Sei sulla pagina 1di 217

A STUDY ON COMPACTION PUNCH DESIGN

ON TABLET PROPERTIES

PARTHIBAN ANBALAGAN
B. Sc. (Pharm.) (Hons.)

A THESIS SUBMITTED FOR THE DEGREE OF


DOCTOR OF PHILOSOPHY

DEPARTMENT OF PHARMACY
NATIONAL UNIVERSITY OF SINGAPORE
2017

Supervisors:
Dr. Celine Valeria Liew, Main Supervisor
Associate Professor Paul Heng Wan Sia, Co-Supervisor

Examiners:
Dr. Ng Wai Kiong, Institute of Chemical & Engineering Sciences
Professor Kunikazu Moribe, Chiba University
Dr. Effendi Widjaja, Johnson & Johnson
DECLARATION

I hereby declare that the thesis is my original work and it has been written by me in

its entirety. I have duly acknowledged all the sources of information which have

been used in the thesis.

The thesis has also not been submitted for any degree in any university previously.

________________________

Parthiban Anbalagan

21 August 2017

ii
ACKNOWLEDGEMENTS

I would like to express my heartfelt gratitude to my supervisors, Dr. Celine Valeria

Liew and Associate Professor Paul Heng Wan Sia, for their guidance, support and

encouragement throughout the candidature. I am also thankful for the teachings and

guidance from Associate Professor Chan Lai Wah and Associate Professor T. R. R.

Kurup.

I would like to thank the Department of Pharmacy, Faculty of Science and the

National University of Singapore for providing me with the research scholarship and

administrative support in these four years. My special appreciation goes to Ms.

Teresa Ang and Ms. Wong Mei Yin for their technical assistance. I would also like

to thank Ms. Vivian Chin Hui Min for her valuable contribution.

I would also like to acknowledge all members in GEA-NUS PPRL, whom I have

had the pleasure of working alongside during the course of my candidature. This

dissertation could not have been completed without your great support.

Finally, I wish to thank my family and friends for their constant support and

encouragement throughout these years.

With my sincere appreciation,

Parthiban Anbalagan

2017

iii
TABLE OF CONTENTS

DECLARATION ................................................................................................... ii
ACKNOWLEDGEMENTS ................................................................................. iii
TABLE OF CONTENTS ......................................................................................iv
SUMMARY ............................................................................................................ix
LIST OF TABLES .................................................................................................xi
LIST OF FIGURES .............................................................................................xiv
LIST OF SYMBOLS AND ABBREVIATIONS ...............................................xix
1. INTRODUCTION............................................................................................... 2
1.1 Pharmaceutical tablet manufacture ................................................................. 2
1.1.1 Tablet compaction process ....................................................................... 2
1.1.2 Particle deformation mechanism .............................................................. 3
1.1.3 Characterization of particle deformation ................................................. 5
1.1.3.1 Heckel analysis ................................................................................. 5
1.1.3.2 Stress relaxation analysis .................................................................. 6
1.1.3.3 Deformation parameters derived from compaction profiles ............. 8
1.1.3.3.1 Compaction force-time profile ................................................... 8
1.1.3.3.2 Compact deformation during high-speed tableting .................. 10
1.1.3.3.3 Compaction force-punch separation profile ............................. 11
1.1.3.4 Compaction parameters................................................................... 13
1.1.4 Instrumentation in tablet compaction ..................................................... 14
1.1.4.1 Rotary tablet press........................................................................... 14
1.1.4.1.1 Turret speed ............................................................................. 14
1.1.4.1.2 Compaction rolls ...................................................................... 15
1.1.4.1.3 Precompaction.......................................................................... 15
1.1.4.1.4 Press technology ...................................................................... 15
1.1.4.2 Compaction simulator ..................................................................... 16
1.1.4.3 Single station tablet press ................................................................ 17
1.2 Tablet physical quality .................................................................................. 17
1.2.1 Tablet tensile strength ............................................................................ 17
1.2.2 Tablet failure tendency .......................................................................... 19
1.2.3 Tablet material adherence tendency ....................................................... 20

iv
1.2.4 Factors influencing tablet physical quality ............................................ 21
1.2.4.1 Deformation characteristics of tableting materials ......................... 21
1.2.4.2 Lubrication...................................................................................... 22
1.2.4.3 Compaction force............................................................................ 23
1.2.4.4 Tableting speed ............................................................................... 24
1.2.4.5 Tablet press and tooling technology ............................................... 25
1.3 Tablet compaction tooling ............................................................................ 26
1.3.1 General concepts .................................................................................... 26
1.3.2 Punch head configuration ...................................................................... 28
1.3.3 Punch face configuration ....................................................................... 30
1.4 Computational simulation of powder compaction ........................................ 34
1.4.1 Finite element modeling (FEM) ............................................................ 34
1.4.2 Discrete element Modeling (DEM) ....................................................... 36
1.5 Research gaps ............................................................................................... 38
1.5.1 Dwell effect on tablet quality ................................................................ 39
1.5.2 Punch head configurational modification .............................................. 40
1.5.3 Punch face configurational modification ............................................... 41
2. HYPOTHESES AND OBJECTIVES ............................................................. 44
3. EXPERIMENTAL............................................................................................ 47
3.1 Part 1: Investigation of the influence of deformation mechanism of tableting
materials during dwell, and the consequent effect on tablet mechanical
properties ............................................................................................................ 47
3.1.1 Rationale of study .................................................................................. 47
3.1.2 Choice of tableting materials ................................................................. 47
3.1.3 Characterization of tableting materials .................................................. 48
3.1.3.1 Heckel analysis ............................................................................... 48
3.1.3.2 Particle size analysis ....................................................................... 49
3.1.3.3 Hausner ratio................................................................................... 50
3.1.3.4 Angle of repose ............................................................................... 50
3.1.4 Preparation of tableting blends .............................................................. 51
3.1.5 Tablet production ................................................................................... 51
3.1.6 Stress relaxation test .............................................................................. 51
3.1.7 Tablet characterization........................................................................... 52
3.1.7.1 Viscoelastic strain recovery ............................................................ 52
3.1.7.2 Tablet tensile strength ..................................................................... 53

