocrystal.13 Among these options,

ASD requires the use of a large amount of excipient as a matrix to maintain adequate

physical stability against crystallization14 as well as a high dissolution rate of the API.15, 16

Salt formation is applicable only to ionizable APIs and is limited by the availability of

pharmaceutically acceptable counterions.17 In comparison to ASD and salt formation,

Page 149: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

127

cocrystals exhibit unique advantages for enhancing dissolution of poorly soluble APIs

because they are physically stable in solid state, applicable to non-ionizable molecules, and

the number of pharmaceutically acceptable co-formers is substantial.17, 18 However, the use

of these more soluble solid form aggravates the manufacturability problem for developing

a high dose DC tablet formulation because the weight percentage of any of these solid

forms is invariably higher than that of the parent API. To ensure acceptable

manufacturability, micromeritic properties of the new API solid form cannot be poor in

order to maintain a high API loading in the tablet.

Spherical cocrystallization has the potential to simultaneously improve both tablet

manufacturability and dissolution performance.19, 20 Although earlier work successfully

demonstrated, that spherical cocrystallization of griseofulvin (GSF) with the co-former

Acesulfame (Acs) profoundly improves flowability and dissolution characteristics, the

tableting performance of spherical GSF-Acs cocrystal agglomerates remained poor, which

limited GSF loadings in each tablet.19 The reason for the poor tabletability of GSF-Acs

was the large average size of the primary crystals in the spherical agglomerates, which led

to small interparticle bonding areas and consequently weak tablets.21 The large size of

primary cocrystals was attributed to the use of a slurry cocrystallization process, which

allowed large cocrystals to grow. Hence, an improved SCC process that can produce

spherical cocrystal agglomerates consisting of fine primary crystals has the potential to

enable the development of high dose DC tablets through simultaneously improve the

flowability, tabletability, and dissolution of APIs. To attain this goal, we developed a novel

integrated crystal and particle engineering approach called quasi-emulsion solvent

diffusion cocrystallization (QESD-CC). In contrast to slurry cocrystallization, we

Page 150: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

128

hypothesized that QESD-CC would yield large and spherical particles consisting of much

smaller primary cocrystals due to the high degree of emulsion supersaturation during

QESD, 22 leading to better tabletability and flowability while retaining a high dissolution

rate. A successful QESD-CC process requires the following two conditions: 1) formation

of sufficiently stable transient emulsions of co-formers; 2) raction between API and co-

former reaction to form a cocrystal upon solvent counter-diffusion. In this work, we chose

indomethacin (IMC)-saccharin (SAC) cocrystal to test the feasibility of QESD-CC because

IMC-SAC cocrystal could be prepared by anti-solvent (AS) precipitation.23 We also used

HPMC as the emulsion stabilizer because of its successful use in other QESD processes.

Since the IMC-SAC cocrystal exhibits much higher solubility than neat IMC (13 times

greater in water and 13-65 times higher at pH 1-3),24 it can likely improve the dissolution

of IMC.

5.3 Materials and methods

5.3.1 Materials

Indomethacin (γ-IMC, Tokyo Chemical Industry, Tokyo, JP) and saccharin (SAC,

Carbosynth Limited, Berkshire, UK), were chosen as the model API and co-former,

respectively. Microcrystalline cellulose (Avicel PH102, FMC, Philadelphia, PA), HPMC

(K100M, MW 1,000 kg/mol, Ashland Specialty Ingredients, Wilmington, DE),

croscarmellose sodium (CCM-Na, SD-711, FMC Biopolymer, Philadelphia, PA),

magnesium stearate (MgSt, Mallinckrodt Inc., St. Louis, MO), dimethylformamide (Fisher

Scientific, Hampton, NH), methanol (Fisher Scientific, Hampton, NH) and ultrapure

deionized water (0.066 μS/cm, Thermo Scientific, USA) were used as received. The

Page 151: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

129

chemical structures of IMC, SAC, and HPMC are shown in Figure 5.2. All crystallization

experiments were performed in glass beakers.

5.3.2 Preparation of IMC-SAC cocrystal powders

5.3.2.1 Fine IMC-SAC cocrystal by anti-solvent (AS) method

Based on a literature study,24 an excess amount of SAC was required to account for

the large solubility difference between IMC and SAC to ensure phase purity of the

produced cocrystal powder. Therefore, 5 mmol of IMC and 15 mmol of SAC were co-

dissolved in 140 mL of methanol. The methanol solution was then poured into 140 mL of

water and mixed with a magnetic stirring bar under ambient conditions for 30 min to

produce fine IMC-SAC cocrystals (AS). The resulting suspension was filtered and dried at

60 °C overnight. The yield of this process was ~ 70%. A small amount of this powder was

dispersed onto a glass slide for analysis under a polarized light microscope (Eclipse E200;

Nikon, Tokyo, Japan).

5.3.2.2 IMC-SAC spherical cocrystal agglomerates by QESD-CC method

A solution of 4 mmol of IMC and 5 mmol of SAC in 6 mL of DMF was poured (in ≤10

s) in into 40 mL of distilled water or 0.01%, 0.1%, or 0.5% (w/v) HPMC aqueous solution.

A HPMC solution at ≥1% concentration was not attempted because they were too viscous

for effective mixing. Our preliminary work also suggested that DMF to water ratios

significantly deviating from 6:40 tended to cause phase impurity and amount adjustment

Page 152: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

130

of SAC was also required to attain phase purity of the cocrystal powders. The suspension

was agitated at 300 rpm using an overhead stirrer for 3 h at room temperature (23.5 oC) to

ensure complete transformation into IMC-SAC. Agitation speed at 300 rpm was used

because it generated spherical agglomerates with a suitable particle size distribution (200-

500 μm). The produced solid particles were filtered and dried, without solvent wash, at

60 °C overnight to yield a powder, hereafter referred as QESD-CC. A small amount of this

powder was dispersed on a glass slide for imaging by polarized light microscopy.

5.3.3 Particle size distribution (PSD)

The particle size of as-received IMC, AS, and QESD-CC powders was determined

using a particle size analyzer (Microtrac SIA, Montgomeryville, PA). Approximately 25

mg of each powder was suspended in Isopar G Fluid (∼5 mL). The as-received IMC and

AS samples were sonicated for 30 s to fully disperse particles. Both IMC-SAC and polymer

HPMC exhibited negligible solubility in this medium. The suspension was added into a

sample delivery controller (SDC, Montgomeryville, PA) and circulated through the closed

system. A high-speed camera captured images of particles in focus. Using the Microtrac

Flex image analysis software provided with the instrument, the diameter of a circle with an

area equivalent to the projected area of individual particles was determined. The volume

of the particle was calculated assuming a spherical shape to determine the volume

distribution of a powder.

5.3.4 Specific surface area (SSA) measurement

Page 153: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

131

Nitrogen adsorption-desorption isotherms for each sample were obtained using a

surface area analyzer (Micromeritics ASAP 2020, Norcross, GA) at 77 K. Before each

measurement, 0.5–1.0 g of accurately weighed sample was placed into a glass tube and

degassed under a vacuum of 15 µm Hg at 27 °C for at least 24 hrs. The amount of the

adsorbed nitrogen at different partial pressures of nitrogen in helium carrier gas was

measured at a temperature of 77.35 K. The data is shown in Table 5.1. The Brunauer–

Emmett–Teller (BET) equation25 was used to fit the data and calculate the sample surface

area. All measurements were made in triplicate.

5.3.5 HPMC content determination via size-exclusion chromatography (SEC)

HPMC concentrations were determined by size exclusion chromatography (SEC)

with refractive index detection.26 SEC measurements were carried out on an Agilent 1260

liquid chromatograph equipped with one guard column and three separating CATSEC

columns with pore sizes of 1000, 300 and 100 Å, operating in 0.1 M Na2SO4(aq) containing

1 wt% acetic acid at an eluent flow rate of 0.4 mL/min at 22 °C. Analyte detection relied

on a Wyatt Optilab T-rEX refractive index (RI) detector and a Wyatt Dawn Heleos II multi-

angle static light scattering detector. The peak area from RI detector signal that corresponds

to HPMC elution was used to determine the HPMC concentration in the ASTRA software.

We determined the refractive index increment (dn/dc) of HPMC as 0.1165 mL/g using a

pure HPMC (K100M) solution with 0.81 mg/mL concentration, assuming 100% mass

recovery. We checked the validity of our analysis by measuring the concentrations of

HPMC solutions from 0.45–1.04 mg/mL and the obtained values agreed within 99.9%

(Figure 5.1). On this basis, we then quantified the amount of HMPC in the QESD-CC

Page 154: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

132

powders by dissolving ~0.65 g QESD-CC sample in 5 mL of HPLC grade tetrahydrofuran

(THF) to precipitate the THF-insoluble HPMC. The precipitate was isolated by

centrifugation (11 644 g, 10 min), after which the supernatant THF solution containing

IMC and SAC was carefully decanted. The isolated HPMC was re-dispersed in 5 mL THF

to wash away residual IMC and SAC by sonication for 10 min, and it was again isolated

by centrifugation. The polymer was then dried under vacuum to remove residual THF. The

isolated HPMC from the QESD-CC samples was dissolved in 5 mL SEC eluent and the

concentration determined from the aforementioned calibration curve. The reported HPMC

content in QESD-CC was determined as the average of three independent measurements.

5.3.6 Powder X-ray Diffractometry (PXRD)

The crystallographic properties of samples were characterized by powder X-ray

diffractometer with Cu Kα radiation (1.54059 Å). Samples were scanned over the range 5

≤ 2θ ≤ 35° with a step size of 0.02° and a dwell time of 1 s/step. The divergence and anti-

scattering slits on the incident beam path were 1/16 and 1/8 degree, respectively, and the

antiscattering slit on the diffracted beam path was 5.5 mm (X’Celerator). The X-ray tube

voltage and amperage were 45 kV and 40 mA, respectively. A theoretical PXRD pattern

of IMC-SAC was calculated from the single crystal structure (CCDC refcode: UFERED).27

5.3.7 Thermal Analyses

Powder samples (~ 5 mg) were loaded into hermetic aluminum pans and heated

from 30 to 190 °C at a heating rate of 10 °C/min on a differential scanning calorimeter

Page 155: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

133

(Q1000, TA Instruments, New Castle, DE) under a continuous helium purge at a flow rate

of 25 mL/min. DSC heat flow was calibrated with the enthalpy of melting of indium, and

the temperature was calibrated with the melting transitions of indium and cyclohexane. To

determine the volatile content, each sample (~ 5 mg) was placed in an open aluminum pan

and analyzed on a thermogravimetry analyzer (Q500, TA Instruments, New Castle, DE).

All samples were heated from room temperature to 300 °C at 10 °C/min under 25 mL/ min

nitrogen purge.

5.3.8 Scanning electron microscopy (SEM)

Samples were mounted onto stubs with carbon tape. A thin layer platinum

(thickness ~ 75 Å) was coated onto the samples using an ion-beam sputter (IBS/TM200S;

VCR Group Inc., San Clemente, CA). The particle morphology and surface features were

evaluated by scanning electron microscopy (JEOL 6500F; JEOL Ltd., Tokyo, Japan),

operated in SEI mode with an accelerating voltage of 5 kV under high vacuum (10-4-10-5

Pa).

5.3.9 Computational methods

5.3.9.1 Energy framework and topological analyses of crystal structures

Pairwise intermolecular interaction energies were estimated using CrystalExplorer

and Gaussian09 from the experimental geometries of the IMC-SAC cocrystal and IMC (γ

form, CSD refcode: INDMET).28, 29 The hydrogen positions were normalized to standard

neutron diffraction values prior to calculation using the CE-B3LYP electron densities

Page 156: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

134

model. The data is shown in Table 5.2. The total intermolecular interaction energy for a

given molecule is the sum of the electrostatic, polarization, dispersion, and exchange-

repulsion components with scale factors of 1.057, 0.740, 0.871, and 0.618, respectively.30

Intermolecular interaction between two molecules is ignored when the closest interatomic

distance was > 3.8 Å. Interaction energies were visualized as a framework by connecting

centers of mass of molecules with cylinder thicknesses proportional to the total

intermolecular interaction energies, where interaction energies weaker than -10 kJ/mol

were omitted for clarity. The crystal packing topology was analyzed using a Python

program.31

5.3.9.2 Drug–Polymer Intermolecular Interaction

To characterize the molecular interactions between IMC-SAC and HPMC,

optimization of their complexation was performed in Gaussian16 using DFT hybrid

functional M06 level of theory and 6-31(d, p) basis sets.32, 33 To reduce the computational

cost, three different monomer units bearing different substituents (R = H, CH3, or

CH2CH(OH)CH3 in Figure 5.2) of HPMC were used. The IMC-SAC complex was from

the reported structure (UFERED). Structure optimization was carried out without any

constraints and local minima energies were confirmed through vibrational frequency

calculations. The final optimized complexes were used to identify the possible

intermolecular hydrogen bond formation.

5.3.10 Preparation of formulated powders

Page 157: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

135

Different powder forms of IMC and all other excipients (Table 5.3), totaling

approximately 5 g, were added to a 30 mL plastic bottle, which was manually rotated about

the long axis at an angle of 45o to the horizon for 15 min (approximately 900 turns). A

tablet formulation containing 46.3 wt% IMC in the same matrix was also prepared as a

control. Avicel PH102 was used as the dry binder in all three formulations due to its high

plasticity, acceptable flowability, and excellent tabletability.34, 35 CCM-Na and MgSt

functioned as a tablet super-disintegrant and a lubricant, respectively.

5.3.11 Powder Flowability

Powder flowability was measured in triplicate using a ring shear tester (RST-XS;

Dietmar Schulze, Wolfenbüttel, Germany), with a 10 mL cell at a 1 kPa pre-shear normal

stress. A yield locus was obtained by plotting maximum shear stress as a function of normal

stress (0.25, 0.40, 0.55, 0.70, and 0.85 kPa), from which unconfined yield strength (fc) and

major principal stress (σn) were obtained by drawing Mohr’s circles. The flowability index

was calculated as ffc = 𝜎𝑛

𝑓c .36, 37

5.3.12 Powder Compaction

A universal material test machine (Model 1485, Zwick/Roell, Germany) was used

to prepare tablets at compaction pressures ranging from 25–350 MPa at a tableting speed

of 4 mm/min. The punch tip and die wall were lubricated using a 5% (w/v) suspension of

MgSt in ethanol, which was air-dried with a fan for ∼1 min to complete dryness. Tablets

were relaxed under ambient environment for at least 24 h before being diametrically broken

using a texture analyzer (TA-XT2i; Texture Technologies Corporation, Scarsdale, NY).

Page 158: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

136

Tablet tensile strength was calculated from the breaking force and tablet dimensions

following the standard procedure.38

5.3.13 Expedited Friability

Tablet friability was determined by an expedited method to quantify the ability of

compressed tablets to resist wearing and abrasion.39 Coded tablets prepared under different

pressures were weighed, individually coded, and loaded into a friabilator (Model F2,

Pharma Alliance Group Inc., Santa Clarita, CA), which rotated at 25 rpm for 4 min. The

initial and final weight of each tablet was used to calculate percent weight loss, which was

plotted against compaction pressure. The compaction pressure corresponding to 0.8%

tablet weight loss was identified from the friability plot.

5.3.14 Dissolution performance

Tablet dissolution performance was determined in 250 mL pH 2.0 HCl(aq) at 37 ±

0.5 °C, which was paddle stirred at 50 rpm. Tablets were prepared at 200 MPa, each

containing ~50 mg of IMC. Aliquots of the medium (3 mL) were taken at predetermined

time points of 1, 3, 5, 10, 20, 30, 40, 50 and 60 min and passed through a filter membrane

with a pore size of 0.45 μm. The IMC concentration was determined using a UV/Vis

spectrophotometer (DU 500, Beckman Instruments, Fullerton, CA) at = 327 nm to avoid

the interference from SAC using a calibration curve with appropriate background

correction.

Page 159: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

137

5.3.15 Solution 1H Nuclear Magnetic Resonance (NMR) Spectroscopy

1H NMR spectra were acquired on a Bruker Advance III HD 400 MHz spectrometer

in DMSO-d6 and referenced to the residual solvent resonance at δ 2.50 ppm. The pulse

repetition delay was set at 10 s in the acquisition spectra for HPMC-containing samples.

Anhydrous DMSO-d6 solvent was distilled from activated molecular sieves and stored in

an argon filled glovebox with [H2O] < 1 ppm. All NMR samples were prepared and sealed

in the glovebox to avoid water uptake during the characterization.

Statistical Analysis

To assess statistical significance, the two-way analysis of variance (ANOVA) was

performed. Difference between two data sets was deemed significant when p < 0.05.

5.4 Results and discussion

5.4.1 Micromeritic properties

For the IMC-SAC system, phase purity of the cocrystal was affected by both the

solvent ratio and relative amount of IMC and SAC. Hence, preliminary work was required

to identify suitable process parameters, i.e., 4 mmol of IMC and 5 mmol of SAC in 6 mL

of DMF poured into 40 mL of distilled water, to ensure phase purity. The use of HPMC

during spherical crystallization significantly affected particle morphology, whereby a

higher HPMC concentration led to larger size aggregates with more spherical shapes

(Figure 5.3), but excessively high HPMC concentrations led to process difficulties due to

Page 160: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

138

high viscosity. Therefore 0.5% (w/v) HPMC in the crystallization medium was used in the

following studies. The resulting QESD-CC particles were highly spherical with a narrower

particle size distribution than both as-received IMC and IMC-SAC prepared by the AS

process (Figure 5.4). In contrast, irregular IMC-SAC crystals were obtained when no

HPMC was present in the crystallization medium (Figure 5.5). The effect of added HPMC

is ascribed to its ability to stabilize the emulsions formed on mixing the drug solution with

water. In the absence of HPMC, the emulsions were unstable and easily broke down under

agitation, resulting in the production of fine individual crystals. Similar observations have

been reported with celecoxib,40 ketoprofen,41 and bucillamine42.

The SEM images revealed that primary IMC-SAC crystals prepared by the AS

process consisted of both small and large micro-size crystals (Figures 5.9a, b). This is

consistent with the broad PSD observed by laser scattering (Figure 5.4). Although large

spherical agglomerates were present in the QESD-CC powder (Figure 5.4c), primary

crystals in AS powder were much larger than those obtained by QESD-CC (Figure 5.9d).

The broad PSD of AS powder was attributed to the preparation method, wherein a methanol

co-solution of IMC and SAC was poured (in ≤ 10 s) into an equal volume of water. In this

process, the water to methanol volume ratio is initially high, leading to a high degree of

supersaturation of the IMC-SAC and instant precipitation of small micro-size crystals

precipitate quickly. With continuous addition of methanol, the degree of supersaturation is

lower than that at the beginning of the process because the more methanol-rich solvent can

dissolve more IMC and SAC. Therefore Consequently, precipitation occurs more slowly,

and the resulting crystals grow to larger sizes. However, during the QESD-CC process, the

primary crystals were small (~ 1 µm) because the fast counter-diffusion between DMF and

Page 161: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

139

water (v/v = 6: 40) led to a high degree of supersaturation of IMC-SAC inside the emulsion

droplets. Hence, fast nucleation limits the growth of individual primary crystals. The

smaller size of primary IMC-SAC crystals in the QESD-CC powder corresponds well with

the specific surface area (1.06 ±0.02 m2/g) that was approximately 2 times of that of IMC-

SAC via anti-solvent (0.59 ±0.09 m2/g). The spherical IMC-SAC particles obtained by

QESD-CC appeared hollow with openings visible at the surface (Figure 5.9c, e). This

hollow structure was confirmed by cutting a spherical cocrystal agglomerate to reveal the

interior (Figure 5.9f).

5.4.2 Phase purity

The PXRD patterns of both AS and QESD-CC powders matched well with that

calculated for IMC-SAC (Figure 5.6). This, along with the absence of characteristic peaks

of individual crystals of IMC and SAC (Figure 5.6), indicates the formation of phase pure

IMC-SAC cocrystals. The DSC thermograms of AS and QESD-CC powders (Figure 5.7)

also corresponded well with the reported profiles of IMC-SAC.27 The TGA profiles showed

negligible weight loss on increasing the temperature to 153 °C (the boiling point of DMF),

indicating the absence of any residual solvent in IMC-SAC prepared by both methods

(Figure 5.8). The amount of HPMC in QESD-CC powder prepared from a 0.5% HPMC

solution was 0.67 0.01% (n = 3).

5.4.3 Flowability and tabletability of various IMC forms

Both adequate flowability and tabletability of APIs are required for successfully

manufacturing a high drug loading tablet by the DC process. The flow property of a powder

Page 162: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

140

is determined by gravity and adhesion, where gravity favors powder flow while adhesion

hinders it.43, 44 In general, larger and more spherical particles with smoother surfaces

exhibit better flowability.45 Given the larger size and more spherical shape of the QESD-

CC particles as compared to those of AS and as-received IMC powders (Figure 5.4), the

flowability of QESD-CC powder was significantly better than other forms of IMC in this

study, as shown by the significantly higher ffc (p < 0.05) (Figure 5.10a). Importantly, the

flowability of the QESD-CC powder was significantly better than the Avicel PH102, which

marks the minimum acceptable powder flowability for high speed tableting.34 Hence, the

flowability of the QESD-CC powder does not require further improvement by

incorporating an excipient to sustain tablet manufacturing.

Both as-received IMC and AS powders showed poor tableting behavior (Figure

5.10b), which indicates inadequate plastic deformation that is required to develop

sufficiently large bonding areas between particles in the tablet.46 Therefore, crystal

plasticity was assessed using a combined energy framework and topological analysis to

gain structural insights.47 In the γ-IMC crystal, the visually identified slip layers parallel to

the (011) plane are energetically unfavored to slip due to strong interlayer interactions. On

the other hand, the energetically more favorable molecular planes parallel to the (001)

plane, recognized by an energy framework analysis, can undergo only very restricted slip

because of the rough layer surface topology (Figure 5.11a).29 In the IMC-SAC cocrystals,

SAC-SAC and IMC-IMC dimers are held together through N-H…O (2.859 Å) and O-H…O

(2.659 Å) hydrogen bonds, respectively. Those dimers assemble in 2D layers extending

along the (001) plane through O-H…O=C (3.005 Å) and O-H…O=S (2.964 Å) hydrogen

bonds. Only weak interactions between the neighboring layers, mediated by the edge to

Page 163: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

141

face π-π interactions (3.490 Å) and C-H…Cl (3.539 Å) hydrogen bonds were observed.

This is consistent with the energy framework, which suggests (001) as the most likely

primary slip plane in IMC-ACS (Figure 5.11b). However, such molecular layers are

interlocked, due to the rough topology of the layer surface, as shown by the negative

interlayer distance (-0.2 Å). The structural analyses above suggest that the poor plasticity

of both IMC and IMC-SAC can be attributed to the absence of active slip systems.

However, the QESD-CC powder exhibited profoundly improved tableting

performance relative to IMC-SAC (p < 0.05), even though they are the same crystal

polymorph. The highest tensile strength of ~ 3.6 MPa was approximately 6 times that of

AS (~ 0.6 MPa) (Figure 5.10b). At 150 MPa compaction pressure, the tablet tensile strength

was already higher than 2 MPa, suggesting adequate tablet mechanical strength for

handling and processing under a relatively mild compaction pressure.48 At least two factors

can contribute to the improved tabletability: 1) coating of the crystals by HPMC and 2)

much smaller primary crystal size. Surface coating is an effective particle engineering

strategy to improve tabletability by increasing bonding area. Therefore, with as little as

0.67% HPMC, tabletability significantly improves.19, 49, 50 Smaller primary crystal size

corresponds to larger surface area available for bonding and hence favors stronger tablets.51

We excluded the fracture of large QESD-CC particles as a major factor, because the

tabletability of the QESD-CC powder after milling for 5 min remained nearly identical

below 250 MPa (Figure 5.10b).

5.4.4 Formulation development

Page 164: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

142

The excellent flow and tableting properties of QESD-CC are expected to enable the

formulation development of DC tablet with a high IMC loading. The blend of components

in a successful DC tablet formulation must be sufficiently free flowing to make tablets with

consistent weight and uniform drug content during high-speed tablet manufacturing. In this

regard, the QESD-CC formulation meets the minimum flowability criterion, since it

exhibits better flowability than Avicel PH102, while IMC and AS formulations do not

(Figure 5.12a). This result highlights the important role of the API on the flowability of

high drug loading formulations.

The tableting performance of formulations followed the same order as that of API

powders, i.e., QESD-CC > as received IMC > AS. When the compaction pressure was

higher than 200 MPa, the tablet tensile strength for all formulations exceeded 2 MPa

(Figure 5.12b), despite both as-received IMC and AS powders exhibiting poor tabletability

with maximum tensile strengths of approximately 0.6 and 0.5 MPa, respectively (Figure

5.10b). The improved tabletability indicates that the presence of plastic Avicel PH102

(24.5 wt% in cocrystal formulations and 48.2 wt% in IMC formulation) could overcome

the poor tabletability of API in both formulations. This is attributed to the excellent

tabletability of Avicel PH102.52 Considering the otherwise identical composition between

QESD-CC and AS formulations (Table 1), the better tabletability of QESD-CC formulation

than AS formulation suggests that the tableting performance of API also influences

tabletability of high dose DC tablet formulations (Figure 5.10b).

Friability characterizes the propensity of a tablet to lose mass when exposed to

external impact during downstream processing, shipping, and handling.53 The friability of

IMC and QESD-CC formulations met the friability requirement that tablets compressed at

Page 165: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

143

compaction pressures higher than 150 MPa and 100 MPa, respectively, exhibited weight

loss of ≤ 0.8%. The tablet friability of AS formulation remained higher than 0.8% in the

entire range of compaction pressure (Figure 5.12c). Therefore, it is unlikely to be a

successful DC tablet formulation.

5.4.5 Dissolution

The final remaining consideration for assessing the suitability of the DC process is

dissolution. Tablets containing 50 mg of as-received IMC or an equivalent amount of

QESD-CC (75.6 mg) were prepared at a compaction pressure of 200 MPa, and their

dissolution performance was determined in simulated gastric fluid (250 mL of pH 2.0 HCl

aqueous solution). The 200 MPa compaction pressure was used because tablets from both

formulations exhibited sufficient mechanical strength, as shown by > 2 MPa tensile

strength (Figure 5.12b) and < 0.8% friability (Figure 5.12c). As expected, the formulated

QESD-CC showed much improved drug release compared to IMC, with the area under

curve of QESD-CC being ~9-fold higher than IMC (Figure 5.12d). Thus, the dissolution

data confirms that the QESD-CC approach can indeed be used to enable the development

of high dose DC tablets of IMC.

5.4.6 Drug-polymer molecular interaction

To better understand the mechanism in formation of the spherical cocrystal

agglomerates, we further probed the underlying intermolecular interactions between

HPMC and IMC/SAC molecules by solution 1H NMR. To mimic the aqueous environment

in the QESD process, deuterium oxide would be the ideal solvent to use. However, the poor

Page 166: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

144

water solubility of IMC led us to characterize these interactions in DMSO-d6 in order to

co-dissolve all three components involved in this process.

When the concentration of IMC-SAC in DMSO-d6 increased from 2 to 8 mg/mL,

the peaks positions correlated with IMC at 7.60–7.70 ppm did not change (Figure 5.13a).

However, downfield shifts of 1H NMR spectral peaks were observed in the range 7.80–

8.20 ppm (Figure 5.13a), which correspond to the H atom resonances on the aromatic ring

of SAC (Figures 1b). The downfield shift indicates decreased electron density in the

aromatic ring, which is induced by a stronger hydrogen bonding between adjacent SAC

molecules through amide moieties, which is analogous to the amide dimer (N-H...O;

2.859(3) Å) synthon (R22 (8) motif) formed between two SAC molecules in the IMC-SAC

crystal structure as depicted in Figure 5.13c. With the addition of 0.5 mg/mL HPMC to a

2 mg/mL IMC-SAC solution, an upfield shift of these peaks is observed that suggests

weakening of the N-H...O H-bond between SAC dimers induced by HPMC (Figure 5.13a).

Additionally, the peaks corresponding to the hydroxyl substituents on HPMC broaden in

the presence of IMC-SAC (Figure 5.13b), indicating conformational restriction of the –OH

side chains of the HPMC, likely due to their interactions with the SAC molecules. Similar

peak shifts and peak smearing in 1H NMR spectra of SAC and HPMC were observed in

the absence of IMC (Figure 5.14). These observations suggest that the hydroxyl groups on

HPMC form hydrogen bonds with SAC, which weaken the original strong intermolecular

N-H...O hydrogen bonds between SAC molecules while IMC is not directly involved. Since

the 1H NMR spectra of IMC remained the same in the presence and absence of HPMC

(Figure 5.13a), IMC does not significantly interact with HPMC.

Page 167: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

145

The preferential interaction of HPMC with SAC instead of IMC was further

investigated using DFT computational simulations, which revealed hydrogen bonds

formed between SAC and HPMC monomer units through S=O…H-O (2.889 Å), N-H…O

(2.704 Å, 2.832 Å), and C=O…H-O (2.775 Å) interactions (Figure 5.15). In comparison,

no hydrogen bonds between HPMC and IMC were suggested by the simulations. These

results corroborate the intermolecular interactions inferred from the 1H NMR data (Figures

7 and S6), i.e., the dominant intermolecular interactions between HPMC and IMC-SAC

are hydrogen bonds between SAC and HPMC.

5.4.7 Formation mechanism of spherical cocrystal agglomerates

The observed hollow structure of the QESD-CC particles (Figures 3e,f) provide

some insights into a possible formation mechanism. Emulsion droplets first form on

mixing the concentrated drug solution and aqueous medium. HPMC in the aqueous

medium interacts with SAC at the organic-aqueous interface through the hydrogen bonds

described above to stabilize the emulsion (Figure 5.16c). The counter diffusion of water

and DMF drives a high degree of supersaturation near the drop surface and results in the

precipitation of very fine IMC-SAC crystals to form a crust. As DMF diffuses into the

aqueous phase, interfacially-adsorbed HPMC precipitates and coats the surface of the

particles because of the poorer solubility of HPMC in the DMF-rich solvent mixture

adjacent to the particle surface. The crust grows inward as the solvent counter diffusion

continue until cocrystallization is complete. This process naturally leads to particles with a

core-shell structure with a HPMC coating. Upon drying, the entrapped solvent escapes

through pores in the shell by evaporation. This particle formation mechanism is consistent

Page 168: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

146

with the greatly enhanced tabletability, which is possible if HPMC coats the QESD-CC

particles instead of distributing throughout the bulk. Additionally, such hollow particles

are more deformable during compression, which maximizes the interparticle contact area

and favors higher tablet strength.

5.5 Conclusion

We have designed and successfully applied a novel quasi-emulsion solvent

diffusion cocrystallization (QESD-CC) method to simultaneously overcome poor

flowability, poor tabletability, and slow dissolution of IMC. HPMC stabilizes the emulsion

through hydrogen bond interactions with saccharin molecules at the emulsion surface.

Subsequent solvent counter-diffusion leads to the initiation of the IMC-SAC crystallization

at the emulsion drop surface. The cocrystal grows inward to form a core-shell structure,

resulting in round shape and large size of QESD-CC particles, which exhibits excellent

powder flowability. The HPMC coating of particle surface and much smaller size of

primary IMC-SAC crystals account for significantly improved tabletability. The improved

flowability and tabletability enable the successful manufacturing of a high IMC loading

(46.3 wt%) tablet formulation using the direct compression process. Tablets also exhibited

excellent dissolution due to the much higher solubility of the IMC-SAC cocrystal than IMC.

Thus, for poorly soluble drugs that also exhibit poor flowability the QESD-CC approach

is a potentially highly effective and efficient engineering strategy to enable the successful

manufacturing of high drug loading tablets using the most economical direct compression

process.

Page 169: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

147

Table 5.1. N2 physisorption isotherms of QESD-CC and AS.

Isotherm Linear Plot Isotherm Linear Plot Isotherm Linear Plot

QESD-CC - Adsorption QESD-CC - Adsorption QESD-CC - Adsorption

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

0.002296 0.069851 0.002296 0.069646 0.002281 0.069826

0.022375 0.147844 0.022316 0.14575 0.022187 0.145737

0.046561 0.183815 0.046806 0.18157 0.046495 0.181907

0.070131 0.207943 0.070155 0.204569 0.070135 0.204573

0.081668 0.218059 0.081686 0.21384 0.081621 0.214242

0.102193 0.233145 0.102187 0.227887 0.10212 0.228296

0.122601 0.246278 0.122938 0.241759 0.122505 0.241146

0.142968 0.258061 0.142931 0.252569 0.142904 0.253097

0.163739 0.26979 0.163227 0.263408 0.163085 0.263622

0.184183 0.280032 0.183657 0.273268 0.183584 0.273683

0.203991 0.290134 0.20411 0.283246 0.204478 0.28405

0.224358 0.300368 0.224472 0.293123 0.224292 0.293487

0.245248 0.310374 0.244644 0.301945 0.245266 0.302919

0.265213 0.319642 0.265231 0.311345 0.2652 0.312214

0.286282 0.328656 0.285598 0.320121 0.285397 0.320344

0.306039 0.338387 0.306036 0.328738 0.306508 0.32901

0.326516 0.347864 0.326506 0.337448 0.32691 0.337886

0.336741 0.352942 0.336692 0.342958 0.336288 0.341999

0.347647 0.357827 0.347909 0.348286 0.34704 0.347842

0.357199 0.362346 0.357178 0.351921 0.356894 0.35175

0.368029 0.367752 0.3685 0.356771 0.367714 0.356661

0.378798 0.372597 0.377592 0.360134 0.377224 0.361035

0.387892 0.37639 0.388289 0.365577 0.388018 0.36518

0.40873 0.385813 0.408918 0.374844 0.408534 0.373417

0.414393 0.389464 0.413323 0.377178 0.412915 0.37598

0.418519 0.391566 0.419062 0.380089 0.419161 0.379727

0.429687 0.396734 0.428786 0.384686 0.428217 0.384009

0.450394 0.407119 0.450171 0.39381 0.449371 0.392946

0.470993 0.417112 0.470718 0.40274 0.469748 0.402941

0.491271 0.426844 0.490848 0.412567 0.490219 0.412714

0.511432 0.43714 0.511013 0.422675 0.510421 0.422718

0.53203 0.448404 0.53208 0.433204 0.531203 0.43305

0.552499 0.460812 0.552211 0.444525 0.551266 0.443542

0.57297 0.472303 0.573233 0.456581 0.572096 0.45527

0.593603 0.485822 0.593186 0.468368 0.592108 0.466559

Page 170: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

148

0.613509 0.498563 0.613559 0.480781 0.61272 0.47981

0.634472 0.513603 0.63402 0.494322 0.633004 0.493226

0.654273 0.527648 0.654546 0.50839 0.653725 0.508163

0.67516 0.544407 0.674924 0.522751 0.673576 0.522959

0.695657 0.562899 0.695384 0.539559 0.694159 0.540075

0.716591 0.583136 0.715858 0.55729 0.714751 0.557629

0.736194 0.601777 0.736317 0.576457 0.735105 0.575719

0.756995 0.625864 0.756818 0.598681 0.755336 0.59691

0.777643 0.652227 0.777319 0.622691 0.775761 0.620937

0.79828 0.681685 0.797711 0.648934 0.796186 0.646323

0.818472 0.715554 0.818197 0.679693 0.816603 0.6772

0.839022 0.756604 0.838621 0.717066 0.837045 0.711893

0.859762 0.806147 0.859068 0.76072 0.85736 0.753309

0.880087 0.86835 0.879387 0.814854 0.877765 0.806106

0.898693 0.939791 0.899984 0.887515 0.898112 0.875111

0.918913 1.04532 0.919252 0.979711 0.919113 0.972597

0.929333 1.12184 0.929266 1.04729 0.929174 1.03899

0.939494 1.21368 0.939464 1.13227 0.939148 1.12185

0.949578 1.33149 0.949502 1.24139 0.949271 1.23216

0.959799 1.49198 0.95964 1.39317 0.959342 1.38419

0.969851 1.74645 0.969489 1.61935 0.969118 1.61627

0.975408 1.95214 0.975097 1.82493 0.97463 1.82567

0.980154 2.19513 0.979947 2.07792 0.979486 2.08368

0.985208 2.56209 0.984938 2.45054 0.984377 2.45561

0.987212 2.75546 0.986894 2.64781 0.986185 2.65953

0.988115 2.88251 0.987768 2.77777 0.98738 2.81067

0.989908 3.10941 0.98967 3.00986 0.988255 2.95666

0.99048 3.21658 0.989861 3.09908 0.989791 3.18788

AS - Adsorption AS - Adsorption AS - Adsorption

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

Relative

Pressure

(P/Po)

Quantity

Adsorbed

(cm³/g STP)

0.002679 0.055584 0.002452 0.054007 0.002418 0.05529

0.023056 0.104172 0.02297 0.109396 0.022999 0.109689

0.046877 0.122074 0.046795 0.129999 0.047113 0.13236

0.07039 0.131833 0.070327 0.143343 0.070344 0.145297

0.081743 0.135894 0.081695 0.147475 0.081706 0.151092

0.102224 0.139538 0.10205 0.155689 0.102456 0.15929

0.122552 0.144073 0.122506 0.162061 0.122525 0.165033

0.1427 0.147478 0.143025 0.168968 0.142892 0.172262

0.16332 0.149634 0.162957 0.173974 0.16349 0.177031

0.183975 0.15207 0.183566 0.179952 0.183545 0.183036

Page 171: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

149

0.204047 0.153155 0.203962 0.182327 0.203888 0.187285

0.224334 0.154635 0.224392 0.186953 0.22425 0.192312

0.24468 0.155088 0.244553 0.191999 0.24507 0.197801

0.265542 0.156772 0.26491 0.197141 0.264991 0.201585

0.286158 0.158076 0.285379 0.199859 0.285345 0.205597

0.306566 0.159003 0.305857 0.203564 0.305787 0.208924

0.32665 0.160649 0.326134 0.207414 0.326227 0.212992

0.337173 0.162431 0.336094 0.209914 0.336332 0.215656

0.347556 0.161867 0.346295 0.213355 0.346551 0.217778

0.358166 0.162598 0.357809 0.215121 0.356852 0.220446

0.367329 0.162145 0.366843 0.217135 0.367734 0.222817

0.37746 0.162912 0.376951 0.219714 0.377204 0.225516

0.38769 0.163321 0.387211 0.220861 0.387934 0.226642

0.408196 0.164739 0.408422 0.225035 0.408785 0.231052

0.413187 0.166265 0.412518 0.225514 0.413585 0.23249

0.419097 0.166357 0.417555 0.228224 0.417904 0.232998

0.42919 0.165325 0.427824 0.2284 0.429121 0.234702

0.449677 0.165296 0.448978 0.232209 0.449731 0.239752

0.469591 0.164869 0.469744 0.236762 0.470189 0.244506

0.491019 0.167629 0.48978 0.242854 0.490292 0.250302

0.511465 0.170625 0.510699 0.246697 0.510767 0.256266

0.531661 0.176543 0.530608 0.251463 0.531177 0.261243

0.552331 0.179419 0.551186 0.257633 0.551702 0.266657

0.572461 0.180196 0.571215 0.265656 0.572067 0.272858

0.593085 0.183683 0.592147 0.274354 0.592525 0.28151

0.613235 0.191017 0.612276 0.282425 0.612612 0.291236

0.633655 0.199402 0.632175 0.292908 0.632882 0.303399

0.65445 0.211224 0.652599 0.307963 0.653408 0.317425

0.674476 0.22465 0.672971 0.323195 0.673658 0.33376

0.694828 0.23669 0.693627 0.336891 0.693889 0.350482

0.715339 0.246493 0.713833 0.350001 0.714526 0.364091

0.735824 0.253985 0.734059 0.362725 0.735175 0.37287

0.756085 0.260854 0.754376 0.374666 0.755518 0.38148

0.776886 0.267156 0.774785 0.386196 0.775812 0.391792

0.79708 0.277895 0.795053 0.394377 0.795785 0.404317

0.817876 0.287299 0.814938 0.40609 0.816494 0.418888

0.837968 0.297879 0.83574 0.424878 0.836236 0.435911

0.858539 0.307645 0.858336 0.439 0.857317 0.450653

0.878889 0.330234 0.878322 0.458573 0.878352 0.472247

0.898619 0.353675 0.898834 0.486355 0.898904 0.499065

0.919382 0.38337 0.9192 0.520258 0.919286 0.535408

0.929355 0.400002 0.928999 0.542827 0.929121 0.559275

0.939411 0.422752 0.939266 0.573748 0.939492 0.587954

Page 172: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

150

0.949562 0.456087 0.949234 0.612801 0.949777 0.62432

0.9597 0.501805 0.959724 0.661852 0.959494 0.678726

0.969364 0.572115 0.969594 0.756994 0.969203 0.77624

0.975016 0.632246 0.974794 0.832113 0.974663 0.849678

0.980069 0.705292 0.979852 0.923025 0.979797 0.943283

0.985482 0.809309 0.98496 1.03842 0.985078 1.06793

0.986783 0.840634 0.986424 1.08499 0.986087 1.10232

0.987817 0.86837 0.987736 1.13276 0.987159 1.13736

0.989104 0.903081 0.988809 1.168 0.98872 1.1909

0.990405 0.944236 0.989884 1.20401 0.989824 1.23648

Page 173: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

151

Figure 5.1. Correlation between the actual measured HPMC concentration by the SEC

method.

Page 174: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

152

Table 5.2. Intermolecular interaction energies estimated using B3LYP+D2/6+31G(d,p)

dispersion corrected DFT model. Both the total energy (E(tot)) and electrostatic (E(ele)),

polarization (E(pol)), dispersion (E(dis)), and exchange repulsion (E(rep)) components of the

energy are listed. R indicated the distance between centers of mass of the pair of molecules.