v
3.1.8 Statistical analysis .................................................................................. 53
3.2 Part 2. Investigation of the effect of punch head modification on compaction
parameters and tablet physical quality ................................................................ 54
3.2.1 Rationale of study .................................................................................. 54
3.2.2 Choice of tableting materials ................................................................. 54
3.2.3 Preparation of tableting blends .............................................................. 54
3.2.4 Characterization of tableting material .................................................... 55
3.2.4.1 Particle size analysis ....................................................................... 55
3.2.4.2 Hausner ratio ................................................................................... 55
3.2.4.3 Angle of repose ............................................................................... 55
3.2.4.4 Heckel analysis ............................................................................... 56
3.2.4.5 Stress relaxation test........................................................................ 56
3.2.5 Tablet production ................................................................................... 56
3.2.5.1 Tableting punch head configuration ............................................... 56
3.2.5.2 Rotary tableting ............................................................................... 58
3.2.6 Compaction parameters.......................................................................... 60
3.2.7 Evaluation of compact deformation ....................................................... 61
3.2.8 Tablet characterization ........................................................................... 62
3.2.8.1 Tablet elastic recovery .................................................................... 62
3.2.8.2 Tablet porosity ................................................................................ 62
3.2.8.3 Tablet tensile strength ..................................................................... 62
3.2.8.4 Capping index ................................................................................. 63
3.2.9 Computational simulation of tablet compaction .................................... 63
3.2.10 Statistical analysis ................................................................................ 66
3.3 Part 3. Investigation of the effect of punch face modification on compaction
parameters and tablet physical quality ................................................................ 67
3.3.1 Rationale of study .................................................................................. 67
3.3.2 Preparation of tableting blends .............................................................. 67
3.3.3 Tablet production ................................................................................... 67
3.3.3.1 Tableting punch face configurations ............................................... 67
3.3.3.2 Rotary tablet production.................................................................. 69
3.3.4 Compaction energy ................................................................................ 71
3.3.5 Tablet characterization ........................................................................... 72
3.3.5.1 Tablet elastic recovery .................................................................... 72
3.3.5.2 Tablet porosity ................................................................................ 72

vi
3.3.5.3 Tablet tensile strength ..................................................................... 72
3.3.5.4 Capping index ................................................................................. 72
3.3.6 Computational simulation of tablet compaction .................................... 72
3.3.6.1 Compact stress analysis .................................................................. 73
3.3.7 Statistical analysis .................................................................................. 74
3.4 Part 4. Investigation of the effect of punch face modification on powder
adhesion tendency to the punch surface during powder compaction ................. 75
3.4.1 Rationale of study .................................................................................. 75
3.4.2 Choice of tableting materials ................................................................. 75
3.4.3 Preparation of tableting blends .............................................................. 75
3.4.4 Characterization of tableting blends ...................................................... 76
3.4.4.1 Particle size analysis ....................................................................... 76
3.4.4.2 Angle of repose measurement......................................................... 76
3.4.5 Tablet production ................................................................................... 76
3.4.5.1 Punch face configuration ................................................................ 76
3.4.5.2 Tablet production using manual press ............................................ 76
3.4.5.3 Tablet production using compaction simulator .............................. 78
3.4.6 Assessment of powder adherence .......................................................... 79
3.4.6.1 Quantification of powder adherence ............................................... 79
3.4.6.2 Force parameters related to powder adherence............................... 80
3.4.7 Statistical analysis .................................................................................. 81
4. RESULTS AND DISCUSSION ....................................................................... 83
4.1 Part 1: Investigation of the influence of deformation mechanism of tableting
materials during dwell and the consequent effect on tablet mechanical properties
............................................................................................................................ 83
4.1.1 Physical characteristics of tableting materials ....................................... 83
4.1.2 Compression behavior of tableting materials ........................................ 83
4.1.3 Compact deformation during dwell ....................................................... 85
4.1.4 Effect of formulation variables on tablet properties .............................. 92
4.1.5 Effect of dwell on tablet properties...................................................... 100
4.1.6 Summary .............................................................................................. 102
4.2 Part 2. Investigation of the effect of punch head modification on compaction
parameters and tablet physical quality .............................................................. 104
4.2.1 Physical characteristics of tableting materials ..................................... 104
4.2.2 Compression behavior of tableting materials ...................................... 104
4.2.3 Influence of punch head configuration on compaction profile ............ 106

vii
4.2.4. Influence of compaction roll configuration on compaction profile .... 110
4.2.5 Influence of punch head configuration on tablet properties................. 114
4.2.6 Finite element analysis ......................................................................... 122
4.2.7 Influence of compaction roll configuration on tablet properties .......... 126
4.2.8 Effect of tableting speed ...................................................................... 132
4.2.9 Summary .............................................................................................. 134
4.3 Part 3. Investigation of the effect of punch face modification on compaction
parameters and tablet physical quality .............................................................. 136
4.3.1 Influence of punch face edge geometry on compaction parameters .... 136
4.3.2 Influence of punch face edge geometry on tablet properties ............... 139
4.3.3 Discrete Element analysis .................................................................... 144
4.3.3.1 Comparison between simulations and experiments ...................... 144
4.3.3.2 Effect of punch face edge geometry on tablet stress distribution . 147
4.3.4 Effect of tableting speed ...................................................................... 151
4.3.5 Effect of precompaction ....................................................................... 153
4.3.6 Effect of roll displacement ................................................................... 156
4.3.7 Summary .............................................................................................. 159
4.4 Part 4. Investigation of the effect of punch face modification on powder
adhesion tendency to the punch surface during powder compaction ................ 162
4.4.1 Physical characteristics of tableting material ....................................... 162
4.4.2 Powder adhesion model ....................................................................... 162
4.4.3 Influence of punch face configuration ................................................. 163
4.4.4 Influence of compaction force on powder adhesion ............................ 167
4.4.5 Influence of magnesium stearate concentration ................................... 169
4.4.6 Influence of compaction method (compaction simulator..................... 171
4.4.7 Influence of punch face configuration (compaction simulator) ........... 172
4.4.8 Influence of compaction force (compaction simulator) ....................... 174
4.4.9 Ejection force as sticking parameter .................................................... 175
4.4.10 Summary ............................................................................................ 177
5. CONCLUSION ............................................................................................... 180
6. REFERENCES................................................................................................ 185
7. LIST OF PUBLICATIONS AND PRESENTATIONS ............................... 195

viii
SUMMARY

Tablet compaction punches have undergone many changes to their design

specifications over the years in an effort to improve manufacturing efficiency, tablet

quality and durability of the tools. This study was initiated with the objective of

understanding the impact of developing new punch designs by modification of

punch head and face configurations, with regard to tablet physical quality.

In the first part of the work, the influence of different deformation mechanisms of

tableting materials during the dwell phase and the consequent effect on tablet

mechanical properties were investigated. Compact deformation during dwell was

observed to adopt distinct phases which were dependent on the degree of plasticity

and porosity in the compact system. Dwell time extension was found to be beneficial

when preparing tablets containing components of high plasticity by forming

mechanically strong tablets. However, the effect of dwell could possibly be

influenced by time-dependent secondary factors.

The specific impact of punch head configuration on the compaction process and

tablet mechanical properties was subsequently studied in a rotary press. Punches

with modified head flat and head radius specifications were used for high speed

tableting. Extension of head flat diameter increased dwell time and this led to

significant improvements in physical quality of tablets prepared. Dwell time

prolongation enabled greater plastic flow to occur, consequently reduced tablet

elasticity and formed more structural bonds. For punches without a head flat, a

change in the head radius was found to influence the compaction process marginally

and it led to insignificant differences in tablet properties.