Symop R Eele Epol Edis Erep Etot

IMC γ Form

-x, -y, -z 6.25 -20.1 -1.6 -79.2 54.6 -57.7

x, y, z 9.24 -7.6 -0.2 -23.3 16.9 -18.0

-x, -y, -z 6.22 -19.6 -6.3 -49.3 34.2 -47.2

x, y, z 13.10 0.0 -2.9 -11.8 6.7 -8.3

-x, -y, -z 12.54 -123.6 -0.6 -13.8 144.1 -54.1

-x, -y, -z 8.57 -7.0 -5.2 -49.3 29.2 -36.1

-x, -y, -z 9.44 -6.9 -0.8 -41.2 27.5 -26.7

-x, -y, -z 7.98 -5.9 -0.4 -18.4 9.9 -16.5

-x, -y, -z 8.52 0.5 -0.5 -20.6 10.5 -11.4

x, y, z 11.79 -6.3 -0.6 -9.2 7.2 -10.6

-x, -y, -z 15.21 -0.6 -0.2 -3.0 2.4 -1.9

IMCSAC

x, y, z 7.11 -9.8 -3.1 -45.2 33.3 -31.4

-x, -y, -z 12.58 -117.4 -27.2 -14.0 141.9 -68.8

-x, -y, -z 8.34 -8.9 -1.8 -31.4 24.6 -22.9

-x, -y, -z 10.27 0.3 -8.2 -11.1 0.9 -14.9

- 4.97 -27.7 -0.7 -57.5 44.8 -52.1

- 7.69 -3.0 -1.2 -5.8 1.5 -8.3

-x, -y, -z 6.05 -17.9 0.0 -56.4 38.9 -44.0

x, y, z 11.45 -0.3 -4.4 -8.4 4.7 -7.9

- 8.59 -20.9 -0.9 -24.5 24.0 -29.3

- 7.99 -5.6 -2.8 -14.5 11.9 -13.3

Page 175: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

153

-x, -y, -z 11.82 -2.4 -1.2 -12.9 6.9 -10.3

- 10.95 3.6 -0.7 -4.0 2.2 1.2

- 12.32 3.1 -0.9 -2.0 0.1 1.0

-x, -y, -z 12.11 -3.3 -1.5 -19.5 14.7 -12.5

- 11.71 1.2 -0.4 -2.4 1.1 -0.4

- 9.42 -9.7 -2.9 -10.3 9.5 -15.5

- 10.46 -3.3 -0.0 -7.0 8.6 -4.3

- 10.06 0.1 -0.2 -1.6 0.0 -1.3

- 10.06 0.1 -0.2 -1.6 0.0 -1.3

- 8.59 -20.9 -5.2 -24.5 24.0 -32.5

- 10.95 3.6 -0.7 -4.0 2.2 1.2

- 7.69 -3.0 -1.2 -5.8 1.5 -8.3

-x, -y, -z 6.99 -69.7 -13.5 -12.5 65.5 -54.1

-x, -y, -z 4.47 -9.4 -2.1 -41.7 28.6 -30.2

- 4.97 -27.7 -6.3 -57.5 44.8 -56.3

- 11.71 1.2 -0.4 -2.4 1.1 -0.4

x, y, z 7.11 -4.4 -0.9 -8.9 1.9 -11.9

- 7.99 -5.6 -2.8 -14.5 11.9 -13.3

- 12.32 3.1 -0.4 -2.0 0.1 1.3

- 9.42 -9.7 -2.9 -10.3 9.5 -15.5

- 10.46 -3.3 -0.6 -7.0 8.6 -4.7

Page 176: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

154

Figure 5.2. Chemical structures of (a) IMC, (b) SAC and (c) HPMC. Hydrogen atoms that

participate in intermolecular interactions responsible for QESD-CC are marked with

different colors in accordance to their chemical shifts in solution 1H NMR spectra.

(a) (b) (c)

Page 177: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

155

Table 5.3. Formulation composition of IMC-containing tablets using different IMC forms

Material Function IMC AS QESD-CC

IMC/IMC-SAC API 46.3 % 70 % 70 %

Avicel PH102 Binder 48.2 % 24.5 % 24.5 %

CCM-Na Super-

disintegrant 5.0 % 5.0 % 5.0 %

MgSt Lubricant 0.5 % 0.5 % 0.5 %

Total 100 % 100 % 100 %

Page 178: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

156

Figure 5.3. Optical microscopy images of particles prepared with aqueous solutions of

HPMC in different concentrations from 0% to 0.5% (scale bar = 500 µm).

0.01 % HPMC

0.1 % HPMC 0.5 % HPMC

0 % HPMC

Page 179: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

157

Figure 5.4. Particle size distribution of as-received IMC, IMC-SAC prepared by AS and

QESD-CC processes, indicating that the QESD-CC yields highly spherical particles with

a relatively narrower size distribution.

Page 180: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

158

Figure 5.5. Optical microscopy images of particles (a) with and (b) without 0.5% HPMC

in the crystallization medium (scale bar = 500 µm).

a) b)

Page 181: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

159

Figure 5.6. PXRD patterns of calculated IMC-SAC, AS, QESD-CC), as-received IMC,

and as-received SAC.

Page 182: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

160

Figure 5.7. DSC profiles of as received IMC, IMC-SAC prepared by AS, and QESD-CC

processes.

Page 183: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

161

Figure 5.8. TGA profiles of as received IMC, IMC-SAC prepared by AS and QESD-CC

processes.

Cocrystal melting point

Page 184: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

162

Figure 5.9. SEM images of AS (a and b) and QESD-CC (c and d) powders at low (x 50)

and high (x 20,000) magnifications. Images e) and f) show the hollow structure of QESD-

CC particles

(a) (b)

(e) (f)

(d) (c)

Page 185: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

163

Figure 5.10. Powder properties of as-received IMC, AS, and QESD-CC powders, a)

flowability and b) tabletability.

a)

Avicel PH102

b)

Page 186: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

164

Figure 5.11. Energy frameworks for (a) IMC and (b) IMC-SAC with a likely slip layer

shaded in pink. The thickness of each cylinder (blue) represents the relative strength of

intermolecular interaction. The energy threshold for the energy framework is set at −10

kJ/mol.

a) b)

Page 187: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

165

Figure 5.12. Performance of tablet formulations based on as-received IMC, AS, and

QESD-CC, (a) flowability, (b) tabletability, c) tablet friability, and d) tablet dissolution.

QESD-CC

IMC

AS

a) b)

Avicel PH102

QESD-CC

IMC

c) d)

Page 188: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

166

Figure 5.13. Partial 1H NMR spectra of (a) 2 mg/mL IMC-SAC, 2 mg/mL IMC-SAC with

0.5 mg/mL HPMC and 8 mg/mL IMC-SAC, and (b) 0.5 mg/mL HPMC with and without

2 mg/mL of IMC-SAC, (c) Intermolecular hydrogen bonding interactions in IMC-SAC

cocrystal. Each group of peaks corresponds to H atoms in Figure 5.2 labelled by the dots

of the same color.

a) b)

c)

Page 189: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

167

Figure 5.14. Partial 1H NMR spectra of (a) 2 mg/mL SAC, 8 mg/mL SAC and 2 mg/mL

SAC with 0.5 mg/mL HPMC, and (b) 0.5 mg/mL HPMC and 2 mg/mL SAC with 0.5

mg/mL HPMC.

a) b)

Page 190: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

168

Figure 5.15. Optimized geometric complexes between IMC-SAC and three HPMC

monomer units with different substituents: a) -H, b) -CH3, c) -CH2CH(OH)CH3.

a) b)

c)

Page 191: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

169

Figure 5.16. Formation process of QESD-CC particles.

HPMC Emulsion droplet Cocrystal

Water phase

Oil phase

Page 192: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

170

5.6 References

1. Van Snick, B.; Holman, J.; Cunningham, C.; Kumar, A.; Vercruysse, J.; De Beer,

T.; Remon, J. P.; Vervaet, C. Continuous direct compression as manufacturing platform

for sustained release tablets. Int J. Pharm. 2017, 519, 390-407.

2. Jivraj, M.; Martini, L. G.; Thomson, C. M. An overview of the different excipients

useful for the direct compression of tablets. Pharm. Sci. Technol. Today 2000, 3, 58-63.

3. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

4. Chen, L.; Ding, X.; He, Z.; Huang, Z.; Kunnath, K. T.; Zheng, K.; Davé, R. N.

Surface engineered excipients: I. Improved functional properties of fine grade

microcrystalline cellulose. Int. J. Pharm. 2018, 536, 127-137.

5. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science (New York,

N.Y.) 1982, 216, 1127-1128.

6. Chen, H.; Wang, C.; Sun, C. C. Profoundly improved plasticity and tabletability

of griseofulvin by in situ solvation and desolvation during spherical crystallization. Cryst.

Growth Des. 2019, 19, 2350-2357.

7. Peña, R.; Nagy, Z. K. Process intensification through continuous spherical

crystallization using a two-stage mixed suspension mixed product removal (msmpr) system.

Cryst. Growth Des. 2015, 15, 4225-4236.

8. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

9. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

10. Chen, H.; Wang, C.; Kang, H.; Zhi, B.; Haynes, C. L.; Aburub, A.; Sun, C. C.

Microstructures and pharmaceutical properties of ferulic acid agglomerates prepared by

different spherical crystallization methods. Int. J. Pharm. 2020, 574, 118914.

11. Saboo, S.; Mugheirbi, N. A.; Zemlyanov, D. Y.; Kestur, U. S.; Taylor, L. S.

Congruent release of drug and polymer: A “sweet spot” in the dissolution of amorphous

solid dispersions. J. Control. Release 2019, 298, 68-82.

12. Serajuddin, A. T. Salt formation to improve drug solubility. Adv. Drug Deliv. Rev.

2007, 59, 603-616.

13. Blagden, N.; de Matas, M.; Gavan, P. T.; York, P. Crystal engineering of active

pharmaceutical ingredients to improve solubility and dissolution rates. Adv. Drug Deliv.

Rev 2007, 59, 617-630.

14. Han, R.; Xiong, H.; Ye, Z.; Yang, Y.; Huang, T.; Jing, Q.; Lu, J.; Pan, H.; Ren, F.;

Ouyang, D. Predicting physical stability of solid dispersions by machine learning

techniques. J. Control. Release 2019, 311-312, 16-25.

Page 193: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

171

15. Indulkar, A. S.; Lou, X.; Zhang, G. G. Z.; Taylor, L. S. Insights into the dissolution

mechanism of ritonavir–copovidone amorphous solid dispersions: Importance of

congruent release for enhanced performance. Mol. Pharm. 2019, 16, 1327-1339.

16. Zheng, K.; Lin, Z.; Capece, M.; Kunnath, K.; Chen, L.; Davé, R. N. Effect of

particle size and polymer loading on dissolution behavior of amorphous griseofulvin

powder. J. Pharm. Sci. 2019, 108, 234-242.

17. Aitipamula, S.; Banerjee, R.; Bansal, A. K.; Biradha, K.; Cheney, M. L.;

Choudhury, A. R.; Desiraju, G. R.; Dikundwar, A. G.; Dubey, R.; Duggirala, N.; Ghogale,

P. P.; Ghosh, S.; Goswami, P. K.; Goud, N. R.; Jetti, R. R. K. R.; Karpinski, P.; Kaushik,

P.; Kumar, D.; Kumar, V.; Moulton, B.; Mukherjee, A.; Mukherjee, G.; Myerson, A. S.;

Puri, V.; Ramanan, A.; Rajamannar, T.; Reddy, C. M.; Rodriguez-Hornedo, N.; Rogers, R.

D.; Row, T. N. G.; Sanphui, P.; Shan, N.; Shete, G.; Singh, A.; Sun, C. C.; Swift, J. A.;

Thaimattam, R.; Thakur, T. S.; Kumar Thaper, R.; Thomas, S. P.; Tothadi, S.; Vangala, V.

R.; Variankaval, N.; Vishweshwar, P.; Weyna, D. R.; Zaworotko, M. J. Polymorphs, salts,

and cocrystals: What’s in a name? Crys. Growth Des. 2012, 12, 2147-2152.

18. Drozd, K. V.; Manin, A. N.; Churakov, A. V.; Perlovich, G. L. Novel drug–drug

cocrystals of carbamazepine with para-aminosalicylic acid: Screening, crystal structures

and comparative study of carbamazepine cocrystal formation thermodynamics.

CrystEngComm 2017, 19, 4273-4286.

19. Chen, H.; Guo, Y.; Wang, C.; Dun, J.; Sun, C. C. Spherical cocrystallization—an

enabling technology for the development of high dose direct compression tablets of poorly

soluble drugs. Cryst. Growth Des. 2019, 19, 2503-2510.

20. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

21. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

22. Kakran, M.; Sahoo, N. G.; Tan, I.-L.; Li, L. Preparation of nanoparticles of poorly

water-soluble antioxidant curcumin by antisolvent precipitation methods. J. Nanopart. Res.

2012, 14, 757.

23. Chun, N.-H.; Wang, I.-C.; Lee, M.-J.; Jung, Y.-T.; Lee, S.; Kim, W.-S.; Choi, G. J.

Characteristics of indomethacin–saccharin (imc–sac) co-crystals prepared by an anti-

solvent crystallization process. Eur. J. Pharm. Biopharm. 2013, 85, 854-861.

24. Alhalaweh, A.; Roy, L.; Rodríguez-Hornedo, N.; Velaga, S. P. Ph-dependent

solubility of indomethacin–saccharin and carbamazepine–saccharin cocrystals in aqueous

media. Mol. Pharm. 2012, 9, 2605-2612.

25. Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of gases in multimolecular

layers. J. Am. Chem. Soc. 1938, 60, 309-319.

26. Chen, M.-H.; Bergman, C. J. Method for determining the amylose content,

molecular weights, and weight- and molar-based distributions of degree of polymerization

of amylose and fine-structure of amylopectin. Carbohydr. Polym. 2007, 69, 562-578.

Page 194: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

172

27. Basavoju, S.; Boström, D.; Velaga, S. P. Indomethacin–saccharin cocrystal:

Design, synthesis and preliminary pharmaceutical characterization. Pharm. Res. 2008, 25,

530-541.

28. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

29. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

30. Turner, M. J.; Grabowsky, S.; Jayatilaka, D.; Spackman, M. A. Accurate and

efficient model energies for exploring intermolecular interactions in molecular crystals. J.

Phys. Chem. Lett. 2014, 5, 4249-4255.

31. Bryant, M. J.; Maloney, A. G. P.; Sykes, R. A. Predicting mechanical properties

of crystalline materials through topological analysis. CrystEngComm 2018, 20, 2698-2704.

32. Gaussian 16, revision b.01, m. J. Frisch, g. W. Trucks, h. B. Schlegel, g. E. Scuseria,

m. A. Robb, j. R. Cheeseman, g. Scalmani, v. Barone, g. A. Petersson, h. Nakatsuji, x. Li,

m. Caricato, a. V. Marenich, j. Bloino, b. G. Janesko, r. Gomperts, b. Mennucci, h. P.

Hratchian, j. V. Ortiz, a. F. Izmaylov, j. L. Sonnenberg, d. Williams-young, f. Ding, f.

Lipparini, f. Egidi, j. Goings, b. Peng, a. Petrone, t. Henderson, d. Ranasinghe, v. G.

Zakrzewski, j. Gao, n. Rega, g. Zheng, w. Liang, m. Hada, m. Ehara, k. Toyota, r. Fukuda,

j. Hasegawa, m. Ishida, t. Nakajima, y. Honda, o. Kitao, h. Nakai, t. Vreven, k. Throssell,

j. A. Montgomery, jr., j. E. Peralta, f. Ogliaro, m. J. Bearpark, j. J. Heyd, e. N. Brothers, k.

N. Kudin, v. N. Staroverov, t. A. Keith, r. Kobayashi, j. Normand, k. Raghavachari, a. P.

Rendell, j. C. Burant, s. S. Iyengar, j. Tomasi, m. Cossi, j. M. Millam, m. Klene, c. Adamo,

r. Cammi, j. W. Ochterski, r. L. Martin, k. Morokuma, o. Farkas, j. B. Foresman, and d. J.

Fox, gaussian, inc., wallingford ct, 2016.

33. Zhao, Y.; Truhlar, D. G. J. T. C. A. The m06 suite of density functionals for main

group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states,

and transition elements: Two new functionals and systematic testing of four m06-class

functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215-241.

34. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

35. Chen, L.; Ding, X.; He, Z.; Fan, S.; Kunnath, K. T.; Zheng, K.; Davé, R. N. Surface

engineered excipients: Ii. Simultaneous milling and dry coating for preparation of fine-

grade microcrystalline cellulose with enhanced properties. Int. J. Pharm. 2018, 546, 125-

136.

36. Shi, L.; Chattoraj, S.; Sun, C. C. Reproducibility of flow properties of

microcrystalline cellulose — avicel ph102. Powder Technol. 2011, 212, 253-257.

37. Schulze, D. Powders and bulk solids: Behavior, characterization, storage and flow,

springer, berlin, heidelberg, new york, tokyo, 2008.

38. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

Page 195: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

173

39. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int. J. Pharm. 2015, 484, 146-155.

40. Chen, H.; Paul, S.; Xu, H.; Wang, K.; Mahanthappa, M. K.; Sun, C. C. Reduction

of punch-sticking propensity of celecoxib by spherical crystallization via polymer assisted

quasi-emulsion solvent diffusion. Mol. Pharm. 2020, 1387–1396.

41. Ribardière, A.; Tchoreloff, P.; Couarraze, G.; Puisieux, F. Modification of

ketoprofen bead structure produced by the spherical crystallization technique with a two-

solvent system. Int. J. Pharm. 1996, 144, 195-207.

42. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

43. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

44. Kendall, K. Adhesion: Molecules and mechanics. Science (New York, N.Y.) 1994,

263, 1720-1725.

45. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

46. Sun, C. C.; Kleinebudde, P. Mini review: Mechanisms to the loss of tabletability

by dry granulation. Eur. J. Pharm. Biopharm. 2016, 106, 9-14.

47. Wang, C.; Sun, C. C. Computational techniques for predicting mechanical

properties of organic crystals: A systematic evaluation. Mol. Pharm. 2019, 16, 1732-1741.

48. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

49. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

50. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

51. Shi, L.; Feng, Y.; Sun, C. C. Roles of granule size in over-granulation during high

shear wet granulation. J. Pharm. Sci. 2010, 99, 3322-3325.

52. Zhang, Y.; Law, Y.; Chakrabarti, S. Physical properties and compact analysis of

commonly used direct compression binders. AAPS PharmSciTech. 2003, 4, 489-499.

53. Sinka, I. C.; Cunningham, J. C.; Zavaliangos, A. Analysis of tablet compaction. Ii.

Finite element analysis of density distributions in convex tablets. J. Pharm. Sci. 2004, 93,

2040-53.

Page 196: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

174

Chapter 6

Reduction of punch-sticking propensity of celecoxib by spherical

crystallization via polymer assisted quasi-emulsion solvent diffusion

Page 197: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

175

6.1 Synopsis

Punch-sticking during tablet compression is a common problem for many active

pharmaceutical ingredients (APIs), which renders tablet formulation development

challenging. Herein, we demonstrate that the punch-sticking propensity of a highly sticky

API, celecoxib (CEL), can be effectively reduced by spherical crystallization enabled by a

polymer-assisted quasi-emulsion solvent diffusion (QESD) process. Among three

commonly used pharmaceutical polymers, poly(vinylpyrrolidone) (PVP), hydroxypropyl

cellulose (HPC), and hydroxypropyl methylcellulose (HPMC), HPMC was the most

effective in stabilizing the transient emulsion during QESD and retarding the coalescence

of emulsion droplets and the initiation of CEL crystallization. These observations may arise

from stronger intermolecular interactions between HPMC and CEL, consistent with

solution 1H NMR analyses. SEM and X-ray photoelectron spectroscopy confirmed the

presence of a thin layer of HPMC on the surface of spherical particles. Thus, the sticking

propensity was significantly reduced because the HPMC coating prevented direct contact

between CEL and the punch tip during tablet compression.

6.2 Introduction

Several powder properties of active pharmaceutical ingredients (APIs), such as

tabletability, flowability, bulk density, and punch-sticking propensity are critical for

efficient and successful tablet formulation design, especially at high drug loadings (> 30%,

w/w).1-3 The punch-sticking propensity refers to the tendency of the API powder to adhere

to the punch tip during tablet compression. Relative to other powder properties, the punch-

sticking propensity of an API is rarely evaluated during early stage therapeutic

Page 198: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

176

development, even though it is a well-recognized common problem in tablet manufacturing

that adversely affects the aesthetic qualities of a finished tablet product.4-6 Punch-sticking

is influenced by process variables such as tableting speed, compaction pressure7, tooling

material8, and tooling design9, as well as API material properties, including mechanical

properties10, surface chemistry10, particle sizes7, and choice of excipients11. As compared

to modifying tablet compaction tools, elimination of punch-sticking through formulation

approaches is more desirable to ensure tablet quality because of its economy, effectiveness,

and broad applicability.

Among the formulation techniques that have been pursued to reduce the punch-

sticking propensity, crystal engineering presents an attractive approach as it addresses the

root cause, instead of the symptoms, of API deficiency by altering the API crystal structure

and its consequent material properties. For example, salt formation can reduce the punch-

sticking propensity by altering the mechanical properties and surface chemistries of

crystalline APIs.10, 12 When crystal structure engineering is not possible, another effective

approach to address API punch-sticking is particle engineering through processes such as

mixing with suitable excipients11 and dry granulation13. Quasi-emulsion solvent diffusion

(QESD) is a particle engineering process for preparing spherical agglomerates (SA)

consisting of fine API crystals.14 QESD has been employed to improve flowability, bulk

density, tabletability, and dissolution characteristics of APIs.15-19 However, it remains

unclear whether API engineering by QESD can be used to overcome the punch-sticking

problem.