The influence of punch face modification on the tableting process was next studied

and tableting performance under different compaction conditions were determined.

Flat-face punches with different face edge geometries (bevel or radius edge

ix
geometry) were used for tablet production in a rotary press. Modification of the

straight bevel edge to a curved edge enabled deeper punch penetration in the die

cavity during the compaction cycle and facilitated greater compact densification.

The presence of a bevel edge led to greater stress anisotropy in the tablet

microstructure and contributed to development of higher elastic energy within the

compact. Application of a precompaction cycle along with dwell time extension to

the tableting process amplified the tableting performance of radius edge punch face

configuration to a greater extent than the bevel edged punch.

In the final part of the study, the influence of punch face configurational

modification on tablet adhesion tendency was studied. The flat-face punches with

radius or bevel edge face configurations and the plain flat-face punches were used

for tablet compaction. Punches possessing face configurations with a concavity

(radius and bevel edge faces) were shown to be advantageous over the flat-face plain

face configuration as the presence of an edge cup reduced residual powder adherence

significantly. Powder adhesion data between the edged punch face configurations

showed similar degrees of material sticking, with only marginal differences,

suggesting equivalent tableting performance in terms of tablet adhesion tendency.

This research study on punch tools provided a comprehensive understanding on the

advantages and limitations of punch head and face configurational changes with

respect to tablet properties. The findings may aid tablet press and tools

manufacturers in process and tool design considerations for a successful tableting

program.

x
LIST OF TABLES

Table 1. Specifications of FFBE and FFRE punch face configurations (FFBE:


flat-face bevel edge, FFRE: flat-face radius edge) ............................... 68

Table 2. Physical characteristics of microcrystalline cellulose (MCC), dicalcium


phosphate (DCP) and paracetamol (APAP) particles ........................... 83

Table 3. Viscoelastic parameters obtained for APAP/MCC compacts of different


proportions (APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75,
F5: 0/100) ............................................................................................. 87

Table 4. Viscoelastic parameters obtained for DCP/MCC compacts of different


proportions (DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75) . 88

Table 5. Viscoelastic parameters obtained for APAP/DCP compacts of different


proportions (APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75) ........... 91

Table 6. Effect of dwell time on stress decay and mechanical properties of


APAP/MCC tablets of different proportions, compacted at 6 kN
(APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100) 93

Table 7. Effect of dwell time on stress decay and mechanical properties of


APAP/MCC tablets of different proportions, compacted at 9 kN
(APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100) 94

Table 8. Effect of dwell time on stress decay and mechanical properties of


DCP/MCC tablets of different proportions, compacted at 6 kN
(DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75) .................... 95

Table 9. Effect of dwell time on stress decay and mechanical properties of


DCP/MCC tablets of different proportions, compacted at 9 kN
(DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75) .................... 96

Table 10. Effect of dwell time on stress decay and mechanical properties of
APAP/DCP tablets of different proportions, compacted at 6 kN
(APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75) .............................. 98

Table 11. Effect of dwell time on stress decay and mechanical properties of
APAP/DCP tablets of different proportions, compacted at 9 kN
(APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75) .............................. 99

xi
Table 12. Physical characteristics of paracetamol-starch and lactose granules .. 104

Table 13. Effect of punch head configuration on compaction parameters obtained


using the larger compaction roll (D240) ............................................. 107

Table 14. Effect of punch head configuration on compaction parameters obtained


using the smaller compaction roll (D150)........................................... 111

Table 15. Effect of punch head configuration on lactose tablet mechanical


properties at the experimental compaction forces for compaction roll of
diameter 240 mm (EHF: extended head flat, SHF: standard head flat,
RHF: reduced head flat, SRH: standard radius head, RRH: reduced radius
head).................................................................................................... 118

Table 16. Effect of punch head configuration on paracetamol-starch tablet


mechanical properties at the experimental compaction forces for
compaction roll of diameter 240 mm (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 121

Table 17. Effect of punch head configuration on lactose tablet mechanical


properties at the experimental compaction forces for compaction roll of
diameter 150 mm (EHF: extended head flat, SHF: standard head flat,
RHF: reduced head flat, SRH: standard radius head, RRH: reduced radius
head).................................................................................................... 128

Table 18. Effect of punch head configuration on paracetamol-starch tablet


mechanical properties at the experimental compaction forces for
compaction roll of diameter 150 mm (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 131

Table 19. Effect of punch head configuration on compaction parameters of lactose


tablets compacted at turret speed 35 rpm (47 m/min) (EHF: extended
head flat, SHF: standard head flat, RHF: reduced head flat) .............. 133

Table 20. Effect of punch head configuration on mechanical properties of lactose


tablets compacted at turret speed 35 rpm (47 m/min) (EHF: extended
head flat, SHF: standard head flat, RHF: reduced head flat) .............. 133

Table 21. Effect of punch face edge geometry on set compaction thickness at the
main compaction event ....................................................................... 137

xii
Table 22. Properties of lactose tablets prepared using single compaction cycle by
FFBE and FFRE punch face configurations ....................................... 140

Table 23. Properties of paracetamol-starch tablets prepared using single


compaction cycle by FFBE and FFRE punch face configurations ..... 143

Table 24. Properties of lactose tablets prepared using single compaction cycle by
FFBE and FFRE punch face configurations at turret speed 35 rpm (47
m/min) ................................................................................................ 152

Table 25. Properties of paracetamol-starch tablets prepared using double


compaction (mode1) cycle by FFBE and FFRE punch face configurations
............................................................................................................ 154

Table 26. Properties of paracetamol-starch tablets prepared using double


compaction (mode2) cycle by FFBE and FFRE punch face configurations
............................................................................................................ 158

xiii
LIST OF FIGURES

Figure 1. Stress relaxation curve showing stress decay in a compact under constant
strain........................................................................................................ 7

Figure 2. Schematic diagrams depicting the motion of punch head in relation to


compaction roll (A) and the graphical representation of compaction
force-time profile with the associated compaction parameters (B). ....... 9

Figure 3. Graphical representation of compaction force-punch separation profile


.............................................................................................................. 12

Figure 4. Model for vertical displacement of a punch head on a rotary press...... 12

Figure 5. Schematic showing the diametrical compression test. .......................... 18

Figure 6. Image showing absence of capping and tablet capping (left to right) after
hardness testing. .................................................................................... 19

Figure 7. Schematics showing the lower (left) and upper (right) punches. .......... 27

Figure 8. Schematic showing the punch head configuration with the associated
head flat and head radius....................................................................... 28

Figure 9. Schematics showing the FFP (A) FFBE (B) and FFRE (C) tablet designs
with difference in tablet edges shown. (FFP: flat-face plain, FFBE: flat-
face bevel edge, FFRE: flat-face radius edge) ...................................... 32

Figure 10. Schematics showing the dimensions (in mm) of the EHF (A), SHF (B),
RHF (C), SRH (D) and RRH (E) punch head configurations in terms of
their head flat (HF), head radius (HR) and head thickness (HT)
specifications. (EHF: extended head flat, SHF: standard head flat, RHF:
reduced head flat, SRH: standard radius head, RRH: reduced radius head)
.............................................................................................................. 57

Figure 11. Schematic showing the rotary press turret with positions of the die filling,
precompaction, main compaction and ejection stages. ......................... 59

Figure 12. Schematic diagrams depicting the (i) motion of the punch at
precompaction event and (ii) motion of the punch at main compaction

xiv
event (A). Graphical representation of the force-time curve at the main
compaction event, with precompaction deactivated (B). Graphical
representation of force-time curve at precompaction event with main
compaction deactivated (C). ................................................................. 59

Figure 13. Drucker-Prager Cap model with associated parameters.