Celecoxib (CEL) is a COX-2 selective nonsteroidal anti-inflammatory drug for

treating chronic pain and inflammation associated with osteoarthritis, rheumatoid arthritis,

Page 199: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

177

and other acute pain symptoms in adults.20 The commercial CEL crystal (Form III), which

exhibits high one dimensional elasticity,21 is a challenging API for tablet formulation

development because of its significant punch-sticking propensity.4 In this work, we

explored the specific use of QESD as a means to reduce the punch-sticking propensity of

CEL. Our studies reveal that QESD processing with (hydroxypropyl)methyl cellulose

(HPMC) can reduce CEL punch-sticking propensity while simultaneously improving its

tabletability and flowability, by forming spherical HPMC-coated CEL crystals.

6.3 Materials and methods

6.3.1 Materials

Celecoxib (CEL, Form III, Aarti Pvt Ltd., Mumbai, India), Microcrystalline

cellulose (Avicel PH102, FMC, Philadelphia, PA), HPMC (K15M, MW 575,000 g/mol,

Ashland Specialty Ingredients, Wilmington, DE), HPC (EF Pharm, MW 80,000 g/mol,

Ashland Specialty Ingredients, Wilmington, DE), PVP (90F, MW 1,000,000–1,500,000

g/mol, BASF, Ludwigshafen, Germany), magnesium stearate (MgSt, Mallinckrodt Inc., St.

Louis, MO), ethyl acetate (EA, Sigma-Aldrich, St. Louis, MO), ultra-pure deionized water

(0.066 μS/cm, Thermo Scientific, USA) were used as received. The chemical structures of

CEL, HPMC, HPC, and PVP are given in Figure 6.1. All crystallization experiments were

performed in glass beakers at the specified temperatures.

6.3.2 Polymer Screening for Quasi-Emulsion Solvent Diffusion (QESD)

Ethyl acetate (EA) and water were selected as good and poor solvent for the QESD

process based on solubility of CEL in the solvents and solvents miscibility. A solution

Page 200: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

178

containing 3 g CEL in 9 mL of EA at 70 °C was added dropwise to 70 mL of an aqueous

solution of 0.5% (w/w) polymer (HPMC, HPC, or PVP) at 22 °C to form spherical

agglomerates over 3 minutes. The mixture was agitated at 600 rpm using an overhead mixer.

Spherical agglomerates were captured on filter paper (Fisher Scientific, Pittsburgh, PA)

supported with a Brinell funnel after 15 min when the suspension no longer appeared to

change, and they were dried at 60 °C overnight in an oven. The morphology of produced

agglomerates was assessed to guide the selection of more appropriate polymer additives.

6.3.3 QESD Spherical Agglomerate Growth Kinetics

An 0.9 mL aliquot of a 0.33 mg/mL CEL solution in EA at 70 °C was added at once

to a 7 mL of a 0.5% (w/w) polymer (HPMC, HPC or PVP) aqueous solution under 600 rpm

agitation using a magnetic stir bar at room temperature. The time-dependence of the QESD

process was monitored in increments appropriate to the rates of drop solidification

associated with the different polymer additives. At each time point, a small volume of

suspension (less than 1 mL) was withdrawn using a disposable plastic pipette and

immediately transferred onto a glass slide for imaging under a polarized light microscope

(Eclipse E200; Nikon, Tokyo, Japan).

6.3.4 Particle Size Distribution

Particle size distribution measurement of CEL and CEL-QESD was carried out

using a particle size analyzer (Microtrac SIA, Montgomeryville, PA). Each powder (~50

mg) was suspended in Isopar G Fluid (∼10 mL), in which the solubility of CEL is

Page 201: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

179

negligible. The suspension was added dropwise into a sample delivery controller (SDC,

Montgomeryville, PA) until the particle density reached the acceptable range for each

measurement. A high-speed camera captured image of particles in focus, from which the

area equivalent diameter of each particle was determined using image analysis software

(Microtrac Flex) provided with the instrument. All experiments were performed in

triplicate.

6.3.5 Powder X-ray Diffractometry (PXRD)

Crystallographic properties of samples were characterized using a X’Pert PRO

powder X-ray diffractometer (PANalytical Inc., West Borough, MA) equipped with a

copper X-ray source (45 kV and 40 mA) to provide Kα radiation (1.5406 Å) over the

angular range 5° ≤ 2θ ≤ 35° using a 0.017° step size and a dwell time of 1.15 s. The

divergence and antiscattering slits on the incident beam path were 1/16 and 1/8 degree,

respectively, and the antiscattering slit on the diffracted beam path was 5.5 mm

(X’Celerator). The expected PXRD pattern of CEL Form III was calculated from its crystal

structure.21

6.3.6 Thermal Analyses

Powder samples (3−5 mg) were loaded into hermetically sealed, aluminum pans

and heated from 20 to 180 °C at a rate of 10 °C/min on a Q1000 differential scanning

calorimeter (DSC, TA Instruments, New Castle, DE) under a continuous nitrogen purge

at a flow rate of 50 mL/min. The DSC cell temperature was calibrated with both indium

Page 202: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

180

and cyclohexane, and the heat flow cell parameter was calibrated using the indium standard.

To measure the residual volatile content in the solids obtained by QESD,

thermogravimetric analyses were performed using a Q500 Thermogravimetric Analyzer

(TA Instruments, New Castle, DE). Solid samples (~3−10 mg) were placed in an open

aluminum pan and heated from 22 to 300 °C at a rate of 10 °C/min under 25 mL/min

nitrogen gas purge.

6.3.7 True density measurement

The true density of CEL was measured using a helium pycnometer (Quantachrome

Instruments, Ultrapycnometer 1000e, Byonton Beach, FL). Accurately weighed powders

(1−2 g) was filled into the sample cell. Sample volume was measured repeatedly until the

coefficient of variation of the last five consecutive measurements was below 0.005%. The

mean and standard deviation of the last five measurements were calculated and reported.

The true density of CEL-QESD was assumed to be the same as CEL since the amount of

HPMC in CEL-QESD was small and the density difference between HPMC and CEL is

also small.

6.3.8 Powder Flowability, Tabletability, and Punch Sticking Assessment

A ring shear cell tester (RST-XS; Dietmar Schulze, Wolfenbüttel, Germany) with

a 10 mL cell was used to assess powder flowability at a 1 kPa pre-shear normal stress. A

yield locus was obtained by conducting shear test under different normal stresses (0.25,

0.40, 0.55, 0.70, and 0.85 kPa), from which the unconfined yield strength (fc) and major

Page 203: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

181

principal stress (σn) were obtained by drawing Mohr’s circles. The flowability index (ffc)

was calculated using Eq. (1).

𝑓𝑓c =𝜎𝑛

𝑓c (1)

Tabletability was analyzed for both pure components and formulations consisting

of 79.5 wt% Avicel PH102, 20 wt% CEL powder, and 0.5 wt% MgSt. Tablets were

prepared by single-sided compression at compaction pressures ranging 50–300 MPa using

a Universal Material Testing Machine (Model 1485, Zwick/Roell, Germany) at a speed of

4 mm/min. A 5% (w/v) suspension of magnesium stearate in ethanol was used to coat the

punch tip and die wall, which was air dried with a fan for ~ 1 min to complete dryness.

Tablets were allowed to relax under ambient conditions for at least 24 h prior to being

broken diametrically using a Texture Analyzer (TA-XT2i, Texture Technologies

Corporation, Scarsdale, NY). Tablet flashing was carefully removed to improve the

accuracy of measured tablet thickness.22 Tablet tensile strength was calculated from the

breaking force (F), tablet thickness (h), and tablet diameter (d) according to Eq. (2):23

𝜎 =2𝐹

𝜋𝑑ℎ (2)

Punch sticking propensity was assessed for formulations consisting of 79.5 wt%

Avicel PH102, 20 wt% CEL powder, and 0.5 wt% MgSt using a Presster Compaction

Simulator (Measurement Control Corp., East Hanover, NJ) with a 12.7 mm diameter flat-

faced upper punch with a removable tip (tare weight 3 g). A total of 50 tablets were

compressed at 250 MPa with a dwell time of 25 ms, simulating a Korsch XL100 press (10

stations) operating at a production rate of 49,300 tablets/h. The punch tip was removed and

Page 204: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

182

weighed with precision of 0.01 mg after punching batches of 10 tablets to determine the

punch sticking kinetics, and the mean weight gain over three independent trials was plotted

against the number of compressions.

6.3.9 Compressibility and Compactibility Analyses

Compressibility, which is derived from a plot of tablet porosity as a function of

compaction pressure, indicates the extent of powder volume reduction under a given

compaction pressure. Porosity of the compacts (𝜀) was calculated from tablet density (𝜌𝑐)

and true density (𝜌𝑡 = 1.4833 ± 0.0004 g/cm3 for CEL) of the material using Eq. (3):

𝜀 = 1 − 𝜌𝑐/𝜌𝑡 (3)

Compactibility, which is derived from a plot of σ as a function of ε, was analyzed by

nonlinear regression according to Eq. (4):24

𝜎 = 𝜎0𝑒−𝑏·𝜀 (4)

where 𝜎0 is the tablet tensile strength at zero porosity and b is a constant, which represents

the sensitivity of σ to a change in 𝜀.

6.3.10 Scanning Electron Microscopy (SEM)

Samples were mounted onto carbon tapes and sputter-coated with a thin layer of

platinum (thickness ~ 75 Å) using an ion-beam sputter (IBS TM200S, VCR Group Inc.,

San Clemente, CA). Particle morphology and surface features were evaluated using a JEOL

Page 205: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

183

6500F scanning electron microscope (JEOL Ltd., Tokyo, Japan), operated at SEI mode

with an acceleration voltage of 5 kV under high vacuum (10-4–10-5 Pa) during imaging.

6.3.11 X-ray Photoelectron Spectroscopy (XPS)

The XPS measurements were performed using a PHI Versa Probe III XPS system

(ULVAC-PHI; Physical Electronics, Inc., Chanhassen, MN, USA) with monochromated

Al K X-ray source (1486.6 eV, power of 50 W under 15 kV) with a base pressure of 5.0

x 10-8 Pa and a 0.2 mm diameter X-ray spot size. During data collection, the pressure in

the sample chamber was maintained near 1.0 x 10-6 Pa. Samples were mounted on a

stainless steel holder using carbon tape. Tablet samples were mounted in such a way that

the surface of interest faced upward for elemental analysis. Charge neutralization was used

since the materials are electrically non-conductive. The incidence angle of X-ray was 90°

and take-off angle of electrons was 45°. The pass energy used was 280 eV with 1.0 eV/step.

Atomic percentages were calculated from the average survey spectrum (n = 3) using the

Multipak software provided with the XPS system.

6.3.12 HPMC Content Determination

HPMC content in CEL samples prepared by QESD was determined by size-

exclusion chromatography (SEC) using an Agilent 1260 liquid chromatography (LC)

operating with an aqueous mobile phase containing 0.1 M Na2SO4 and 1 wt% acetic acid

at a flow rate of 0.4 mL/min. The LC was equipped with one guard column and three

CATSEC separation columns with pore sizes of 1000, 300 and 100 Å, respectively.

Page 206: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

184

Analyte detection by an inline Wyatt Optilab T-rEX refractive-index detector (Wyatt

Technologies, Santa Barbara, CA) enabled HPMC concentration determination from the

peak area corresponding to polymer fraction in the ASTRA software. 25, 26 The refractive

index increment was dn/dc = 0.1165 mL/g for HPMC under the analysis conditions, which

was determined by SEC analyses of HPMC (K15M) solutions of known concentrations

assuming 100% mass recovery. For each measurement, CEL-QESD powder (0.65 g) was

dissolved in 5 mL HPLC grade tetrahydrofuran (THF) to separate the insoluble HPMC.

The insoluble HPMC polymer was then isolated by centrifugation force at 11,644 x g for

10 min and the CEL solution in THF was carefully decanted. In order to remove any

residual CEL, the HPMC polymer was re-dispersed in 5 mL THF by sonication for 10 min

and isolated again by centrifugation. Upon removal of residual THF under vacuum, the

isolated HPMC was dissolved in 5 mL of the aqueous SEC mobile phase, and the

concentration was determined by SEC. The reported HPMC content in each CEL-QESD

sample represents an average of three independent measurements.

6.3.13 1H NMR Spectroscopy

1D solution 1H NMR spectra in DMSO-d6 were acquired at 298 K on a Bruker AV-

400 NMR spectrometer (Bruker BioSpin GmbH, Rheinstetten, Germany) operating at a

proton resonance frequency of 400 MHz. Chemical shifts were referenced with respect to

the residual proton resonance of DMSO-d6 (2.50 ppm).

6.4 Results and discussion

Page 207: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

185

6.4.1 Polymer Screening and Granule Growth Process

During the QESD process, the polymer used in aqueous solution significantly

influenced the morphology of the resulting agglomerates. The sphericity of the CEL

agglomerates qualitatively decreased in the order of HPMC >> HPC > PVP ≈ no polymer

(Figure 6.2). The addition of HPMC notably led to spherical agglomerates with smooth

surfaces, while only irregular agglomerates with rough surfaces were produced in the

presence of HPC. The agglomerates produced from PVP solutions were large and even

more irregular than those from the HPC solution, in a manner similar to those obtained

from QESD without any polymer (Figure 6.2).

To better understand the origin of the distinct granule morphologies, the time-

dependent process of granule growth was monitored for each polymer additive (Figure 6.3).

In all aqueous polymer solutions examined here, emulsions were formed upon addition of

CEL solution. The emulsion formed by the HPMC solution was stable for 6 min, after

which the phase separated CEL droplets gradually solidified to form spherical

agglomerates (Figure 6.3a). In contrast, emulsions formed in the HPC and PVP solutions

were less stable, since crystallization of CEL droplets began after about 1 min and 30 sec,

respectively (Figures 6.3b & 6.3c). In the PVP solution, coalescence of emulsion droplets

also rapidly took place, which led to the formation of large granules (Figure 6.3c).

Thus, sufficient emulsion stability against coalescence is required for

crystallization of the droplets to occur at the surfaces and propagate into the droplet core.

This confined crystallization mode results in a spherical particle morphology that mimics

that of the parent emulsion droplets. Unstable emulsions, wherein coalescence of small

droplets of CEL takes place more quickly than crystallization, leads to the formation of

Page 208: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

186

large and irregular agglomerates. Therefore, the crystallization onset time represents a good

criterion for selecting appropriate polymers to enable the successful development of a

QESD process. This was achieved with 0.5% (w/w) HPMC, which was then used in

subsequent studies as the crystallization medium to produce high quality, spherical

agglomerates of CEL, designated as CEL-QESD.

6.4.2 Sticking propensity reduction by CEL-QESD

During tablet compression, a significant amount of the as-received CEL powder

visibly adhered to the punch surface after a single compression (Figure 6.4a), consistent

with the known severe punch-sticking propensity of CEL.11 In contrast, the CEL-QESD

powders exhibited significantly less punch-sticking propensity than as-received CEL.

Moreover, the degree of reduction was visually evident as the HPMC concentration in the

QESD process increased from 0.1 to 0.5 wt% (Figures 6.4b–6.4d). The CEL powder

prepared from the 0.5% HPMC aqueous solution exhibited a completely sticking free

performance upon compression (Figure 6.4d).

Quantitative evaluation of punch-sticking propensity was carried out following an

established method using formulated CEL,4 comprising of Avicel PH102 (79.5 wt%), CEL

(20 wt%), and MgSt (0.5 wt%). Results showed severe sticking propensity of the as-

received CEL formulation, where >1.80 mg weight gain was recorded after 50

compressions (Figure 6.4e). In comparison, the CEL-QESD formulation only had ~ 0.10

mg weight gain recorded after 50 compressions. Having established the successful

reduction of punch sticking by spherical crystallization of CEL (CEL-QESD), we further

examined the possible mechanisms in the following sections.

Page 209: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

187

6.4.3 Solid form and residual solvent

Both DSC (Figure 6.5a) and powder XRD (Figure 6.5b) analysis revealed that the

CEL-QESD, derived from the 0.5% HMPC emulsion, and the as-received CEL exhibited

the same thermal properties and crystal structure consistent with those of Form III.

Thermogravimetric analyses further demonstrated the absence of detectable residual

solvent in the CEL-QESD, since no sample weight loss was observed with increasing the

temperature to 180 °C (Figure 6.5c). Therefore, the drastically reduced punch sticking

propensity cannot be ascribed to the presence of residual solvent nor change in CEL crystal

form.

6.4.4 Manufacturability of CEL Powders

To evaluate the manufacturability of CEL-QESD, we characterized the flowability

and tabletability. The flowability index ffc of CEL-QESD was much higher than that of the

as-received CEL (Figure 6.6a). By the measure of ffc, CEL-QESD exhibited flowability

suitable for high-speed tableting given that it flows better than Avicel PH102, but the as-

received CEL did not (Figure 6.6a).27 The improvement in powder flowability is an

expected outcome for SA powders, due to the combination of the spherical morphology,

smooth surface, and larger particle size of the CEL-QESD. Compared to the elongated fine

crystals of the as-received CEL, CEL-QESD agglomerates were larger and more spherical.

The spherical shape of the fine CEL-QESD particles in the range of 10–40 µm (Figures 6.2

and 6.7a) suggests that they were likely resulted from the solidification of small droplets

during the QESD process.

Page 210: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

188

CEL-QESD also exhibited profoundly improved tabletability, with tablet tensile

strength more than two-fold higher than that of as-received CEL (Figure 6.6b). Analysis

of the results shows that the concentration of HMPC content in the prepared CEL-QESD

samples positively correlates with the HPMC concentration in the aqueous medium for

QESD (Figure 6.7b). From the 0.5% HPMC solution, the CEL-QESD sample contained

1.04 ± 0.07% (w/w) HPMC (Figure 6.7b). Thus, the much improved tabletability may arise

from a HMPC surface coating layer on the spherical agglomerate that facilitates the

formation of a 3D bonding network upon powder compression, which was shown to

extremely effective in improving powder tabletability.28

Consistent with the higher tabletability of the pure CEL-QESD powders (Figure

6.6b), CEL-QESD also exhibited much better tabletability than the CEL formulation

(Figure 6.8a). According to the bonding area (BA) – bonding strength (BS) interplay

model,29 a higher tabletability is a consequence of large BA, higher BS, or both among the

adjacent particles. BA and BS may be assessed using compressibility and compactibility

plots, respectively. The nearly identical compressibility profiles of the two formulations

(Figure 6.9) imply similar BA in tablets formed under the same compaction pressures. On

the other hand, the compactibility profile of the CEL-QESD formulation was much higher

than that of as-received CEL formulation, as shown by the higher σ at the same tablet

porosity (Figure 6.8b), indicative of its higher apparent BS. Therefore, the higher

tabletability of CEL-QESD formulation was driven by the higher BS. A higher BS among

particles tends to lower sticking propensity, because the inter-particle adhesion is preferred

to punch-particle adhesion, when the tablet is separated from the punch during tablet

manufacturing.

Page 211: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

189

6.4.5 Polymer Coating on CEL-QESD

To confirm the suspected surface HPMC coating, based on the improved

tabletability, we first compared the surface properties of the different CEL and CEL-QESD

particles. When viewed at a high magnification, the irregular and rough surfaces of CEL-

QESD sharply contrasted the smooth surfaces of as-received CEL crystals (Figure 6.10).

The presence of irregular surfaces can suggest the presence of non-crystalline (amorphous)

material. However, given the high degree of CEL crystallinity observed by DSC and PXRD

(Figures 6.5 and 6.9), this possibility could be eliminated. Another possible origin of the

irregular surface features of the CEL-QESD is the presence of a thin HPMC coating. Such

HPMC coating would be consistent with the reduced punch-sticking propensity (Figure

6.4e), because it prevents the direct contact between CEL and punch.

The presence of a surface HPMC layer is further supported by the XPS data, which

provides the surface elemental composition. The oxygen-rich nature of the cellulosic

backbone of HMPC coupled with the unique occurrence of F and N in CEL (Figure 6.1)

makes it possible to probe particle surface compositions by assessing the relative

abundances of O, F, and N. Compared to the neat CEL powder, the CEL-QESD powder

exhibited a significantly higher O 1s signal with concomitant suppression of the N 1s and

F 1s signals (Figure 6.11a, Table 6.1). This result is consistent with the aforementioned

notion that HMPC decorated the outer surface of the CEL-QESD. Upon compression into

tablets, the CEL-QESD tablet also exhibited significantly lower F 1s and N 1s signals as

compared to the tablets of the as-received CEL (Figure 6.11b, Table 6.1). This confirms

the presence of a surface HPMC layer on the CEL-QESD particles and at the surface of

Page 212: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

190

tablet. Given the shallow take-off angle (45 o) employed in the XPS measurements, the

observation of a F 1s signal in the CEL-QESD samples suggests that the minimum

thickness of the HPMC layer coating the particle surfaces was thinner than the typical

sampling depth of organic compounds by XPS, ~5 nm.10, 30 Otherwise, N and F in CEL

would have not been detected if the HPMC layer was always thicker than the maximum

sampling thickness by XPS.

6.4.6 Interactions between CEL and polymers

To gain molecular-level insight into the very different performances of PVP, HPC,

and HPMC in facilitating the QESD process of CEL (Figures 6.2 and 6.3), we sought to

identify preferential intermolecular interactions between CEL and the various polymers

using solution 1H NMR (Figure 6.12). We directed our attention to changes in the

appearance of the peaks associated with protons ortho to the aryl sulfonamide (H15 and

17), meta to the arylsulfonamide (H14 and H18), the sulfonamide –NH2 (H23), tolyl

protons (H7, H8, H10, H11, H12), and the pyrazole proton (H4) (see the numbering scheme

in Figure 6.1). Although deuterium oxide is the most relevant solvent to investigate the

drug-polymer interactions in the QESD process carried out in water, deuterium oxide could

not be used because CEL is insoluble in it. Therefore, we employed DMSO-d6 as the

alternative solvent because of its high polarity and ability to dissolve both the polymers

and CEL. Addition of HPMC (3 mg/mL) into a DMSO-d6 solution of CEL (3 mg/mL) led

to clear upfield shifts in the 1H NMR resonances associated with H14 and H18 ( 7.84–

7.90 ppm), H15 and H17 ( 7.52–7.60 ppm), H4, H7, H8, and H10 ( 7.16–7.24 ppm), and

H12 ( 2.28–2.36 ppm). The observed upfield shifts of the 1H NMR resonances suggest

Page 213: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

191

intermolecular hydrophobic interactions between CEL molecules and/or CEL and HMPC

that resulted in increased electron densities at these positions.31, 32 We note that upfield

shifts of the 1H NMR resonances of aromatic compounds are often observed in their guest-

host inclusion complexes with cyclodextrins,33 the latter of which are cyclic saccharides

that are structurally related to HPMC. The more hydrophilic variant of HPC, apparently

resulted in much smaller shifts in the 1H NMR resonances of CEL, indicating weaker

intermolecular polymer/CEL interactions. Finally, very minimal peak shifts were observed

with the most hydrophilic polymer PVP. These 1H NMR observations lead us to conclude

that hydrophobic interactions between the aryl and aliphatic moieties of the CEL and the

permethylated HMPC additive drive their association in DMSO-d6 solutions. We note that

HMPC addition drove a downfield shift of the peak associated with H23 ( 7.50–7.52 ppm).

This last downfield shift of peaks indicated a decrease in electron density at these protons,

suggestive of sulfonamide hydrogen bond donation to either the HPMC or to other CEL

molecules that are hydrophobically associated with the polymer. No such peak shifts were

observed with the more hydrophilic HPC and PVP additives, even though PVP is solely a

hydrogen bond acceptor that could form complementary interactions with free CEL. This

last observation suggests that intermolecular interactions between hydrophobically

associated CEL molecules along a HPMC chain drive substantial changed in the chemical

environment around the amino group that lead to the observed chemical shift.

The NMR data shown in Figure 6.12 suggests that CEL-polymer association

strength follows the descending order of HPMC >> HPC > PVP. This order is consistent

with their CEL emulsion stabilization abilities discerned from the time-dependent optical

micrographs (Figure 6.3). Thus, hydrophobic association between the CEL and HPMC

Page 214: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

192

appears to play a critical role in stabilizing the quasi-emulsion and enabling the formation

of spherical agglomerates. Compared to that observed by 1H NMR in DMSO-d6, the

hydrophobic association is likely stronger in the higher polarity H2O during the QESD

process. HPMC chains may thus be driven from the aqueous medium to adsorb onto the

emulsion droplet interfaces, leading to their temporal stabilization that allows for spherical

agglomerate formation with a thin HPMC coating. These findings concur with previous

studies that demonstrated the ability of HPMC to form specific intermolecular interactions

with CEL to significantly reduce the crystal growth in a supersaturated solution,34 whereas

the more hydrophilic PVP did not.35

6.5 Conclusions

We have shown that the strength of the intermolecular interactions between CEL

and HPMC, HPC, and PVP is positively correlated with their ability in facilitating spherical

agglomeration of CEL by the QESD process. CEL-QESD agglomerates exhibit

significantly reduced the punch sticking propensity of CEL with substantially improved

tabletability and flowability. These significant enhancements in CEL powder properties are

attributed to QESD emulsion droplet stabilization by HPMC adsorption to the droplet

surfaces. This results in a thin HPMC coating on the surfaces of the resulting CEL particles,

which reduces punch sticking propensity by forming a physical barrier to separate CEL

from punch surface during compression. This first successful example of reducing punch-

sticking propensity of a highly sticky API by QESD, coupled with molecular-level insight

into the underlying mechanism, adds another powerful particle engineering strategy to the

arsenal of tools that can be used to solve common tablet manufacturing problems.

Page 215: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

193

Figure 6.1. Chemical structures of (a) the model API CEL, and commonly used polymeric

stabilizers (b) HPMC, (c) HPC, and (d) PVP.

(a) (b) (c) (d)

Page 216: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

194

Figure 6.2. Micrographs of agglomerates prepared by QESD method in the absence and

presence of various polymers.

Without polymer 0.5% HPMC

0.5% HPC 0.5% PVP

100μm

m

500μm

m 500μm

m

500μm

m

Page 217: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

195

Figure 6.3. Time-dependent growth of CEL spherical agglomerates in various aqueous

polymer solutions assessed by a polarized light microscopy: a) HPMC, b) HPC, and c)

PVP. All scale bars are 500 µm.

30 sec 3 min 6 min 15 min

a)

30 sec 1 min 1.5 min 2 min

b)

5 sec 15 sec 30 sec 2 min

c)

Page 218: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

196

Figure 6.4. Punch sticking propensity of CEL after compaction at 50 MPa: a) as-received

CEL, and CEL-QESD prepared from different concentrations of HPMC solutions b) 0.1%,

c) 0.3%, d) 0.5%. (e) Quantitative sticking propensity assessment using formulated CEL

powders (79.5% (w/w) Avicel PH102, 20% CEL, and 0.5% MgSt).

a) b)

c) d)

e)

Page 219: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

197

Figure 6.5. Thermal behavior of CEL as-received and CEL QESD (from a 0.5% HPMC

solution) a) differential scanning calorimetry, b) thermogravimetric analysis of as-received

and CEL-QESD samples, and c) PXRD patterns for calculated, as-received, and CEL-

QESD samples.

a) b)

c)

Page 220: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

198

Figure 6.6. a) Flowability (at 1 kPa pre-shear normal stress) and b) Tabletability of CEL-

QESD as compared to as-received CEL.

a) b)

Page 221: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

199

Figure 6.7. a) particle size distributions of as-received CEL and CEL-QESD powders, b)

HPMC content in solid CEL-QESD plotted as a function of the HPMC solution

concentration (n = 3).

a) b)

Page 222: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

200

Figure 6.8. Compaction properties of as-received CEL and CEL-QESD formulations

consisting of 79.5 wt% Avicel PH102, 20 wt% CEL powder, and 0.5 wt% MgSt: a)

tabletability and b) compactibility.

a) b)

Page 223: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

201

Figure 6.9. Compressibility plot of 20% (w/w) as-received CEL and CEL-QESD

formulations in 79.5% Avicel PH102 and 0.5% Magnesium stearate.

Page 224: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

202

Figure 6.10. SEM images of CEL-QESD (top row) and as-received CEL (bottom row) at

low (left column) and high (right column) magnifications.

100 µm

100 µm 10 µm

10 µm

Page 225: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

203

Figure 6.11. X-ray photoelectron spectral analysis of the particle surface elemental

compositions of as-received CEL and CEL-QESD, (a) powder, and (b) tablets.

a) b)

Page 226: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

204

Table 6.1. Elemental composition of sample surfaces

CEL (%) CEL-QESD (%) CEL/Tablet (%) CEL-QESD/Tablet (%)

C1s 69.13 (±1.5) 60.27 (±1.55) 69.1 (±1.5) 86.30 (±0.30)

N1s 8.93 (±0.32) 3.37 (±0.42) 3.4 (±0.4) 0

O1s 9.83 (±0.29) 32.00 (±2.65) 9.8 (±0.3) 11.20 (±1.51)

F1s 9.57 (±0.95) 3.17 (±0.70) 9.1 (±1.3) 2.07 (±1.44)

Page 227: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

205

Figure 6.12. 1D 1H NMR spectra of CEL (3 mg/mL) with different polymer additives (3

mg/mL) in DMSO-d6 over the chemical shift ranges a) 7.80-7.90 ppm, b) 7.40-7.60

ppm, c) 7.12-7.28 ppm, and d) 2.20-2.40 ppm. The numbering of the resonances

corresponds to that given in Figure 6.1.

b) a)

c) d)

H14, 18 H15, 17 H23

H7, 8, 10, and 11 H4 H12

Page 228: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

206

6.6 References

1. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

2. Lakshman, J. P.; Kowalski, J.; Vasanthavada, M.; Tong, W. Q.; Joshi, Y. M.;

Serajuddin, A. T. Application of melt granulation technology to enhance tabletting

properties of poorly compactible high-dose drugs. J. Pharm. Sci. 2011, 100, 1553-1565.

3. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high

drug loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

4. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.;

Meenan, P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J.

Pharm. Sci. 2017, 106, 151-158.

5. Wang, J. J.; Guillot, M. A.; Bateman, S. D.; Morris, K. R. Modeling of adhesion

in tablet compression. Ii. Compaction studies using a compaction simulator and an

instrumented tablet press. J. Pharm. Sci. 2004, 93, 407-417.

6. Chattoraj, S.; Daugherity, P.; McDermott, T.; Olsofsky, A.; Roth, W. J.; Tobyn, M.

Sticking and picking in pharmaceutical tablet compression: An iq consortium review. J.

Pharm. Sci. 2018, 107, 2267-2282.

7. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J. F.; Dawson, N.; Mullarney, M.

P.; Meenan, P.; Sun, C. C. Powder properties and compaction parameters that influence

punch sticking propensity of pharmaceuticals. Int. J. Pharm. 2017, 521, 374-383.

8. Abdel-Hamid, S.; Betz, G. A novel tool for the prediction of tablet sticking during

high speed compaction. Pharm. Dev. Technol. 2012, 17, 747-754.

9. Roberts, M.; Ford, J. L.; MacLeod, G. S.; Fell, J. T.; Smith, G. W.; Rowe, P. H.

Effects of surface roughness and chrome plating of punch tips on the sticking tendencies

of model ibuprofen formulations. J. Pharm. Pharmacol. 2003, 55, 1223-1228.

10. Paul, S.; Wang, C.; Wang, K.; Sun, C. C. Reduced punch sticking propensity of

acesulfame by salt formation: Role of crystal mechanical property and surface chemistry.

Mol. Pharm. 2019, 16, 2700-2707.

11. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through

excipient selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

12. Al-Karawi, C.; Leopold, C. S. A comparative study on the sticking tendency of

ibuprofen and ibuprofen sodium dihydrate to differently coated tablet punches. Eur. J.

Pharm. Biopharm. 2018, 128, 107-118.

13. Simmons, D. M.; Gierer, D. S. A material sparing test to predict punch sticking

during formulation development. Drug Dev. Ind. Pharm. 2012, 38, 1054-1060.

14. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

15. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

Page 229: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

207

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion method. Chem. Pharm. Bull. 1990, 38, 733-739.

16. Kawashima, Y.; Niwa, T.; Takeuchi, H.; Hino, T.; Itoh, Y.; Furuyama, S.

Characterization of polymorphs of tranilast anhydrate and tranilast monohydrate when

crystallized by two solvent change spherical crystallization techniques. J. Pharm. Sci. 1991,

80, 472-478.

17. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

18. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

19. Tahara, K.; Kono, Y.; Myerson, A. S.; Takeuchi, H. Development of continuous

spherical crystallization to prepare fenofibrate agglomerates with impurity complexation

using mixed-suspension, mixed-product removal crystallizer. Cryst. Growth Des. 2018, 18,

6448-6454.

20. Ray, W. A.; Stein, C. M.; Daugherty, J. R.; Hall, K.; Arbogast, P. G.; Griffin, M.

R. Cox-2 selective non-steroidal anti-inflammatory drugs and risk of serious coronary

heart disease. The Lancet 2002, 360, 1071-1073.

21. Wang, K.; Mishra, M. K.; Sun, C. C. Exceptionally elastic single-component

pharmaceutical crystals. Chem. Mater. 2019, 31, 1794-1799.

22. Paul, S.; Chang, S.-Y.; Sun, C. C. The phenomenon of tablet flashing — its impact

on tableting data analysis and a method to eliminate it. Powder Technol. 2017, 305, 117-

124.

23. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

24. Ryshkewitch, E. Compression strength of porous sintered alumina and zirconia. J.

Am. Ceram. Soc. 1953, 36, 65-68.

25. Chen, M.-H.; Bergman, C. J. Method for determining the amylose content,

molecular weights, and weight- and molar-based distributions of degree of polymerization

of amylose and fine-structure of amylopectin. Carbohydr. Polym. 2007, 69, 562-578.

26. Gapper, L. W.; Copestake, D. E. J.; Otter, D. E.; Indyk, H. E. J. A.; Chemistry, B.

Analysis of bovine immunoglobulin g in milk, colostrum and dietary supplements: A

review. Anal. Bioanal. Chem. 2007, 389, 93-109.

27. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

28. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

Page 230: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

208

29. Osei-Yeboah, F.; Chang, S.-Y.; Sun, C. C. J. P. R. A critical examination of the

phenomenon of bonding area - bonding strength interplay in powder tableting. Pharm. Res.

2016, 33, 1126-1132.

30. Tanuma, S.; Powell, C. J.; Penn, D. R. Calculations of electorn inelastic mean free

paths. Ii. Data for 27 elements over the 50–2000 ev range. Surf. Interface Anal. 1991, 17,

911-926.

31. Chen, Y.; Pui, Y.; Chen, H.; Wang, S.; Serno, P.; Tonnis, W.; Chen, L.; Qian, F.

Polymer-mediated drug supersaturation controlled by drug–polymer interactions persisting

in an aqueous environment. Mol. Pharm. 2019, 16, 205-213.

32. Jasani, M. S.; Kale, D. P.; Singh, I. P.; Bansal, A. K. Influence of drug–polymer

interactions on dissolution of thermodynamically highly unstable cocrystal. Mol. Pharm.

2019, 16, 151-164.

33. Schneider, H.-J.; Hacket, F.; Rüdiger, V.; Ikeda, H. Nmr studies of cyclodextrins

and cyclodextrin complexes. Chemical Reviews 1998, 98, 1755-1786.

34. Ilevbare, G. A.; Liu, H.; Edgar, K. J.; Taylor, L. S. Impact of polymers on crystal

growth rate of structurally diverse compounds from aqueous solution. Mol. Pharm. 2013,

10, 2381-2393.

35. Xie, T.; Taylor, L. S. Dissolution performance of high drug loading celecoxib

amorphous solid dispersions formulated with polymer combinations. Pharm. res. 2016, 33,

739-750.

Page 231: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

209

Chaper 7

Research summary and future work

Page 232: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

210

Research summary

The thesis work aims to develop robust high drug loading direct compression tablet

formulations using the spherical crystallization technique. The key points of each chapter

are summarized.

In Chapter 2, an extremely high drug loading (99%) direct compression tablet

formulation was successfully developed via the spherical crystallization on the model API

ferulic acid. A tablet formulation requires to meet different crucial properties such as

adequate tablet tensile strength, powder free-flowing, fast dissolution, short disintegration

time, small tablet ejection force, low tablet friability, etc. Among these essential

requirements, tabletability, flowability and dissolution are the most challenging criteria to

deliver in a tablet formulation. Spherical crystallization, a technique that can

simultaneously improve the tabletability and flowability of API, provides a unique

advantage to enable the formulation development. Herein, spherical agglomerates of

ferulic acid was successfully prepared using the QESD method, generating QESD powder

with excellent flow and tableting performance. Owing to these exceptional properties, a

99% API loading direct compression tablet formulation was successfully developed.

In Chapter 3, an in-situ solvation and desolvation process during spherical

crystallization was observed, which generated spherical agglomerates with excellent

plasticity and consequently strong tablet. The solvent DCM not only functioned as the

bridging liquid but also reacted with the API GSF to form a solvate GSF-DCM during the

SA process. Since GSF-DCM is not stable even at room temperature, the spherical

agglomerates of solvate desolvated into the parent drug GSF completely during the drying

Page 233: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

211

process. This solvation-desolvation process led to the porous structure of the primary

crystals owing to the escaping of the solvent DCM. During tablet compaction, the

nanopores collapse and the particles are more deformable, which explains the origin of

superior plasticity of spherical agglomerates.

In Chapter 4 and 5, spherical cocrystallization techniques were developed to

simultaneously improve the manufacturability and dissolution performance of BCS Class

II drugs and thus enable high drug loading tablet formulation development. In Chapter 4,

the spherical cocrystallization, based on the combination of SA and cocrystallization, was

applied on the model drug GSF with coformer Acs. In addition, a highly water-soluble

polymer HPC was incorporated to enhance tabletability as well as the wettability of

spherical cocrystal agglomerates. The excellent properties of spherical cocrystal

agglomerates enabled the successful high drug loading (55.7%) tablet formulation

development. In Chapter 5, the spherical cocrystallization technique, a combination of

QESD method and cocrystallization, was applied on the model drug IMC and coformer

SAC, and the polymer HPMC in the water phase functioned as a stabilizer of the emulsions.

1H NMR and DFT computational simulation studies suggest that HPMC interacted with

the SAC molecule through the hydrogen bonds, through which the emulsions were

stabilized. The resulting agglomerates presented spherical shape and large size, leading to

excellent powder flowability. Polymer coating on the particle surface as well as much

smaller primary crystals consisted of agglomerates resulted in excellent tabletability.

Additionally, the much higher solubility of cocrystal over the parent drug IMC implies the

better dissolution. Owing to the exceptional flowability, tabletability and enhanced

Page 234: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

212

solubility, a high IMC loading (46.3%) tablet formulation using the direct compression

process was developed.

Based on the mechanism of QESD process, in Chapter 6, the QESD method on the

model drug CEL was applied to reduce the punch sticking propensity. The polymer coating

on the particle surface confirmed by the SEM and XPS analysis, form a physical barrier

between the CEL and punch, which is the key for the reduced the punch sticking propensity.

In addition, the polymer coating substantially improved tabletability of CEL owing to the

formation of a strong 3D bonding structure during tablet compaction. To understand the

origin of the polymer coating, 1H NMR study was carried out to investigate the molecular

interactions between CEL and different polymers. The result suggests that the

intermolecular interactions between CEL and HPMC is the strongest, which is essential to

stabilize the emulsions and form the spherical agglomerates.

Page 235: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

213

Future work

Spherical cocrystallization

In Chapter 4 and 5, we successfully developed spherical cocrystallization

techniques based on the SA and QESD methods, and the generated spherical cocrystal

agglomerates presented promising properties (e.g., manufacturability, solubility) that are

attractive in the pharmaceutical product development. However, these properties

improvement are API dependent, and therefore intensive preliminary investigations are

required to ensure the spherical cocrystallization process development.

Selection of a suitable cocrystal with favored mechanical property and solubility is

the first step but crucial to the following development. To improve the solubility of the

drug, a highly water-soluble coformer is preferred since the cocrystal solubility may be

directly proportional to the solubility of constituent reactants.1 Caution should be taken

here since a highly soluble cocrystal has a higher free energy than the parent drug. The

high free energy difference is prone to lead to the phase transformation from cocrystal form

to parent drug during the dissolution, which can be detrimental to the dissolution

performance.

Secondly, screening of a bridging liquid that can wet the cocrystal well to generate

spherical cocrystal agglomerates. Solubility measurement in various solvents, a routine

way to screen the bridging liquid in the spherical crystallization technique, may not be

easily applicable here since the solubility measurement of cocrystal is more tedious. From

parent drug to cocrystal, transferring from one component to multiple components,

selection of bridging liquids using calculation tools may require more calculation resources

Page 236: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

214

and longer time to finish. Therefore, the development of a quicker and easier screening

method is necessary.

Thirdly, although the cocrystal has higher solubility than the parent drug, the

dissolution of spherical cocrystal agglomerates may be even worse due to the reduced

surface area exposed to the medium2. In this scenario, incorporation of superdisintegrant

or highly water-soluble polymer in the spherical cocrystal agglomerates will be helpful,

which can break down the large agglomerates into small primary crystals or wet the

agglomerates to increase the exposure of cocrystal to the medium.

Spherical amorphization

Amorphous drugs are becoming more and more attractive owing to its solubility

advantage over its crystalline form. Spherical amorphization, a technique that combines

the spherical agglomeration and amorphization, may be developed to simultaneously

improve the manufacturability and dissolution. Spherical amorphization may have

advantages comparing to the commonly used amorphization techniques such as spray

drying and hot melt extrusion. The particles produced by the spray drying method are

relatively small and require further unit operations such as dry granulation to improve the

flow property. The hot melt extrusion has high energy consumption and requires the drug

to be stable at high temperature.