.............................................................................................................. 64

Figure 14. Schematic showing the uniaxial compression test for material property
calibration. ............................................................................................ 64

Figure 15. Schematics showing the dimensions (in mm) of the FFBE (A) and FFRE
(B) punch face configurations. (FFBE: flat-face bevel edge, FFRE: flat-
face radius edge) ................................................................................... 68

Figure 16. Schematic diagrams depicting the motion of the punch at precompaction
(PC) event (i) with air compensator deactivated (mode1) and (ii) motion
of the punch at main compaction (MC) event (A). Graphical
representation of force-time curve at precompaction event with air
compensator deactivated (mode1) and the force time-time curve at the
main compaction event (B)................................................................... 70

Figure 17. Schematic diagrams depicting the motion of the punch at precompaction
event (i) with air compensator activated (mode2) and (ii) motion of the
punch at main compaction event (A). Graphical representation of force-
time curve at precompaction event with air compensator activated
(mode2) with the associated roll displacement curve and the force-time
curve at the main compaction event (B). .............................................. 70

Figure 18. Manual press used for tablet production.


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

Figure 19. Tooling setup (A) and the force feeder with the take-off force sensor (B)
in the compaction simulator.................................................................. 78

Figure 20. Schematics showing the stress factors associated with tablet ejection (A)
and take-off (B). (Fe: ejection force, Ft: take-off force, Fa: adhesive force,
Fr: compact-die wall force) ................................................................... 81

Figure 21. Heckel plots for microcrystalline cellulose (♦), dicalcium phosphate (■)
and paracetamol (●) particles ............................................................... 84

xv
Figure 22. Stress relaxation profiles of APAP/MCC compacts of different
proportions. (APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75,
F5: 0/100).............................................................................................. 87

Figure 23. Stress relaxation profiles of DCP/MCC compacts of different


proportions. (DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75) 88

Figure 24. Stress relaxation profiles of APAP/DCP compacts of different


proportions. (APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75) .......... 91

Figure 25. Heckel plots (A) and stress relaxation curves (B) for lactose (  ) and
paracetamol-starch (  ) granules. ...................................................... 105

Figure 26. Comparison of compaction profiles (6 and 9 kN) using the larger
compaction roll (D240) between SHF and RHF (A), SRH and RRH (B)
head configurations. (SHF: standard head flat, RHF: reduced head flat,
SRH: standard radius head, RRH: reduced radius head) .................... 108

Figure 27. Comparison of compaction profiles (6 and 9 kN) between compaction


rolls of diameters 240 (─) and 150 (---) mm using the SHF (A) and EHF
(B) head configurations. (EHF: extended head flat, SHF: standard head
flat) ...................................................................................................... 112

Figure 28. Effect of punch head configuration on lactose tablet tensile strength (A)
and area quotient (B) at the experimental compaction forces for
compaction roll of diameter 240 mm. (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 115

Figure 29. Effect of punch head configuration on paracetamol-starch tablet tensile


strength (A) and area quotient (B) at the experimental compaction forces
for compaction roll of diameter 240 mm. (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 120

Figure 30. Stress distribution contours of simulated paracetamol-starch tablets


compacted at dwell times representative of RRH (A), SRH (B), RHF (C),
SHF (D) and EHF (E) punch head configurations. ............................. 124

Figure 31. Effect of punch head configuration on lactose tablet tensile strength (A)
and area quotient (B) at the experimental compaction forces for

xvi
compaction roll of diameter 150 mm. (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 127

Figure 32. Effect of punch head configuration on paracetamol-starch tablet tensile


strength (A) and area quotient (B) at the experimental compaction forces
for compaction roll of diameter 150 mm. (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head,
RRH: reduced radius head) ................................................................. 130

Figure 33. Tensile strength of lactose tablets prepared using single compaction
cycle by FFBE (■) and FFRE (□) punch face configurations. ........... 140

Figure 34. Compaction force-punch separation curves for the FFBE (●) and FFRE
(○) punch face configurations. ............................................................ 141

Figure 35. Tensile strength of paracetamol-starch tablets prepared using single


compaction cycle by FFBE (■) and FFRE (□) punch face configurations.
............................................................................................................ 143

Figure 36. Comparison between simulation and experimental tablet thickness (A)
and RFA values (B) for FFP (●), FFRE (♦) and FFBE (▲) tablets. .. 145

Figure 37. (i) Compaction of particles with grid bin group in the vertical Y-Z cross-
section obtained for FFP (A), FFRE (B) and FFBE (C) tablets. Fig. 37
(ii-iii) Stress distribution in the vertical Y-Z cross-sections obtained for
FFP (A), FFRE (B) and FFBE (C) tablets at compaction force (ii) 8.0 and
(iii) 9.0 kN. The X and Y axes of the contour plots represent the distance
(mm) from the edge and bottom of the tablet respectively. ................ 148

Figure 38. Roll displacement-time curves for the FFBE (—) and FFRE (---) punch
face configurations with the activation of the air compensator (mode2)
during double compaction cycle. ........................................................ 157

Figure 39. FFP punch face showing first-degree (A) and second-degree (B) powder
adhesion. ............................................................................................. 163

Figure 40. Effect of punch face configuration on powder adherence on the upper (■)
and lower (□) punch faces after compacting 30 (A) and 50 (B) tablets
using an ibuprofen-lactose formulation containing 1 %, w/w magnesium
stearate. ............................................................................................... 164

xvii
Figure 41. Distribution of residual powder on flat-face bevel (A) and radius (B)
edge punch faces. ................................................................................ 167

Figure 42. Effect of punch face configuration on powder adherence on the upper (■)
and lower (□) punch faces after compacting 30 (A) and 50 (B) tablets
using an ibuprofen-lactose formulation containing 1.5 %, w/w
magnesium stearate. ............................................................................ 170