Continuous manufacturing of spherical crystallization

As discussed in the Chapter 1, many parameters are involved in the spherical

crystallization process and have the potential to influence the properties of spherical

agglomerates. Batch manufacturing process in large scale can cause noticeable batch to

Page 237: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

215

batch variation and therefore is not preferred. Instead, a continuous manufacturing process

of spherical crystallization is recommended where the spherical agglomerate is generated

continuously, and the properties can be monitored on-time. Recent advances on the

continuous spherical crystallization process development imply the potential to

manufacture the spherical agglomerates continuously. Still, more investigations are needed

to explore the influence of processing parameters during the continuous manufacturing on

the generated spherical agglomerates comparing to batch manufacturing in small scale. In

addition, no work has been reported on the development of continuous spherical

cocrystallization, which is preferred especially for the water insoluble BCS Class II and IV

drugs.

7.1 Reference

1. Good, D. J.; Rodríguez-Hornedo, N. Solubility advantage of pharmaceutical

cocrystals. Crys. Growth Des. 2009, 9, 2252-2264.

2. Chen, H.; Wang, C.; Kang, H.; Zhi, B.; Haynes, C. L.; Aburub, A.; Sun, C. C.

Microstructures and pharmaceutical properties of ferulic acid agglomerates prepared by

different spherical crystallization methods. Int. J. Pharm. 2020, 574, 118914.

Page 238: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

216

Bibliography

Chapter 1.

1. Gupta, H.; Bhandari, D.; Sharma, A. Recent trends in oral drug delivery: A review.

Recent Pat. Drug Deliv. Formul. 2009, 3, 162-73.

2. Roopwani, R.; Buckner, I. S. Co-processed particles: An approach to transform

poor tableting properties. J. Pharm. Sci. 2019, 108, 3209-3217.

3. Taylor, M. K.; Ginsburg, J.; Hickey, A. J.; Gheyas, F. Composite method to

quantify powder flow as a screening method in early tablet or capsule formulation

development. AAPS PharmSciTech 2000, 1, 20-30.

4. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

5. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine

active pharmaceutical ingredient powder blends and tablets at high drug loading via dry

particle coating. Int. J. Pharm. 2018, 543, 288-299.

6. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

7. Braig, V.; Konnerth, C.; Peukert, W.; Lee, G. Enhanced dissolution of naproxen

from pure-drug, crystalline nanoparticles: A case study formulated into spray-dried

granules and compressed tablets. Int. J. Pharm. 2019, 554, 54-60.

8. Ullah, M.; Hussain, I.; Sun, C. C. The development of carbamazepine-succinic

acid cocrystal tablet formulations with improved in vitro and in vivo performance. Drug

Dev. Ind. Pharm. 2016, 42, 969-976.

9. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

10. Chen, H.; Wang, C.; Sun, C. C. Profoundly improved plasticity and tabletability

of griseofulvin by in situ solvation and desolvation during spherical crystallization. Cryst.

Growth Des. 2019, 19, 2350-2357.

11. Chen, L.; Ding, X.; He, Z.; Huang, Z.; Kunnath, K. T.; Zheng, K.; Davé, R. N.

Surface engineered excipients: I. Improved functional properties of fine grade

microcrystalline cellulose. Int. J. Pharm. 2018, 536, 127-137.

12. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

13. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

14. Garala, K. C.; Patel, J. M.; Dhingani, A. P.; Dharamsi, A. T. Preparation and

evaluation of agglomerated crystals by crystallo-co-agglomeration: An integrated approach

Page 239: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

217

of principal component analysis and box-behnken experimental design. Int. J. Pharm. 2013,

452, 135-56.

15. Garala, K. C.; Patel, J. M.; Dhingani, A. P.; Dharamsi, A. T. Quality by design

(qbd) approach for developing agglomerates containing racecadotril and loperamide

hydrochloride by crystallo-co-agglomeration. Powder Technol. 2013, 247, 128-146.

16. Ribardière, A.; Tchoreloff, P.; Couarraze, G.; Puisieux, F. Modification of

ketoprofen bead structure produced by the spherical crystallization technique with a two-

solvent system. Int. J. Pharm. 1996, 144, 195-207.

17. Subero-Couroyer, C.; Mangin, D.; Rivoire, A.; Blandin, A. F.; Klein, J. P.

Agglomeration in suspension of salicylic acid fine particles: Analysis of the wetting period

and effect of the binder injection mode on the final agglomerate size. Powder Technol.

2006, 161, 98-109.

18. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

19. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

20. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

21. Maghsoodi, M.; Taghizadeh, O.; Martin, G. P.; Nokhodchi, A. Particle design of

naproxen-disintegrant agglomerates for direct compression by a crystallo-co-

agglomeration technique. Int. J. Pharm. 2008, 351, 45-54.

22. Sano, A.; Kuriki, T.; Handa, T.; Takeuchi, H.; Kawashima, Y. Particle design of

tolbutamide in the presence of soluble polymer or surfactant by the spherical crystallization

technique: Improvement of dissolution rate. J. Pharm. Sci. 1987, 76, 471-474.

23. Masumi Ueda, Y. N. a. Y. K. Particle design of enoxacin by spherical

crystallization technique. I principle of ammonia diffusion system (ads). Chem. Pharm.

Bull. 1990, 38, 2537-2541.

24. Chow, A. H. L.; Leung, M. W. M. A study of the mechanisms of wet spherical

agglomeration of pharmaceutical powders. Drug Dev. Ind. Pharm. 1996, 22, 357-371.

25. Peña, R.; Jarmer, D. J.; Burcham, C. L.; Nagy, Z. K. Further understanding of

agglomeration mechanisms in spherical crystallization systems: Benzoic acid case study.

Crys. Growth Des. 2019, 19, 1668-1679.

26. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

27. Butt, H.-J.; Kappl, M. Normal capillary forces. Adv. Colloid Interface Sci. 2009,

146, 48-60.

28. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

Page 240: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

218

29. Amaro-Gonza, D.; Biscans, B. a. Spherical agglomeration during crystallization of

an active pharmaceutical ingredient. Powder Technol. 2002, 128, 188-194.

30. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

31. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Cryst. Growth Des. 2013, 13, 4162-4167.

32. Chen, M.; Liu, X.; Yu, C.; Yao, M.; Xu, S.; Tang, W.; Song, X.; Dong, W.; Wang,

G.; Gong, J. Strategy of selecting solvent systems for spherical agglomeration by the

lifshitz-van der waals acid-base approach. Chem. Eng. Sci. 2020, 220, 115613.

33. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

34. Yoshiaki Kawashima, F. C. Parameters determining the agglomeration behaviour

and the micromeritic properties of spherically agglomerated crystals prepared by the

spherical crystallization. Int. J. Pharm. 1995, 119, 139-147.

35. Blandin, A. F.; Mangin, D.; Rivoire, A.; Klein, J. P.; Bossoutrot, J. M.

Agglomeration in suspension of salicylic acid fine particles: Influence of some process

parameters on kinetics and agglomerate final size. Powder Technol. 2003, 130, 316-323.

36. Thati, J.; Rasmuson, A. C. On the mechanisms of formation of spherical

agglomerates. Eur. J. Pharm. Sci. 2011, 42, 365-79.

37. Kawashima, Y.; Okumura, M.; Takenaka, H. The effects of temperature on the

spherical crystallization of salicylic acid. Powder Technol. 1984, 39, 41-47.

38. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

39. Nocent, M.; Bertocchi, L.; Espitalier, F.; Baron, M.; Couarraze, G. Definition of a

solvent system for spherical crystallization of salbutamol sulfate by quasi‐ emulsion

solvent diffusion (qesd) method. J. Pharm. Sci. 2001, 90, 1620-1627.

40. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion method. Chem. Pharm. Bull. 1990, 38, 733-739.

41. Zhou, X.; Zhang, Q.; Xu, R.; Chen, D.; Hao, S.; Nie, F.; Li, H. A novel spherulitic

self-assembly strategy for organic explosives: Modifying the hydrogen bonds by polymeric

additives in emulsion crystallization. Cryst. Growth Des. 2018, 18, 2417–2423.

42. Kawashima, Y.; Cui, F.; Takeuchi, H.; Niwa, T.; Hino, T.; Kiuchi, K.

Improvements in flowability and compressibility of pharmaceutical crystals for direct

tabletting by spherical crystallization with a two-solvent system. Powder Technol. 1994,

78, 151-157.

Page 241: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

219

43. Kedia, K.; Wairkar, S. Improved micromeritics, packing properties and

compressibility of high dose drug, cycloserine, by spherical crystallization. Powder

Technol. 2019, 344, 665-672.

44. Orlewski, P. M.; Ahn, B.; Mazzotti, M. Tuning the particle sizes in spherical

agglomeration. Cryst. Growth Des. 2018, 10, 6257-6265.

45. Sun, W.; Chen, H.; Aburub, A.; Sun, C. C. A platformdirect compression

formulation for low dose sustained-release tablets enabled by a dual particle engineering

approach. Powder Technol. 2019, 342, 856-863.

46. Teunou, E.; Fitzpatrick, J. J. Effect of relative humidity and temperature on food

powder flowability. J. Food Eng. 1999, 42, 109-116.

47. Li, Q.; Rudolph, V.; Weigl, B.; Earl, A. Interparticle van der waals force in powder

flowability and compactibility. Int. J. Pharm. 2004, 280, 77-93.

48. Kawashima, Y.; Okumura, M.; Takenaka, H.; Kojima, A. Direct preparation of

spherically agglomerated salicylic acid crystals during crystallization. J. Pharm. Sci. 1984,

73, 1535-1538.

49. Kenji, M.; Yoshiaki, K.; Hirofumi, T.; Toshiyuki, N.; Tomoaki, H.; Yoichi, K.

Tabletting properties of bucillamine agglomerates prepared by the spherical crystallization

technique. Int. J. Pharm. 1994, 105, 11-18.

50. Chen, M.; Du, S.; Zhang, T.; Wang, Q.; Wu, S.; Gong, J. Spherical crystallization

and the mechanism of clopidogrel hydrogen sulfate. Chem. Eng. Technol. 2018, 41, 1259-

1265.

51. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

52. Joiris, E.; Martino, P. D.; Berneron, C.; Guyot-Hermann, A.-M.; Guyot, J.-C.

Compression behavior of orthorhombic paracetamol. Pharm. Res. 1998, 15, 1122-1130.

53. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

54. Malkowska, S.; Khan, K. A. Effect of re-conpression on the properties of tablets

prepared by dry granulation. Drug Dev. Ind. Pharm. 1983, 9, 331-347.

55. Arndt, O. R.; Baggio, R.; Adam, A. K.; Harting, J.; Franceschinis, E.; Kleinebudde,

P. Impact of different dry and wet granulation techniques on granule and tablet properties:

A comparative study. J. Pharm. Sci. 2018, 107, 3143-3152.

56. Toshiyuki Niwa, H. T., Tomoaki Hino, Akira Itoh, Yoshiaki Kawashima, Katsumi

Kiuchi. Preparation of agglomerated crystals for direct tabletting and microencapsulation

by the spherical crystallization technique with a continuous system. Pharm. Res. 1994, 11,

478-484.

57. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. V. Improvement of

dissolution and bioavailability of direct compressed tablets prepared using tolbutamide

agglomerated crystals. Chem. Pharm. Bull. 1992, 40, 3030-3035.

Page 242: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

220

58. Martino, P. D.; Cristofaro, R. D.; Barthelemy, C.; Joiris, E.; Filippo, G. P.; Sante,

M. Improved compression properties of propyphenazone spherical crystals. Int. J. Pharm.

2000, 197, 95-106.

59. Kawashima, Y.; Imai, M.; Takeuchi, H.; Yamamoto, H.; Kamiya, K.; Hino, T.

Improved flowability and compactibility of spherically agglomerated crystals of ascorbic

acid for direct tableting designed by spherical crystallization process. Powder Technol.

2003, 130, 283-289.

60. Hutchins A, M. B., Mullerney MP. Assessing tablet sticking propensity. Pharm.

Tech. 2012, 36, 31-34.

61. Paul, S.; Wang, C.; Wang, K.; Sun, C. C. Reduced punch sticking propensity of

acesulfame by salt formation: Role of crystal mechanical property and surface chemistry.

Mol. Pharm. 2019, 16, 2700-2707.

62. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.;

Meenan, P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J.

Pharm. Sci. 2017, 106, 151-158.

63. Samiei, L.; Kelly, K.; Taylor, L.; Forbes, B.; Collins, E.; Rowland, M. The

influence of electrostatic properties on the punch sticking propensity of pharmaceutical

blends. Powder Technol. 2017, 305, 509-517.

64. Al-Karawi, C.; Leopold, C. S. A comparative study on the sticking tendency of

ibuprofen and ibuprofen sodium dihydrate to differently coated tablet punches. Eur. J.

Pharm. Biopharm. 2018, 128, 107-118.

65. Simmons, D. M.; Gierer, D. S. A material sparing test to predict punch sticking

during formulation development. Drug Dev. Ind. Pharm. 2012, 38, 1054-1060.

66. Bejugam, N. K.; Mutyam, S. K.; Shankar, G. N. Tablet formulation of an active

pharmaceutical ingredient with a sticking and filming problem: Direct compression and

dry granulation evaluations. Drug Dev. Ind. Pharm. 2015, 41, 333-341.

67. Zhang, Y.; Maier, W. J.; Miller, R. M. Effect of rhamnolipids on the dissolution,

bioavailability, and biodegradation of phenanthrene. Environ. Sci. Technol. 1997, 31,

2211-2217.

68. Amidon, G. L.; Lennernäs, H.; Shah, V. P.; Crison, J. R. A theoretical basis for a

biopharmaceutic drug classification: The correlation of in vitro drug product dissolution

and in vivo bioavailability. Pharm. Res. 1995, 12, 413-420.

69. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

70. O'Driscoll, C. M.; Griffin, B. T. Biopharmaceutical challenges associated with

drugs with low aqueous solubility—the potential impact of lipid-based formulations. Adv.

Drug Deliv. Rev. 2008, 60, 617-624.

71. Hörter, D.; Dressman, J. B. Influence of physicochemical properties on dissolution

of drugs in the gastrointestinal tract. Adv. Drug Deliv. Rev. 2001, 46, 75-87.

Page 243: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

221

72. Savjani, K. T.; Gajjar, A. K.; Savjani, J. K. Drug solubility: Importance and

enhancement techniques. ISRN pharmaceutics 2012, 2012, 195727-195727.

73. Indulkar, A. S.; Lou, X.; Zhang, G. G. Z.; Taylor, L. S. Insights into the dissolution

mechanism of ritonavir–copovidone amorphous solid dispersions: Importance of

congruent release for enhanced performance. Mol. Pharm. 2019, 16, 1327-1339.

74. Cao, F.; Amidon, G. L.; Rodríguez-Hornedo, N.; Amidon, G. E. Mechanistic basis

of cocrystal dissolution advantage. J. Pharm. Sci. 2018, 107, 380-389.

75. Li, S.; Wong, S.; Sethia, S.; Almoazen, H.; Joshi, Y. M.; Serajuddin, A. T. M.

Investigation of solubility and dissolution of a free base and two different salt forms as a

function of ph. Pharm. Res. 2005, 22, 628-635.

76. Sun, J.; Wang, F.; Sui, Y.; She, Z.; Zhai, W.; Wang, C.; Deng, Y. Effect of particle

size on solubility, dissolution rate, and oral bioavailability: Evaluation using coenzyme q₁₀

as naked nanocrystals. Int. J. Nanomed. 2012, 7, 5733-5744.

77. Bolhuis, G. K.; Zuurman, K.; te Wierik, G. H. P. Improvement of dissolution of

poorly soluble drugs by solid deposition on a super disintegrant. Ii. The choice of super

disintegrants and effect of granulation. Eur. J. Pharm. Sci. 1997, 5, 63-69.

78. Kawashima, Y.; Lin, S. Y.; Ogawa, M.; Handa, T.; Takenaka, H. Preparations of

agglomerated crystals of polymorphic mixtures and a new complex of indomethacin—

epirizole by the spherical crystallization technique. J. Pharm. Sci. 1985, 74, 1152-1156.

79. Varshosaz, J.; Tavakoli, N.; Salamat, F. A. Enhanced dissolution rate of

simvastatin using spherical crystallization technique. Pharm. Dev. Technol. 2011, 16, 529-

535.

80. Peña, R.; Nagy, Z. K. Process intensification through continuous spherical

crystallization using a two-stage mixed suspension mixed product removal (msmpr) system.

Cryst. Growth Des. 2015, 15, 4225-4236.

81. Konstantinov, K. B.; Cooney, C. L. White paper on continuous bioprocessing. May

20–21, 2014 continuous manufacturing symposium. J. Pharm. Sci. 2015, 104, 813-820.

82. Mascia, S.; Heider, P. L.; Zhang, H.; Lakerveld, R.; Benyahia, B.; Barton, P. I.;

Braatz, R. D.; Cooney, C. L.; Evans, J. M. B.; Jamison, T. F.; Jensen, K. F.; Myerson, A.

S.; Trout, B. L. End-to-end continuous manufacturing of pharmaceuticals: Integrated

synthesis, purification, and final dosage formation. Angew. Chem. 2013, 52, 12359-12363.

83. Lee, S. In Present and future for continuous manufacturing: Fda perspective,

Proceedings of the 3rd FDA/PQRI Conference on Advancing Product Quality, March,

2017; pp 22-24.

84. Lee, S. L.; O’Connor, T. F.; Yang, X.; Cruz, C. N.; Chatterjee, S.; Madurawe, R.

D.; Moore, C. M. V.; Yu, L. X.; Woodcock, J. Modernizing pharmaceutical manufacturing:

From batch to continuous production. J. Pharm. Innov. 2015, 10, 191-199.

85. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

Page 244: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

222

86. Tahara, K.; Kono, Y.; Myerson, A. S.; Takeuchi, H. Development of continuous

spherical crystallization to prepare fenofibrate agglomerates with impurity complexation

using mixed-suspension, mixed-product removal crystallizer. Cryst. Growth Des. 2018, 18,

6448-6454.

87. Lee, T.; Chen, J. W.; Lee, H. L.; Lin, T. Y.; Tsai, Y. C.; Cheng, S.-L.; Lee, S.-W.;

Hu, J.-C.; Chen, L.-T. Stabilization and spheroidization of ammonium nitrate: Co-

crystallization with crown ethers and spherical crystallization by solvent screening. Chem.

Eng. J. 2013, 225, 809-817.

88. Wang, J. R.; Bao, J.; Fan, X.; Dai, W.; Mei, X. Ph-switchable vitamin b9 gels for

stoichiometry-controlled spherical co-crystallization. Chem. Commun. 2016, 52, 13452-

13455.

Page 245: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

223

Chapter 2.

1. Jivraj, M.; Martini, L. G.; Thomson, C. M. An overview of the different excipients

useful for the direct compression of tablets. Pharm Sci Technolo Today 2000, 3, 58-63.

2. Hirschberg, C.; Sun, C. C.; Rantanen, J. Analytical method development for

powder characterization: Visualization of the critical drug loading affecting the

processability of a formulation for direct compression. J. Pharm. Biomed. Anal. 2016, 128,

462-468.

3. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine

active pharmaceutical ingredient powder blends and tablets at high drug loading via dry

particle coating. Int. J. Pharm. 2018, 543, 288-299.

4. Capece, M.; Huang, Z.; Davé, R. Insight into a novel strategy for the design of

tablet formulations intended for direct compression. J. Pharm. Sci. 2017, 106, 1608-1617.

5. Martinello, T.; Kaneko, T. M.; Velasco, M. V. R.; Taqueda, M. E. S.; Consiglieri,

V. O. Optimization of poorly compactable drug tablets manufactured by direct

compression using the mixture experimental design. Int. J. Pharm. 2006, 322, 87-95.

6. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high

drug loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

7. Thoorens, G.; Krier, F.; Leclercq, B.; Carlin, B.; Evrard, B. Microcrystalline

cellulose, a direct compression binder in a quality by design environment—a review. Int.

J. Pharm. 2014, 473, 64-72.

8. Chattoraj, S.; Sun, C. C. Crystal and particle engineering strategies for improving

powder compression and flow properties to enable continuous tablet manufacturing by

direct compression. J. Pharm. Sci. 2018, 107, 968-974.

9. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

10. Geldart, D. Types of gas fluidization. Powder Technol. 1973, 7, 285-292.

11. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization

of aceclofenac and paracetamol spherical crystals and their tableting properties. Powder

Technol. 2015, 274, 446-454.

12. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

13. Fadke, J.; Desai, J.; Thakkar, H. Formulation development of spherical crystal

agglomerates of itraconazole for preparation of directly compressible tablets with enhanced

bioavailability. AAPS PharmSciTech 2015, 16, 1434-1444.

14. Yoshiaki Kawashima, M. O. Spherical crystallization direct spherical

agglomeration of salicylic acid crystals during crystallization. Science 1982, 216, 1127-

1128.

15. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

Page 246: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

224

16. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

17. Kovačič, B.; Vrečer, F.; Planinšek, O. Spherical crystallization of drugs. Acta

Pharm 2012, 62, 1-14.