Figure 43. Effect of punch face configuration on powder adherence on the upper
punch face after tablet production in the compaction simulator, using an
ibuprofen-lactose formulation containing 1 %, w/w magnesium stearate.
............................................................................................................ 171

Figure 44. Take-off force profile during the time course of tablet production using
the compaction simulator for the FFP (○), FFRE (●) and FFBE (▲)
punch face geometries at compaction force 5.0 kN (A), 7.5 kN (B) and
10 kN (C) using an ibuprofen-lactose formulation containing 1 %, w/w
magnesium stearate. ............................................................................ 173

Figure 45. Ejection force profile during the time course of tablet production using
the compaction simulator for the FFP (○), FFRE (●) and FFBE (▲)
punch face geometries at compaction force 5.0 kN (A), 7.5 kN (B) and
10 kN (C) using an ibuprofen-lactose formulation containing 1 %, w/w
magnesium stearate. ............................................................................ 176

xviii
LIST OF SYMBOLS AND ABBREVIATIONS

AOR Angle of repose

APAP Paracetamol

API Active pharmaceutical ingredient

AQ Area quotient

a Volumetric plastic strain

CSmax Force at the beginning of the dwell phase

CS0 Force at the end of the dwell phase

Cp Compactibility coefficient

DCP Dicalcium phosphate

DEM Discrete element modeling

D150 Compaction roll of diameter 150 mm

D240 Compaction roll of diameter 240 mm

dt Tablet diameter

D Cohesion parameter

dx Particle size corresponding to Xth volume percentile


under the cumulative undersize distribution curve

EHF Extended head flat

xix
ER Tablet elastic recovery

EU European standards

F Formulation

F Tablet fracture force

FEM Finite element modeling

FFP Flat-face plain

FFBE Flat-face bevel edge

FFRE Flat-face radius edge

Ht Tablet thickness

HF Head flat

HT Head thickness

HR Head radius

MCC Microcrystalline cellulose

Np Number of tablets that capped off the press

Nh Number of tablets that capped during hardness testing

Nt Total number of tablets

Pb Cap surface parameter

Ptapped Tapped density

xx
Pbulk Bulk density

R Material parameter

RFA Rate of force application

RHF Reduced head flat

RRH Reduced radius head

rpm Revolutions per minute

RQ Relaxation quotient

SHF Standard head flat

SRH Standard radius head

STmain Set main compaction thickness

STpre Set precompaction thickness

Tcon Consolidation time

Tdwell Dwell time

Tdec Decompression time

Ttotal Total compaction time

VER Tablet viscoelastic recovery

VS Viscoelastic slope

Wnet Net work of compaction

xxi
Wexp Work of expansion

σz Axial stresses

σr Radial stresses

f
σd Radial tensile strength

f
σc Axial tensile strength

β Friction angle

ɛ Compact porosity

xxii
CHAPTER 1

INTRODUCTION

1
1. INTRODUCTION

1.1 Pharmaceutical tablet manufacture

The tablet dosage form is the most common form of solid medicament that have

been used in medicine for many years. Tablet manufacturing process has evolved

over time, with the first recorded hand-operated tablet press patented by William

Brockedon in 1843 [1]. Tablet production rate greatly increased with the invention

of the automated tablet press by John Wyeth and Bros in the late 1870s [2]. Since

then, tablet manufacturing process has evolved with advances in press and tooling

technology.

1.1.1 Tablet compaction process

A tablet is formed by reducing the volume of powder particles confined in a die into

a consolidated solid body with a pair of punches. Powder compaction process

comprises a series of sequential phases - compression, consolidation and

decompression - in a particulate solid-gas system as a result of an applied force [3,

4]. Compression involves a reduction in bulk volume caused by particle

rearrangement at relatively low pressures. Compact porosity reduces when particles

occupy the voids between them thus forming a closely packed structure. At a certain

point, the interparticulate friction and particulate packing characteristics, e.g. shape

and surface roughness, will restrict any further particle rearrangement. Application

of pressure beyond this stage results in deformation of the particles, new particle

surfaces more interparticulate contact points are created [5]. Compression processes

next lead to the consolidation phase during which bonds are formed at the contact

points. The consolidation phase represents the stage when an increase in mechanical

strength of a compact is obtained when the compaction pressure is further increased.

It is generally accepted that interparticulate forces, such as van der Waals forces and

2
hydrogen bonding, are the major bonding mechanisms [6]. Other types of bonding

mechanisms include mechanical interlocking and formation of solid bridges between

particles. Decompression occurs when the punches start to move away from each

other after reaching minimum distance apart. As the displacement between the

punches increases, the compaction pressure reduces quickly. During this phase,

some of the elastic strain present under pressure will recover with the compact still

within the confines of the die. This event consequently results in partial disruption

of the interparticulate bonds formed during the consolidation phase and brings about

a reduction in the relative density of the compact. The tablet compaction process is

therefore greatly influenced by the deformation mechanisms of the materials to be

compacted.

1.1.2 Particle deformation mechanism

Application of an external force to the powder bed results in the force being

transmitted through interparticulate points of contact leading to local deformation of

particles. Three main types of particle deformation mechanisms - elastic, plastic and

brittle fracture deformation - have been identified by researchers. Plastic

deformation and fragmentation are irreversible processes that result in permanent

change in the particle shape/size upon removal of the compaction load, whereas

elastic deformation is a reversible process with deformed particles recovering their

original shape after removal of compaction load. As a consequence of these

processes, new particle surfaces are created and interparticulate contact points

increase, leading to the formation of interparticulate bonds. Plastic deformation and

fragmentation of particles are the two processes that control the evolution of the

tablet microstructure in terms of interparticulate bonding. Examples of materials that

consolidate predominantly by plastic deformation include microcrystalline cellulose

3
and sodium chloride, and materials that undergo brittle fracture include dicalcium

phosphate and some classes of crystalline lactose [7-11]. Nonetheless, all materials

experience some degree of elastic, plastic and fragmenting characteristics. However,

usually one type of volume reduction mechanism will predominate for

pharmaceutical materials, dependent on factors such as compaction force and rate.

Some materials undergo deformation processes that are time dependent and occur at

various rates during the compaction sequence, considering that the compact mass in

the die is never in a stress/strain equilibrium during the actual compaction cycle. The

extent of compact formation for such materials is determined by its time dependent

viscoelastic behavior. Increasing the speed of the compaction process may have a

marked effect on powder compactibility, causing mechanical failure of the resultant

tablets for a formulation that produces acceptable tablets at a slower machine speed.

The rate at which the compaction load is applied and removed, and the duration

during which the compact strain is sustained have been shown to be critical factors

when viscoelastic materials are involved. During tableting, compacts usually

undergo a period of static strain known as dwell phase. Particles under compressive

forces during the dwell phase would undergo further plastic deformation and reduce

the interstitial spaces within the compact by better packing, causing a reduction in

porosity and an increase in interparticulate bonding [12]. These realignment

processes allow the internal structure of the tablet to concurrently undergo stress

relief, as characterized by a reduction in the applied force due to time dependent

plastic deformation of the materials under compression. The extent of compact

relaxation has been reported to be largely influenced by compact porosity and

amount of stored energy in the compact [13, 14].