18. Leon, R. A. L.; Wan, W. Y.; Badruddoza, A. Z. M.; Hatton, T. A.; Khan, S. A.

Simultaneous spherical crystallization and co-formulation of drug(s) and excipient from

microfluidic double emulsions. Cryst. Growth Des. 2014, 14, 140-146.

19. Yoshiaki Kawashima, T. N. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

20. Sgarbossa, A.; Giacomazza, D.; Di Carlo, M. Ferulic acid: A hope for alzheimer’s

disease therapy from plants. Nutrients 2015, 7, 5764-5782.

21. Janicke, B.; Hegardt, C.; Krogh, M.; Onning, G.; Akesson, B.; Cirenajwis, H. M.;

Oredsson, S. M. The antiproliferative effect of dietary fiber phenolic compounds ferulic

acid and p-coumaric acid on the cell cycle of caco-2 cells. Nutr. Cancer 2011, 63, 611-22.

22. Pagidipati, N. J.; Gaziano, T. A. Estimating deaths from cardiovascular disease: A

review of global methodologies of mortality measurement. Circulation 2013, 127, 749-56.

23. Prabhakar, P. K.; Prasad, R.; Ali, S.; Doble, M. Synergistic interaction of ferulic

acid with commercial hypoglycemic drugs in streptozotocin induced diabetic rats.

Phytomedicine 2013, 20, 488-494.

24. Saija, A.; Tomaino, A.; Trombetta, D.; De Pasquale, A.; Uccella, N.; Barbuzzi, T.;

Paolino, D.; Bonina, F. In vitro and in vivo evaluation of caffeic and ferulic acids as topical

photoprotective agents. Int. J. Pharm. 2000, 199, 39-47.

25. Sun, C. C. Materials science tetrahedron--a useful tool for pharmaceutical research

and development. J. Pharm. Sci. 2009, 98, 1671-1687.

26. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

27. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int J Pharm 2015, 484, 146-55.

28. Paul, S.; Chang, S.-Y.; Dun, J.; Sun, W.-J.; Wang, K.; Tajarobi, P.; Boissier, C.;

Sun, C. C. Comparative analyses of flow and compaction properties of diverse mannitol

and lactose grades. Int. J. Pharm. 2018, 546, 39-49.

29. Kawashima, K. M. a. Y. Micromeritic characteristics and agglomeration

mechanisms in the spherical crystallization of bucillamine by the spherical agglomeration

and the emulsion solvent diffusion methods. Powder Technol. 1993, 76, 57-64.

30. Cui, F.; Yang, M.; Jiang, Y.; Cun, D.; Lin, W.; Fan, Y.; Kawashima, Y. Design of

sustained-release nitrendipine microspheres having solid dispersion structure by quasi-

emulsion solvent diffusion method. J. Control. Release 2003, 91, 375-384.

Page 247: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

225

31. Kumar, N.; Pruthi, V. Structural elucidation and molecular docking of ferulic acid

from parthenium hysterophorus possessing cox-2 inhibition activity. 3 Biotech 2015, 5,

541-551.

32. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J Adhes Sci Technol. 2011, 25, 483-499.

33. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

34. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

35. WHO. Revision of monograph of tablets. 2011, Document QAS/09.324/Final.

36. USP. Usp41(online), general chapters. Weight variation of dietary supplements-

tablet. Rockville: United states pharmacopoeial convention. 2018.

37. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int. J. Pharm. 2015, 484, 146-155.

38. USP. Usp41(online), general chapters. Tablet friability. Rockville: United states

pharmacopoeial convention. 2018.

39. USP. Usp34 (online), general chapter. Disintegration and dissolution of dietary

supplements. Rockville: United states pharmacopoeial convention. 2011.

40. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.;

Meenan, P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J.

Pharm. Sci. 2017, 106, 151-158.

41. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through

excipient selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

Page 248: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

226

Chapter 3.

1. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

2. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through

excipient selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

3. Kristensen, H. G.; Schaefer, T. Granulation: A review on pharmaceutical wet-

granulation. Drug Dev. Ind. Pharm. 1987, 13, 803-872.

4. Rouan, E.; Rhodes, C. T. The spray drying of pharmaceuticals. Drug Dev. Ind.

Pharm. 1992, 18, 1169-1206.

5. Remenar, J. F.; Morissette, S. L.; Peterson, M. L.; Moulton, B.; MacPhee, J. M.;

Guzmán, H. R.; Almarsson, Ö. Crystal engineering of novel cocrystals of a triazole drug

with 1,4-dicarboxylic acids. J. Am. Chem. Soc. 2003, 125, 8456–8457.

6. Chattoraj, S.; Sun, C. C. Crystal and particle engineering strategies for improving

powder compression and flow properties to enable continuous tablet manufacturing by

direct compression. J. Pharm. Sci. 2017, 107, 968-974.

7. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

8. Maghsoodi, M.; Taghizadeh, O.; Martin, G. P.; Nokhodchi, A. Particle design of

naproxen-disintegrant agglomerates for direct compression by a crystallo-co-

agglomeration technique. Int. J. Pharm. 2008, 351, 45-54.

9. Geldart, D. Types of gas fluidization. Powder Technol. 1973, 7, 285-292.

10. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

11. Kim, E. H.; Chen, X. D.; Pearce, D. Effect of surface composition on the

flowability of industrial spray-dried dairy powders. Colloids Surf. B 2005, 46, 182-187.

12. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

13. Vromans, H.; Bolhuis, G. K.; Lerk, C. F.; van de Biggelaar, H.; Bosch, H. Studies

on tableting properties of lactose. Vii. The effect of variations in primary particle size and

percentage of amorphous lactose in spray dried lactose products. Int. J. Pharm. 1987, 35,

29-37.

14. Jain, S. Mechanical properties of powders for compaction and tableting: An

overview. Pharm. Sci. Technol. Today 1999, 2, 20-31.

15. Jones, W.; Motherwell, W. D. S.; Trask, A. V. Pharmaceutical cocrystals: An

emerging approach to physical property enhancement. MRS Bulletin 2011, 31, 875-879.

Page 249: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

227

16. Chang, S. Y.; Sun, C. C. Superior plasticity and tabletability of theophylline

monohydrate. Mol. Pharm. 2017, 14, 2047-2055.

17. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

18. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

19. Martino, P. D.; Cristofaro, R. D.; Barthelemy, C.; Joiris, E.; Filippo, G. P.; Sante,

M. Improved compression properties of propyphenazone spherical crystals. Int. J. Pharm.

2000, 197, 95-106.

20. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

21. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

22. Kawashima, Y.; Okumura, M.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

23. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

24. Nokhodchi, A.; Maghsoodi, M. Preparation of spherical crystal agglomerates of

naproxen containing disintegrant for direct tablet making by spherical crystallization

technique. AAPS PharmSciTech 2008, 9, 54-59.

25. Kawashima, Y.; Imai, M.; Takeuchi, H.; Yamamoto, H.; Kamiya, K.; Hino, T.

Improved flowability and compactibility of spherically agglomerated crystals of ascorbic

acid for direct tableting designed by spherical crystallization process. Powder Technol.

2003, 130, 283-289.

26. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization

of aceclofenac and paracetamol spherical crystals and their tableting properties. Powder

Technol. 2015, 274, 446-454.

27. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

28. Thati, J.; Rasmuson, A. C. Particle engineering of benzoic acid by spherical

agglomeration. Eur. J. Pharm. Sci. 2012, 45, 657-667.

29. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

30. Zhang, H.; Chen, Y.; Wang, J.; Gong, J. Investigation on the spherical

crystallization process of cefotaxime sodium. Ind. Eng. Chem. Res. 2010, 49, 1402-1411.

Page 250: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

228

31. Su, Y.; Xu, J.; Shi, Q.; Yu, L.; Cai, T. Polymorphism of griseofulvin: Concomitant

crystallization from the melt and a single crystal structure of a metastable polymorph with

anomalously large thermal expansion. Chem. Commun. 2018, 54, 358-361.

32. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

33. Heckel, W. An analysis of powder compaction phenomena. Trans. Metall. Soc.

AIME 1961, 221, 671-675.

34. Kuentz, M.; Leuenberger, H. Pressure susceptibility of polymer tablets as a critical

property: A modified heckel equation. J. Pharm. Sci. 1999, 88, 174-179.

35. Chen, Y.; Liu, C.; Chen, Z.; Su, C.; Hageman, M.; Hussain, M.; Haskell, R.;

Stefanski, K.; Qian, F. Drug–polymer–water interaction and its implication for the

dissolution performance of amorphous solid dispersions. Mol. Pharm. 2015, 12, 576-589.

36. Tao, J.; Sun, Y.; Zhang, G. G. Z.; Yu, L. J. P. R. Solubility of small-molecule

crystals in polymers: D-mannitol in pvp, indomethacin in pvp/va, and nifedipine in pvp/va.

Pharm. Res. 2009, 26, 855-864.

37. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

38. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

39. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

40. Chen, H.; Guo, Y.; Wang, C.; Dun, J.; Sun, C. C. Spherical cocrystallization - an

enabling technology for the development of high dose direct compression tablets of poorly

soluble drugs. Crys. Growth Des. 2019, 19, 2503-2510.

41. Wu, S. J.; Sun, C. Insensitivity of compaction properties of brittle granules to size

enlargement by roller compaction. J. Pharm. Sci. 2007, 96, 1445-50.

42. Paul, S.; Sun, C. C. The suitability of common compressibility equations for

characterizing plasticity of diverse powders. Int. J. Pharm. 2017, 532, 124-130.

43. Patel, S.; Sun, C. C. Macroindentation hardness measurement—modernization and

applications. Int. J. Pharm. 2016, 506, 262-267.

44. Shirotani, K.-i.; Suzuki, E.; Morita, Y.; Sekiguchi, K. Solvate formation of

griseofulvin with alkyl halide and alkyl dihalides. Chem. Pharm. Bull. 1988, 36, 4045-

4054.

45. Fachaux, J. M.; Guyot Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Huvenne, J. P.; Bouche, R. Compression ability improvement by solvation/desolvation

process: Application to paracetamol for direct compression. Int. J. Pharm. 1993, 99, 99-

107.

Page 251: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

229

46. Fachaux, J. M.; Guyot-Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Veesler, S.; Boistelle, R. Pure paracetamol for direct compression part i. Development of

sintered-like crystals of paracetamol. Powder Technol. 1995, 82, 123-128.

47. Fachaux, J. M.; Guyot-Hermann, A. M.; Guyot, J. C.; Conflant, P.; Drache, M.;

Veesler, S.; Boistelle, R. Pure paracetamol for direct compression part ii. Study of the

physicochemical and mechanical properties of sintered-like crystals of paracetamol.

Powder Technol. 1995, 82, 129-133.

48. Sekiguchi, K.; Suzuki, E.; Tsuda, Y.; Morita, Y. Thermal analysis of griseofulvin-

chloroform system by high pressure differential scanning calorimetry. Chem. Pharm. Bull.

1983, 31, 2139-2141.

49. Sekiguchi, K.; Horikoshi, I.; Himuro, I. Studies on the method of size reduction of

medicinal compounds. Iii. Size reduction of griseofulvin by solvation and desolvation

method using chloroform. (3). Chem. Pharm. Bull. 1968, 16, 2495-2502.

50. Sekiguchi, K.; Ito, K.; Owada, E.; Ueno, K. Studies on the method of size reduction

of medicinal compounds. Ii. Size reduction of griseofulvin by solvation and desolvation

method using chloroform (2). Chem. Pharm. Bull. 1964, 12, 1192-1197.

Page 252: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

230

Chapter 4.

1. Sun, W.-J.; Aburub, A.; Sun, C. C. Particle engineering for enabling a formulation

platform suitable for manufacturing low-dose tablets by direct compression. J. Pharm. Sci.

2017, 106, 1772-1777.

2. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high

drug loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

3. Lakshman, J. P.; Kowalski, J.; Vasanthavada, M.; Tong, W. Q.; Joshi, Y. M.;

Serajuddin, A. T. Application of melt granulation technology to enhance tabletting

properties of poorly compactible high-dose drugs. J. Pharm. Sci. 2011, 100, 1553-1565.

4. Sun, W.-J.; Chen, H.; Aburub, A.; Sun, C. C. A platformdirect compression

formulation for low dose sustained-release tablets enabled by a dual particle engineering

approach. Powder Technol. 2019, 342, 856-863.

5. Babu, N. J.; Nangia, A. Solubility advantage of amorphous drugs and

pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662-2679.

6. Rabinow, B.; Kipp, J.; Papadopoulos, P.; Wong, J.; Glosson, J.; Gass, J.; Sun, C.

S.; Wielgos, T.; White, R.; Cook, C.; Barker, K.; Wood, K. Itraconazole iv nanosuspension

enhances efficacy through altered pharmacokinetics in the rat. Int. J. Pharm. 2007, 339,

251-60.

7. Keck, C. M.; Muller, R. H. Drug nanocrystals of poorly soluble drugs produced by

high pressure homogenisation. Eur. J. Pharm. Biopharm. 2006, 62, 3-16.

8. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

9. Sastry, S. V.; Nyshadham, J. R.; Fix, J. A. Recent technological advances in oral

drug delivery – a review. Pharm. Sci. Technolo. Today 2000, 3, 138-145.

10. Williams, H. D.; Trevaskis, N. L.; Charman, S. A.; Shanker, R. M.; Charman, W.

N.; Pouton, C. W.; Porter, C. J. H. Strategies to address low drug solubility in discovery

and development. Pharmacol. Rev. 2013, 65, 315-499.

11. Hickey, M. B.; Peterson, M. L.; Scoppettuolo, L. A.; Morrisette, S. L.; Vetter, A.;

Guzman, H.; Remenar, J. F.; Zhang, Z.; Tawa, M. D.; Haley, S.; Zaworotko, M. J.;

Almarsson, O. Performance comparison of a co-crystal of carbamazepine with marketed

product. Eur. J. Pharm. Biopharm. 2007, 67, 112-119.

12. Martinello, T.; Kaneko, T. M.; Velasco, M. V. R.; Taqueda, M. E. S.; Consiglieri,

V. O. Optimization of poorly compactable drug tablets manufactured by direct

compression using the mixture experimental design. Int. J. Pharm. 2006, 322, 87-95.

13. Patra, C. N.; Swain, S.; Mahanty, S.; Panigrahi, K. C. Design and characterization

of aceclofenac and paracetamol spherical crystals and their tableting properties. Powder

Technol. 2015, 274, 446-454.

14. Thakur, A.; Thipparaboina, R.; Kumar, D.; Sai Gouthami, K.; Shastri, N. R. Crystal

engineered albendazole with improved dissolution and material attributes. CrystEngComm

2016, 18, 1489-1494.

Page 253: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

231

15. Fadke, J.; Desai, J.; Thakkar, H. Formulation development of spherical crystal

agglomerates of itraconazole for preparation of directly compressible tablets with enhanced

bioavailability. AAPS PharmSciTech 2015, 16, 1434-1444.

16. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science 1982, 216,

1127-1128.

17. Lamešić, D.; Planinšek, O.; Lavrič, Z.; Ilić, I. Spherical agglomerates of lactose

with enhanced mechanical properties. Int. J. Pharm. 2017, 516, 247-257.

18. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion metho. Chem. Pharm. Bull. 1990, 38, 733-739.

19. Kawashima, Y.; Cui, F.; Takeuchi, H.; Niwa, T.; Hino, T.; Kiuchi, K.

Improvements in flowability and compressibility of pharmaceutical crystals for direct

tabletting by spherical crystallization with a two-solvent system. Powder Technol. 1994,

78, 151-157.

20. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

21. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient - a tale of spherical crystallization. J. Pharm. Sci. 2018, 108, 1396-1400.

22. Lee, T.; Chen, J. W.; Lee, H. L.; Lin, T. Y.; Tsai, Y. C.; Cheng, S.-L.; Lee, S.-W.;

Hu, J.-C.; Chen, L.-T. Stabilization and spheroidization of ammonium nitrate: Co-

crystallization with crown ethers and spherical crystallization by solvent screening. Chem.

Eng. J. 2013, 225, 809-817.

23. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

24. Wang, J. R.; Bao, J.; Fan, X.; Dai, W.; Mei, X. Ph-switchable vitamin b9 gels for

stoichiometry-controlled spherical co-crystallization. Chem. Commun. 2016, 52, 13452-

13455.

25. Aitipamula, S.; Vangala, V. R.; Chow, P. S.; Tan, R. B. H. Cocrystal hydrate of an

antifungal drug, griseofulvin, with promising physicochemical properties. Cryst Growth

Des. 2012, 12, 5858-5863.

26. Talukder, R.; Reed, C.; Durig, T.; Hussain, M. Dissolution and solid-state

characterization of poorly water-soluble drugs in the presence of a hydrophilic carrier.

AAPS PharmSciTech 2011, 12, 1227-1233.

27. Chen, Y.; Liu, C.; Chen, Z.; Su, C.; Hageman, M.; Hussain, M.; Haskell, R.;

Stefanski, K.; Qian, F. Drug–polymer–water interaction and its implication for the

dissolution performance of amorphous solid dispersions. Mol. Pharm. 2015, 12, 576-589.

Page 254: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

232

28. Tao, J.; Sun, Y.; Zhang, G. G. Z.; Yu, L. J. P. R. Solubility of small-molecule

crystals in polymers: D-mannitol in pvp, indomethacin in pvp/va, and nifedipine in pvp/va.

Pharm. Res. 2009, 26, 855-864.

29. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

30. Bryant, M. J.; Maloney, A. G. P.; Sykes, R. A. Predicting mechanical properties

of crystalline materials through topological analysis. CrystEngComm 2018, 20, 2698-2704.

31. Sun, C. C.; Himmelspach, M. W. Reduced tabletability of roller compacted

granules as a result of granule size enlargement. J. Pharm. Sci. 2006, 95, 200-206.

32. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

33. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

34. Paul, S.; Chang, S.-Y.; Sun, C. C. The phenomenon of tablet flashing — its impact

on tableting data analysis and a method to eliminate it. Powder Technol. 2017, 305, 117-

124.

35. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int. J. Pharm. 2015, 484, 146-155.

36. Wang, C.; Chopade, S. A.; Guo, Y.; Early, J. T.; Tang, B.; Wang, E.; Hillmyer, M.

A.; Lodge, T. P.; Sun, C. C. Preparation, characterization, and formulation development

of drug-drug protic ionic liquids of diphenhydramine with ibuprofen and naproxen. Mol.

Pharm. 2018, 15, 4190-4201.

37. Zhang, Y.; Huo, M.; Zhou, J.; Xie, S. Pksolver: An add-in program for

pharmacokinetic and pharmacodynamic data analysis in microsoft excel. Comput. Methods

Programs Biomed. 2010, 99, 306-314.

38. Jitkar, S.; Thipparaboina, R.; Chavan, R. B.; Shastri, N. R. Spherical agglomeration

of platy crystals: Curious case of etodolac. Cryst. Growth Des. 2016, 16, 4034-4042.

39. Rajniak, P.; Mancinelli, C.; Chern, R. T.; Stepanek, F.; Farber, L.; Hill, B. T.

Experimental study of wet granulation in fluidized bed: Impact of the binder properties on

the granule morphology. Int. J. Pharm. 2007, 334, 92-102.

40. Skinner, G. W.; Harcum, W. W.; Barnum, P. E.; Guo, J.-H. The evaluation of fine-

particle hydroxypropylcellulose as a roller compaction binder in pharmaceutical

applications. Drug Dev. Ind. Pharm. 1999, 25, 1121-1128.

41. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

42. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

Page 255: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

233

43. Sun, C.; Grant, D. J. W. Influence of crystal structure on the tableting properties

of sulfamerazine polymorphs. Pharm. Res. 2001, 18, 274-280.

44. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

45. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

46. Paul, S.; Sun, C. C. Dependence of friability on tablet mechanical properties and a

predictive approach for binary mixtures. Pharm. Res. 2017, 34, 2901-2909.

47. Sun, C. C. A classification system for tableting behaviors of binary powder

mixtures. Asian J. Pharm. 2016, 11, 486-491.

48. Wu, C. Y.; Best, S. M.; Bentham, A. C.; Hancock, B. C.; Bonfield, W. Predicting

the tensile strength of compacted multi-component mixtures of pharmaceutical powders.

Pharm. Res. 2006, 23, 1898-1905.

49. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

Page 256: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

234

Chapter 5.

1. Van Snick, B.; Holman, J.; Cunningham, C.; Kumar, A.; Vercruysse, J.; De Beer,

T.; Remon, J. P.; Vervaet, C. Continuous direct compression as manufacturing platform

for sustained release tablets. Int J. Pharm. 2017, 519, 390-407.

2. Jivraj, M.; Martini, L. G.; Thomson, C. M. An overview of the different excipients

useful for the direct compression of tablets. Pharm. Sci. Technol. Today 2000, 3, 58-63.

3. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

4. Chen, L.; Ding, X.; He, Z.; Huang, Z.; Kunnath, K. T.; Zheng, K.; Davé, R. N.

Surface engineered excipients: I. Improved functional properties of fine grade

microcrystalline cellulose. Int. J. Pharm. 2018, 536, 127-137.

5. Kawashima, Y.; Okumura, O.; Takenaka, H. Spherical crystallization direct

spherical agglomeration of salicylic acid crystals during crystallization. Science (New York,

N.Y.) 1982, 216, 1127-1128.