4
1.1.3 Characterization of particle deformation

The study of powder compaction mechanism is an important part in the development

of tablets. Characterization of compaction mechanism is greatly facilitated by the

availability of equipment such as instrumented tablet presses and universal testing

machines that are fitted with force sensors and also, preferably, displacement

measurement probes.

1.1.3.1 Heckel analysis

Heckel analysis is a popular method used in pharmaceutical research to determine

the basic volume reduction mechanism during compaction at high compaction

pressures. It assumes that powder compaction adopts first-order kinetics with the

interparticulate spaces as the reactants and powder densification as the product [15-

17]. The Heckel equation (Eq. (1)) is expressed as:

1
ln = 𝑘P + A (1)
ɛ

where k and A are constants, and porosity (ɛ) of the compacts can be calculated from

the relative density of the compact at applied pressure (P) and the true density of the

material. The reciprocal of k gives the yield pressure, a material dependent constant.

The yield pressure is inversely related to the material resistance to plastic flow; high

values indicate onset of plastic deformation at higher compaction pressures. The

main utility of Heckel plots arises from their ability to identify the predominant

deformation behavior of the materials, particularly to distinguish fragmenting and

plastically deforming materials. Heckel plots allow the classification of powder

compaction behavior based on the progression of the Heckel curves [18, 19]. A

linear relationship with increasing applied pressure indicates deformation

mechanism dominated by plastic deformation [20]. An initial curved region

5
followed by a straight line indicates particles fragmentation at the early stages of the

compaction process with brittle fracture preceding plastic flow [11]. The use of

Heckel analysis for deformation characterization is debatable, attributed to the high

variability in the results from different studies largely due to employment of different

experimental conditions [21]. Variables that affect Heckel plots include compaction

kinematics (rate and duration of compaction), degree of lubrication and

configuration of tools used for compact formation. Even the method of compact

density measurement (in-die or out-of-die) has been reported to significantly impact

the Heckel plots, particularly when elastic materials are compacted. These variables

should be taken into consideration during Heckel analysis. Nonetheless, the widely

used Heckel analysis remains a valuable tool in the field of pharmaceutical

compaction studies.

1.1.3.2 Stress relaxation analysis

While the Heckel analysis evaluates compact deformation at high applied pressures,

it is not suitable for studying compact deformation during dwell phase. Stress

relaxation during the dwell phase has been used by several researchers to evaluate

the viscoelastic behavior of tableting material [22]. In a stress relaxation test,

compacts are formed at a specified force level, and the position of the punches at

maximal stated force is paused for a stipulated period of time. The reduction in the

applied force with time, which occurs due to plastic deformation, is monitored [23].

A typical stress relaxation curve is shown in Fig. 1. The amount of plastic flow

during dwell can be quantified by determining the total amount of compaction force

decay or by evaluating the viscoelastic slope from the linear portions of the stress

relaxation curves [24-26].

6
The stress relaxation test has been adopted by several authors to analyze and

differentiate particle deformation under constant strain. Augsburger et al.

demonstrated the extent of plastic flow during the dwell phase in some commonly

used directly compressible fillers and related the viscoelastic behavior to a Maxwell

model [24]. While compaction force was shown to decay with time for all fillers, a

greater total amount of compaction force decay and a more accelerated rate of decay

was found for microcrystalline cellulose and compressible starch than for lactose

and dicalcium phosphate. This was attributed to the predominantly plastically

deforming nature of the former excipients. Similar results were reported in a study

by Rees and co-workers [27]. It was reported that brittle fracturing materials showed

minimal stress relaxation during the period of constant strain, and a large total stress

relaxation was observed for materials such as pregelatinized starch and cellulose

which are relatively more ductile in nature. The contribution of lubrication to the

viscoelastic behavior of tableting materials on compact deformation during dwell

phase was further investigated by Ebba et al [28]. It was shown that the type of

lubricant used influenced the extent of stress relaxation during dwell phase, with

granules lubricated with magnesium stearate exhibiting the highest stress relaxation

followed by talc and glyceryl palmitostearate.


Compaction force

Time

Figure 1. Stress relaxation curve showing stress decay in a compact under constant
strain.

7
1.1.3.3 Deformation parameters derived from compaction profiles

While the Heckel and stress relaxation curves allow the overall deformation and

viscoelastic character of various tableting materials to be quantified, it may not

reflect the behavior of the compacts during high speed tableting, in which the extent

of time-dependent deformation processes are often different. Compaction conditions

during Heckel and stress relaxation tests are often not representative of the rapid

compaction kinematics encountered during high speed tableting. Hence the

extrapolation of data on particulate deformation from reported studies obtained

under various compaction conditions may not be reliable. Deformation parameters

derived from compaction profiles can be alternatively used to characterize compact

deformation during high speed tableting. Two main types of compaction profiles

(force-time and force-punch separation) have been used extensively in high speed

powder compaction studies.

1.1.3.3.1 Compaction force-time profile

Compaction force-time profiles can possibly be utilized to study dwell phase

deformation during high speed tableting. Fig. 2 shows the movement of the tableting

punch head and the associated compaction force-time profile with the associated

compaction parameters - consolidation (Tcon), dwell (Tdwell) and decompression (Tdec)

times. Due to the importance of the dwell period in relation to viscoelastic

deformation within the compact, it is necessary that the concept of dwell phase is

understood in greater detail. Dwell time is defined in terms of tableting punch

displacement or compaction force profile. Tableting punches experience vertical and

linear motion upon contact with a compaction roll in a rotary press, with maximum

penetration achieved in the die at peak displacement as shown in Fig. 2A, the second

position of the punch head.

8
A

Compaction roll

Punch head

B
7000

6000
Compaction force (N)

A1 A2
5000

4000

3000

2000

1000
Time (ms)
0
0 Tcon
20 40 Tdwell 60 Tdec 80 100

Figure 2. Schematic diagrams depicting the motion of punch head in relation to


compaction roll (A) and the graphical representation of compaction force-time
profile with the associated compaction parameters (B).

Dwell time is often defined as the duration taken for the compaction roll to transit

across the head flat, as it is related to constant vertical punch displacement. This

definition might not be applicable to tableting instrumentation setups in which the

rolls are absent, e.g. compaction simulators. In such scenarios, the period under

which the separation distance between the upper and lower punches remains

constant can be identified as dwell time.

Some authors have defined dwell time as the interval during which maximum

compaction pressure is maintained by the punches during the compaction cycle [29].

However, the geometry of the majority of high speed rotary machines yield rounded

9
pressure peaks, where the true time interval at constant peak pressure may not be

identifiable despite the absence of vertical movement of the punches during the

dwell period as shown in Fig. 2B. On such machines, it is more practical to define

dwell time in terms of compaction force-time profile width at a given fixed fraction

of the peak height. Time interval in a compaction force-time profile during which

compaction force is more or equal to 90% of its peak value, is a commonly

recognized definition of dwell time among practitioners [30].

1.1.3.3.2 Compact deformation during high-speed tableting

Dwell phase compact deformation during high speed tableting can be evaluated from

specific parameters derived from compaction force-time profiles. Dwivedi and

coworkers proposed peak offset time as a deformation parameter to quantify the

extent of compact deformation during dwell phase [31]. The peak offset time is the

difference between the time of peak compaction force and the middle of the dwell

time. Peak offset time was proven as a valid measure of plastic flow during dwell;

punch stress decreased during the peak offset time as the strain remained essentially

constant. Increasing peak compaction force, tableting speed and the presence of

fragmenting materials affected the peak offset time. In a similar attempt to quantify

dwell phase deformation, Schmidt and Vogel proposed an alternative method based

on the degree of asymmetry of the compaction force-time profiles (Fig. 2B) [32].

Dwell period was identified from the compaction profile and divided into halves.

The areas under the compaction force-time curve were calculated for each half, with

both areas decreased by the rectangles under the minimum compaction force in the

dwell time. The ratio of the spatial areas (A2/A1) provided an informative parameter

that quantified deformation behavior and was demonstrated to be sensitive to

compaction load and velocities [33]. The quotient enabled differentiation of the

10
deformation characteristics of different tablet excipients; plastically deforming

materials occupying lower levels of the quotient values and materials that undergo

predominantly brittle and elastic deformations exhibiting higher quotient values [34,

35].

Both methods of plastic flow quantification have been adopted in compaction

studies. Leitriz et al. studied the impact of moisture content on deformation

mechanism of starch during the dwell phase and showed that with increasing

moisture content, the plastic deformation of excipient is amplified so that an

increasing part of the plastic deformation occurs during the consolidation phase, and

a decreasing extent of plastic flow takes place during dwell phase [36]. The

deformation quotient was also demonstrated to be more robust and sensitive to

deformational changes in the compact than peak offset time which exhibited larger

error bars.

1.1.3.3.3 Compaction force-punch separation profile

Compaction force-punch separation profiles allow the calculation of the net energy

input into the compact system which is a function of the material deformation

property [37, 38]. Fig. 3 schematically illustrates a typical compaction force-punch

separation plot. Study of the work of compaction provides information on

elastic/plastic deformation, extent of bond formation and bond disruption during

high speed tableting. Punch separation is the distance between the two punch faces

during the compaction cycle, which can be approximated based on the spatial

relation between tableting punch head, compaction roll and turret configuration as

shown in Fig. 4. Radii of the compaction roll and punch head curvature are

represented by r1 and r2 respectively, wt is the angular velocity of the turret with a

radius of r3, and x2 is the radius of the punch head flat.

11
4500
4000 C
Compaction force

3500
3000
2500
2000
1500
1000
500
0
B D Punch separation A
0.0035 0.0045 0.0055 0.0065

Figure 3. Graphical representation of compaction force-punch separation profile.

r1 Compaction x2
roll

r3 Turret
r2 Punch
wt
head
x2

Figure 4. Model for vertical displacement of a punch head on a rotary press.

12
Alternatively, punch distance can also be obtained from sensitive displacement

probes fitted into the tableting machines that are able to provide accurate punch

displacement data. The amount of energy utilized to compress a powder mass into a

consolidated tablet (net work of compaction; Wnet) and the energy recovered during

the decompression phase (work of expansion; Wexp) can be obtained by calculating

specific areas under the curve in Fig. 3. A is the punch separation at the first

measurable force and B is the minimum punch separation at peak compaction force

C. Punch separation after the decompression phase is represented by D. Area BCD

is equivalent to Wexp while Wnet is determined by the difference between area ABC

and area BCD. Wnet is an indicator of the ability of bonds formed during compaction

to survive the decompression process [39, 40]. An elastic material will exhibit a high

Wexp value and a low Wnet as the input energy for deformation will be largely

recovered during decompression. On the other hand, plastically deforming materials

usually possess high Wnet values as the energy input will be dissipated to a greater

extent for bonding.

1.1.3.4 Compaction parameters

The changes in certain compaction parameters are determined by the extent of

powder bed densification and this is dependent on the predominant deformation

mechanism of the tableting material. Ejection force (force required to push the

compact out of die) and residual die wall stress (pressure exerted onto the die wall

surface by the compact after formation) have been used by researchers to study

compact deformation [41]. A Compact that experiences a high degree of elastic

recovery usually exhibits lower ejection force and residual die wall stress values due

to the axial dimensional recovery of the compact during the decompression phase.

However, the above mentioned parameters may also be influenced by other factors

13
apart from compact deformation such as degree of adhesion at the compact-die wall

interphase. Punch penetration depth, which determines the tablet thickness at

maximum compaction in the die, is another parameter which can potentially be

utilized to study particle resistance to deformation; elastically deforming materials

generally allow reduced punch penetration.

1.1.4 Instrumentation in tablet compaction

The compaction behavior of tableting materials could be significantly altered by the

type of tablet making instrument used [42]. Research pertaining to tablet compaction

usually revolves around three types of compaction systems – rotary press,

compaction simulator and single station press.

1.1.4.1 Rotary tablet press

Commercial tablets are usually produced using rotary tablet presses for such presses

are capable of very high volume and low unit cost production. The basic compaction

operational cycle of a rotary tablet press is essentially the same from machine to

machine. It involves biaxial compaction with both the upper and lower punches

moving in unison towards one another to a specified target punch separation, upon

contact with the compaction rolls. The important operational features of the rotary

press related to tableting performance will be discussed in the subsequent sections.

1.1.4.1.1 Turret speed

Turret speed (tableting speed) is often expressed as revolutions-per-min (rpm). An

increase in turret speed narrows the compaction profile with concomitant reductions

in consolidation (increased strain rate), dwell and decompression times. Turret speed

and dwell time follow an inverse relationship, slowing the turret speed increases the

dwell time as it takes more time for the roll to transit across the head flat [43]. Tablet

14
physical quality as a function of tableting speed has been widely investigated, and

will be discussed under a later section.

1.1.4.1.2 Compaction rolls

The diameter of the compaction roll affects the compaction cycle significantly and

may impact the tablet manufacturing performance as a result. The compaction roll

diameter controls the consolidation and decompression rate, contact time with the

punch head, and the dwell time [44]. A reduction in the roll diameter reduces the

overall contact time and increases the strain rate during the consolidation phase

significantly. The changes in the roll dimensions could particularly be of importance

for materials that undergo time-dependent deformational changes.