6. Chen, H.; Wang, C.; Sun, C. C. Profoundly improved plasticity and tabletability

of griseofulvin by in situ solvation and desolvation during spherical crystallization. Cryst.

Growth Des. 2019, 19, 2350-2357.

7. Peña, R.; Nagy, Z. K. Process intensification through continuous spherical

crystallization using a two-stage mixed suspension mixed product removal (msmpr) system.

Cryst. Growth Des. 2015, 15, 4225-4236.

8. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

9. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

10. Chen, H.; Wang, C.; Kang, H.; Zhi, B.; Haynes, C. L.; Aburub, A.; Sun, C. C.

Microstructures and pharmaceutical properties of ferulic acid agglomerates prepared by

different spherical crystallization methods. Int. J. Pharm. 2020, 574, 118914.

11. Saboo, S.; Mugheirbi, N. A.; Zemlyanov, D. Y.; Kestur, U. S.; Taylor, L. S.

Congruent release of drug and polymer: A “sweet spot” in the dissolution of amorphous

solid dispersions. J. Control. Release 2019, 298, 68-82.

12. Serajuddin, A. T. Salt formation to improve drug solubility. Adv. Drug Deliv. Rev.

2007, 59, 603-616.

13. Blagden, N.; de Matas, M.; Gavan, P. T.; York, P. Crystal engineering of active

pharmaceutical ingredients to improve solubility and dissolution rates. Adv. Drug Deliv.

Rev 2007, 59, 617-630.

14. Han, R.; Xiong, H.; Ye, Z.; Yang, Y.; Huang, T.; Jing, Q.; Lu, J.; Pan, H.; Ren, F.;

Ouyang, D. Predicting physical stability of solid dispersions by machine learning

techniques. J. Control. Release 2019, 311-312, 16-25.

Page 257: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

235

15. Indulkar, A. S.; Lou, X.; Zhang, G. G. Z.; Taylor, L. S. Insights into the dissolution

mechanism of ritonavir–copovidone amorphous solid dispersions: Importance of

congruent release for enhanced performance. Mol. Pharm. 2019, 16, 1327-1339.

16. Zheng, K.; Lin, Z.; Capece, M.; Kunnath, K.; Chen, L.; Davé, R. N. Effect of

particle size and polymer loading on dissolution behavior of amorphous griseofulvin

powder. J. Pharm. Sci. 2019, 108, 234-242.

17. Aitipamula, S.; Banerjee, R.; Bansal, A. K.; Biradha, K.; Cheney, M. L.;

Choudhury, A. R.; Desiraju, G. R.; Dikundwar, A. G.; Dubey, R.; Duggirala, N.; Ghogale,

P. P.; Ghosh, S.; Goswami, P. K.; Goud, N. R.; Jetti, R. R. K. R.; Karpinski, P.; Kaushik,

P.; Kumar, D.; Kumar, V.; Moulton, B.; Mukherjee, A.; Mukherjee, G.; Myerson, A. S.;

Puri, V.; Ramanan, A.; Rajamannar, T.; Reddy, C. M.; Rodriguez-Hornedo, N.; Rogers, R.

D.; Row, T. N. G.; Sanphui, P.; Shan, N.; Shete, G.; Singh, A.; Sun, C. C.; Swift, J. A.;

Thaimattam, R.; Thakur, T. S.; Kumar Thaper, R.; Thomas, S. P.; Tothadi, S.; Vangala, V.

R.; Variankaval, N.; Vishweshwar, P.; Weyna, D. R.; Zaworotko, M. J. Polymorphs, salts,

and cocrystals: What’s in a name? Crys. Growth Des. 2012, 12, 2147-2152.

18. Drozd, K. V.; Manin, A. N.; Churakov, A. V.; Perlovich, G. L. Novel drug–drug

cocrystals of carbamazepine with para-aminosalicylic acid: Screening, crystal structures

and comparative study of carbamazepine cocrystal formation thermodynamics.

CrystEngComm 2017, 19, 4273-4286.

19. Chen, H.; Guo, Y.; Wang, C.; Dun, J.; Sun, C. C. Spherical cocrystallization—an

enabling technology for the development of high dose direct compression tablets of poorly

soluble drugs. Cryst. Growth Des. 2019, 19, 2503-2510.

20. Pagire, S. K.; Korde, S. A.; Whiteside, B. R.; Kendrick, J.; Paradkar, A. Spherical

crystallization of carbamazepine/saccharin co-crystals: Selective agglomeration and

purification through surface interactions. Crys. Growth Des. 2013, 13, 4162-4167.

21. Sun, C. C. Decoding powder tabletability: Roles of particle adhesion and plasticity.

J. Adhes. Sci. Technol. 2011, 25, 483-499.

22. Kakran, M.; Sahoo, N. G.; Tan, I.-L.; Li, L. Preparation of nanoparticles of poorly

water-soluble antioxidant curcumin by antisolvent precipitation methods. J. Nanopart. Res.

2012, 14, 757.

23. Chun, N.-H.; Wang, I.-C.; Lee, M.-J.; Jung, Y.-T.; Lee, S.; Kim, W.-S.; Choi, G. J.

Characteristics of indomethacin–saccharin (imc–sac) co-crystals prepared by an anti-

solvent crystallization process. Eur. J. Pharm. Biopharm. 2013, 85, 854-861.

24. Alhalaweh, A.; Roy, L.; Rodríguez-Hornedo, N.; Velaga, S. P. Ph-dependent

solubility of indomethacin–saccharin and carbamazepine–saccharin cocrystals in aqueous

media. Mol. Pharm. 2012, 9, 2605-2612.

25. Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of gases in multimolecular

layers. J. Am. Chem. Soc. 1938, 60, 309-319.

26. Chen, M.-H.; Bergman, C. J. Method for determining the amylose content,

molecular weights, and weight- and molar-based distributions of degree of polymerization

of amylose and fine-structure of amylopectin. Carbohydr. Polym. 2007, 69, 562-578.

Page 258: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

236

27. Basavoju, S.; Boström, D.; Velaga, S. P. Indomethacin–saccharin cocrystal:

Design, synthesis and preliminary pharmaceutical characterization. Pharm. Res. 2008, 25,

530-541.

28. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A. Energy

frameworks: Insights into interaction anisotropy and the mechanical properties of

molecular crystals. Chem. Comm. 2015, 51, 3735-3738.

29. Wang, C.; Sun, C. C. Identifying slip planes in organic polymorphs by combined

energy framework calculations and topology analysis. Cryst. Growth Des. 2018, 18, 1909-

1916.

30. Turner, M. J.; Grabowsky, S.; Jayatilaka, D.; Spackman, M. A. Accurate and

efficient model energies for exploring intermolecular interactions in molecular crystals. J.

Phys. Chem. Lett. 2014, 5, 4249-4255.

31. Bryant, M. J.; Maloney, A. G. P.; Sykes, R. A. Predicting mechanical properties

of crystalline materials through topological analysis. CrystEngComm 2018, 20, 2698-2704.

32. Gaussian 16, revision b.01, m. J. Frisch, g. W. Trucks, h. B. Schlegel, g. E. Scuseria,

m. A. Robb, j. R. Cheeseman, g. Scalmani, v. Barone, g. A. Petersson, h. Nakatsuji, x. Li,

m. Caricato, a. V. Marenich, j. Bloino, b. G. Janesko, r. Gomperts, b. Mennucci, h. P.

Hratchian, j. V. Ortiz, a. F. Izmaylov, j. L. Sonnenberg, d. Williams-young, f. Ding, f.

Lipparini, f. Egidi, j. Goings, b. Peng, a. Petrone, t. Henderson, d. Ranasinghe, v. G.

Zakrzewski, j. Gao, n. Rega, g. Zheng, w. Liang, m. Hada, m. Ehara, k. Toyota, r. Fukuda,

j. Hasegawa, m. Ishida, t. Nakajima, y. Honda, o. Kitao, h. Nakai, t. Vreven, k. Throssell,

j. A. Montgomery, jr., j. E. Peralta, f. Ogliaro, m. J. Bearpark, j. J. Heyd, e. N. Brothers, k.

N. Kudin, v. N. Staroverov, t. A. Keith, r. Kobayashi, j. Normand, k. Raghavachari, a. P.

Rendell, j. C. Burant, s. S. Iyengar, j. Tomasi, m. Cossi, j. M. Millam, m. Klene, c. Adamo,

r. Cammi, j. W. Ochterski, r. L. Martin, k. Morokuma, o. Farkas, j. B. Foresman, and d. J.

Fox, gaussian, inc., wallingford ct, 2016.

33. Zhao, Y.; Truhlar, D. G. J. T. C. A. The m06 suite of density functionals for main

group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states,

and transition elements: Two new functionals and systematic testing of four m06-class

functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215-241.

34. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

35. Chen, L.; Ding, X.; He, Z.; Fan, S.; Kunnath, K. T.; Zheng, K.; Davé, R. N. Surface

engineered excipients: Ii. Simultaneous milling and dry coating for preparation of fine-

grade microcrystalline cellulose with enhanced properties. Int. J. Pharm. 2018, 546, 125-

136.

36. Shi, L.; Chattoraj, S.; Sun, C. C. Reproducibility of flow properties of

microcrystalline cellulose — avicel ph102. Powder Technol. 2011, 212, 253-257.

37. Schulze, D. Powders and bulk solids: Behavior, characterization, storage and flow,

springer, berlin, heidelberg, new york, tokyo, 2008.

38. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

Page 259: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

237

39. Osei-Yeboah, F.; Sun, C. C. Validation and applications of an expedited tablet

friability method. Int. J. Pharm. 2015, 484, 146-155.

40. Chen, H.; Paul, S.; Xu, H.; Wang, K.; Mahanthappa, M. K.; Sun, C. C. Reduction

of punch-sticking propensity of celecoxib by spherical crystallization via polymer assisted

quasi-emulsion solvent diffusion. Mol. Pharm. 2020, 1387–1396.

41. Ribardière, A.; Tchoreloff, P.; Couarraze, G.; Puisieux, F. Modification of

ketoprofen bead structure produced by the spherical crystallization technique with a two-

solvent system. Int. J. Pharm. 1996, 144, 195-207.

42. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

43. Chattoraj, S.; Shi, L.; Sun, C. C. Profoundly improving flow properties of a

cohesive cellulose powder by surface coating with nano-silica through comilling. J. Pharm.

Sci. 2011, 100, 4943-4952.

44. Kendall, K. Adhesion: Molecules and mechanics. Science (New York, N.Y.) 1994,

263, 1720-1725.

45. Hou, H.; Sun, C. C. Quantifying effects of particulate properties on powder flow

properties using a ring shear tester. J. Pharm. Sci. 2008, 97, 4030-4039.

46. Sun, C. C.; Kleinebudde, P. Mini review: Mechanisms to the loss of tabletability

by dry granulation. Eur. J. Pharm. Biopharm. 2016, 106, 9-14.

47. Wang, C.; Sun, C. C. Computational techniques for predicting mechanical

properties of organic crystals: A systematic evaluation. Mol. Pharm. 2019, 16, 1732-1741.

48. Sun, C. C.; Hou, H.; Gao, P.; Ma, C.; Medina, C.; Alvarez, F. J.; Hou, H.; Gao, P.

Development of a high drug load tablet formulation based on assessment of powder

manufacturability: Moving towards quality by design. J. Pharm. Sci. 2009, 98, 239-247.

49. Shi, L.; Sun, C. C. Overcoming poor tabletability of pharmaceutical crystals by

surface modification. Pharm. Res. 2011, 28, 3248-3255.

50. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

51. Shi, L.; Feng, Y.; Sun, C. C. Roles of granule size in over-granulation during high

shear wet granulation. J. Pharm. Sci. 2010, 99, 3322-3325.

52. Zhang, Y.; Law, Y.; Chakrabarti, S. Physical properties and compact analysis of

commonly used direct compression binders. AAPS PharmSciTech. 2003, 4, 489-499.

53. Sinka, I. C.; Cunningham, J. C.; Zavaliangos, A. Analysis of tablet compaction. Ii.

Finite element analysis of density distributions in convex tablets. J. Pharm. Sci. 2004, 93,

2040-53.

Page 260: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

238

Chapter 6.

1. Chen, H.; Aburub, A.; Sun, C. C. Direct compression tablet containing 99% active

ingredient—a tale of spherical crystallization. J. Pharm. Sci. 2019, 108, 1396-1400.

2. Lakshman, J. P.; Kowalski, J.; Vasanthavada, M.; Tong, W. Q.; Joshi, Y. M.;

Serajuddin, A. T. Application of melt granulation technology to enhance tabletting

properties of poorly compactible high-dose drugs. J. Pharm. Sci. 2011, 100, 1553-1565.

3. Cai, L.; Farber, L.; Zhang, D.; Li, F.; Farabaugh, J. A new methodology for high

drug loading wet granulation formulation development. Int. J. Pharm. 2013, 441, 790-800.

4. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J.; Dawson, N.; Mullarney, M. P.;

Meenan, P.; Sun, C. C. Mechanism and kinetics of punch sticking of pharmaceuticals. J.

Pharm. Sci. 2017, 106, 151-158.

5. Wang, J. J.; Guillot, M. A.; Bateman, S. D.; Morris, K. R. Modeling of adhesion

in tablet compression. Ii. Compaction studies using a compaction simulator and an

instrumented tablet press. J. Pharm. Sci. 2004, 93, 407-417.

6. Chattoraj, S.; Daugherity, P.; McDermott, T.; Olsofsky, A.; Roth, W. J.; Tobyn, M.

Sticking and picking in pharmaceutical tablet compression: An iq consortium review. J.

Pharm. Sci. 2018, 107, 2267-2282.

7. Paul, S.; Taylor, L. J.; Murphy, B.; Krzyzaniak, J. F.; Dawson, N.; Mullarney, M.

P.; Meenan, P.; Sun, C. C. Powder properties and compaction parameters that influence

punch sticking propensity of pharmaceuticals. Int. J. Pharm. 2017, 521, 374-383.

8. Abdel-Hamid, S.; Betz, G. A novel tool for the prediction of tablet sticking during

high speed compaction. Pharm. Dev. Technol. 2012, 17, 747-754.

9. Roberts, M.; Ford, J. L.; MacLeod, G. S.; Fell, J. T.; Smith, G. W.; Rowe, P. H.

Effects of surface roughness and chrome plating of punch tips on the sticking tendencies

of model ibuprofen formulations. J. Pharm. Pharmacol. 2003, 55, 1223-1228.

10. Paul, S.; Wang, C.; Wang, K.; Sun, C. C. Reduced punch sticking propensity of

acesulfame by salt formation: Role of crystal mechanical property and surface chemistry.

Mol. Pharm. 2019, 16, 2700-2707.

11. Paul, S.; Sun, C. C. Modulating sticking propensity of pharmaceuticals through

excipient selection in a direct compression tablet formulation. Pharm. Res. 2018, 35, 113.

12. Al-Karawi, C.; Leopold, C. S. A comparative study on the sticking tendency of

ibuprofen and ibuprofen sodium dihydrate to differently coated tablet punches. Eur. J.

Pharm. Biopharm. 2018, 128, 107-118.

13. Simmons, D. M.; Gierer, D. S. A material sparing test to predict punch sticking

during formulation development. Drug Dev. Ind. Pharm. 2012, 38, 1054-1060.

14. Kawashima, Y.; Niwa, T. Preparation of controlled-release microspheres of

ibuprofen with acrylic polymers by a novel quasi-emulsion solvent diffusion method. J.

Pharm. Sci. 1988, 78, 68-72.

15. Sano, A.; Kuriki, T.; Kawashima, Y.; Takeuchi, H.; Hino, T.; Niwa, T. Particle

design of tolbutamide by the spherical crystallization technique. Iii. Micromeritic

Page 261: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

239

properties and dissolution rate of tolbutamide spherical agglomerates prepared by the

quasi-emulsion solvent diffusion method. Chem. Pharm. Bull. 1990, 38, 733-739.

16. Kawashima, Y.; Niwa, T.; Takeuchi, H.; Hino, T.; Itoh, Y.; Furuyama, S.

Characterization of polymorphs of tranilast anhydrate and tranilast monohydrate when

crystallized by two solvent change spherical crystallization techniques. J. Pharm. Sci. 1991,

80, 472-478.

17. Morishima, K.; Kawashima, Y.; Kawashima, Y.; Takeuchi, H.; Niwa, T.; Hino, T.

Micromeritic characteristics and agglomeration mechanisms in the spherical crystallization

of bucillamine by the spherical agglomeration and the emulsion solvent diffusion methods.

Powder Technol. 1993, 76, 57-64.

18. Tahara, K.; O’Mahony, M.; Myerson, A. S. Continuous spherical crystallization of

albuterol sulfate with solvent recycle system. Cryst. Growth Des. 2015, 15, 5149-5156.

19. Tahara, K.; Kono, Y.; Myerson, A. S.; Takeuchi, H. Development of continuous

spherical crystallization to prepare fenofibrate agglomerates with impurity complexation

using mixed-suspension, mixed-product removal crystallizer. Cryst. Growth Des. 2018, 18,

6448-6454.

20. Ray, W. A.; Stein, C. M.; Daugherty, J. R.; Hall, K.; Arbogast, P. G.; Griffin, M.

R. Cox-2 selective non-steroidal anti-inflammatory drugs and risk of serious coronary

heart disease. The Lancet 2002, 360, 1071-1073.

21. Wang, K.; Mishra, M. K.; Sun, C. C. Exceptionally elastic single-component

pharmaceutical crystals. Chem. Mater. 2019, 31, 1794-1799.

22. Paul, S.; Chang, S.-Y.; Sun, C. C. The phenomenon of tablet flashing — its impact

on tableting data analysis and a method to eliminate it. Powder Technol. 2017, 305, 117-

124.

23. Fell, J. T.; Newton, J. M. Determination of tablet strength by the diametral-

compression test. J. Pharm. Sci. 1970, 59, 688-691.

24. Ryshkewitch, E. Compression strength of porous sintered alumina and zirconia. J.

Am. Ceram. Soc. 1953, 36, 65-68.

25. Chen, M.-H.; Bergman, C. J. Method for determining the amylose content,

molecular weights, and weight- and molar-based distributions of degree of polymerization

of amylose and fine-structure of amylopectin. Carbohydr. Polym. 2007, 69, 562-578.

26. Gapper, L. W.; Copestake, D. E. J.; Otter, D. E.; Indyk, H. E. J. A.; Chemistry, B.

Analysis of bovine immunoglobulin g in milk, colostrum and dietary supplements: A

review. Anal. Bioanal. Chem. 2007, 389, 93-109.

27. Sun, C. C. Setting the bar for powder flow properties in successful high speed

tableting. Powder Technol. 2010, 201, 106-108.

28. Shi, L.; Sun, C. C. Transforming powder mechanical properties by core/shell

structure: Compressible sand. J. Pharm. Sci. 2010, 99, 4458-4462.

Page 262: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

240

29. Osei-Yeboah, F.; Chang, S.-Y.; Sun, C. C. J. P. R. A critical examination of the

phenomenon of bonding area - bonding strength interplay in powder tableting. Pharm. Res.

2016, 33, 1126-1132.

30. Tanuma, S.; Powell, C. J.; Penn, D. R. Calculations of electorn inelastic mean free

paths. Ii. Data for 27 elements over the 50–2000 ev range. Surf. Interface Anal. 1991, 17,

911-926.

31. Chen, Y.; Pui, Y.; Chen, H.; Wang, S.; Serno, P.; Tonnis, W.; Chen, L.; Qian, F.

Polymer-mediated drug supersaturation controlled by drug–polymer interactions persisting

in an aqueous environment. Mol. Pharm. 2019, 16, 205-213.

32. Jasani, M. S.; Kale, D. P.; Singh, I. P.; Bansal, A. K. Influence of drug–polymer

interactions on dissolution of thermodynamically highly unstable cocrystal. Mol. Pharm.

2019, 16, 151-164.

33. Schneider, H.-J.; Hacket, F.; Rüdiger, V.; Ikeda, H. Nmr studies of cyclodextrins

and cyclodextrin complexes. Chemical Reviews 1998, 98, 1755-1786.

34. Ilevbare, G. A.; Liu, H.; Edgar, K. J.; Taylor, L. S. Impact of polymers on crystal

growth rate of structurally diverse compounds from aqueous solution. Mol. Pharm. 2013,

10, 2381-2393.

35. Xie, T.; Taylor, L. S. Dissolution performance of high drug loading celecoxib

amorphous solid dispersions formulated with polymer combinations. Pharm. res. 2016, 33,

739-750.

Page 263: IMPROVE THE MANUFACTURABILITY AND DISSOLUTION …

241

Chapter 7.

1. Good, D. J.; Rodríguez-Hornedo, N. Solubility advantage of pharmaceutical

cocrystals. Crys. Growth Des. 2009, 9, 2252-2264.

2. Chen, H.; Wang, C.; Kang, H.; Zhi, B.; Haynes, C. L.; Aburub, A.; Sun, C. C.

Microstructures and pharmaceutical properties of ferulic acid agglomerates prepared by

different spherical crystallization methods. Int. J. Pharm. 2020, 574, 118914.


Recommended