1.1.4.1.3 Precompaction

Modern rotary tablet presses will invariably be equipped with dual compaction rolls

known as precompaction roll, positioned just after the feed frame, followed by the

main compaction roll. A compaction force that is a fraction of the main compaction

force is usually applied to the powder bed at the precompaction event. The primary

aim of this activity is to facilitate removal of air in the powder bed and increase the

effective contact time between powder particles under an applied force. During this

extended contact time, some interparticulate bonds may form and partial

neutralization of elasticity may occur [45]. Hence, application of precompaction may

aid in reducing tablet elastic recovery during decompression and strengthen the

tablets, which could reduce incidences of capping and lamination.

1.1.4.1.4 Press technology

Tablet press manufacturers have developed effective tableting systems incorporated

with innovative technologies that enable the modification of compaction parameters,

independent of turret speed and tooling configuration. Dwell duration is often the

15
target parameter for several existing press technologies. A press manufacturer,

Courtoy, manufactures tablet presses with a technology involving the use of air

compensators mounted over the compaction rolls (floating roll technology) [46].

When the compaction force applied exceeds a preset force value, the air compensator

would withdraw the roll upwards and cause displacement of the roll, thereby

maintaining a relatively constant applied force on the compact and consequently,

extending the dwell time without compromising the tableting speed. Another

innovative press technology is the use of a compaction dwell bar (Korsch) which

maintains the pressure on the compact between the precompaction and main

compaction rolls and thereby, extends the dwell time in the process [47].

1.1.4.2 Compaction simulator

A compaction simulator is a tableting machine usually fitted with a typical tablet

tooling setup - upper punch, lower punch and a die - and able to mimic vertical punch

movement during a full compaction cycle of the unit operation [48]. Simulators are

usually instrumented with sensitive sensors that allow monitoring of the

displacement and force profiles associated with the compaction event. Compaction

simulators can be programmed to mimic cycles of tablet press models or other

production conditions with relatively small amount of material. Data collected under

specified compaction parameters using the compaction simulator are relatable to

data from a high speed rotary press, due to the largely similar rapid compaction

kinematics involved [49, 50]. Compaction profiles obtained from the compaction

simulator may exhibit some subtle differences and this is partially attributed to the

lack of a compaction roll in the simulator.

16
1.1.4.3 Single station tablet press

A single station tablet press may be operated electrically or manually. The main

operational difference with single station tablet presses, differing from rotary

presses, is that the lower punch remains stationary during the compaction cycle

(single-ended compaction). The compaction kinematics of the single station presses

are often not comparable to the rapid compaction cycles observed in the above

mentioned tableting instruments. Thus, data collected from the single station tablet

presses may not be reflective of powder compaction during high speed rotary press

tableting, particularly when time-dependent particle deformation is of concern.

Nonetheless, compaction using a single station tablet press is of value when specific

tableting properties related to punch face configurational design, e.g. powder

adhesion due to embossing, or tablet properties independent of strain sensitivity are

studied.

1.2 Tablet physical quality

Tablet physical quality is often associated with the stresses necessary to disrupt the

intact tablet produced. Tablets must possess acceptable mechanical strength and

remain intact during handling and transport, between production and administration.

Optimization of tablet physical quality is an integral part of tablet formulation

development, process up-scaling to bulk manufacture. Tablet physical quality can

be characterized by evaluation of tablet tensile strength, tendency towards tablet

failure and tablet material adherence events.

1.2.1 Tablet tensile strength

Tablet tensile strength for round shaped tablets can be calculated using Eq. (2) [51].

2𝐹
Tensile strength = 𝜋 × 𝑑 (2)
𝑡 × 𝐻t

17
where F is the force needed to fracture a round shaped tablet of thickness Ht along

its diameter dt. The fracture force, often known as tablet hardness value is obtained

from a diametrical compression test (hardness test) in which the tablet is placed

against a platen and force is applied along an axis of the tablet by a movable platen

(Fig. 5). Force applied continuously increases until the tablet fails and the force at

failure is recorded as the fracture force. Tablet fracture force is influenced by the

physical dimensions of the tablet. Therefore, differences in the tablet shape must be

considered during the interpretation of tensile strength values. Tensile strength of

tablets derived from tablets of other shapes can be calculated by modification of the

existing formula. However comparison of values between different classes of tablet

shapes remains debatable.

Alternative methods of tensile strength measurement have been used in solid dosage

development. Flexure test, in which a tablet is bent, has been proposed as an

alternative to the diametrical compression test [52]. Another procedure of tensile

strength measurement involves pulling the tablet along the tablet’s main axes until

failure occurs, used primarily as a means to detect weaknesses in the compact in the

axial direction [53]. Nonetheless, the diametrical compression test remains the most

commonly used method for tablet tensile strength measurement.

Figure 5. Schematic showing the diametrical compression test.

18
1.2.2 Tablet failure tendency

Tablet failure often occurs in the form of capping. Capping refers to the partial or

complete removal of the top and/or bottom crowns of a tablet from the main body

(Fig. 6). This disruption of the tablet system can occur during decompression in the

die or upon the application of an external stress such as during diametrical

compression test. It is generally accepted that extensive tablet elastic recovery during

decompression is one of the primary factors responsible for the occurrence of defects

[54-56]. When the stresses produced by elastic recovery during decompression are

sufficient enough to disrupt the bonds that are formed during compaction, capping

occurs. Tablet capping has also been attributed to a number of other root causes: air

entrapment, mechanism of volume reduction, tableting speed, non-uniform stress

and density distribution as well as internal shear stresses due to die wall pressure.

Figure 6. Image showing absence of capping and tablet capping (left to right) after
hardness testing.

Capping tendency and tablet tensile strength share a positive relation to a certain

extent; formulation with higher capping propensity usually possess low tablet tensile

strength. Several descriptors of tablet failure propensity have been proposed by

researchers. The brittle fracture index (Eq. (3)) proposed by Hiestand and coworkers,

reflects the ability of the tablet to resist fracture during tableting [57, 58].

19
𝑇
𝐵𝐹𝐼 = 0.5 ( ) (3)
𝑇0

where T and T0 are the tensile strength of a compact without and with a center hole

respectively. Askeli et al. proposed calculation of a dimensionless index known as

capping index, an alternative method in which capping occurrence during the

production and testing stages are considered, using Eq. (4).

5𝑁p + 𝑁h
Capping Index = 𝑁t
(4)

where Np is the number of tablets that capped off the press, Nh is the number of

tablets that capped during hardness testing, and Nt is the total number of tablets tested

[59]. The equation differentiated the severity of capping incidences that occur during

rotary tablet production and hardness testing. Presence of capping during the

manufacturing process was considered more severe than capping observed during

the hardness testing.

1.2.3 Tablet material adherence tendency

Tablet material adhesion to punch face and die surfaces is a commonly encountered

problem in tablet manufacture, often referred to as ‘sticking’. Significant sticking

during tablet manufacturing leads to compromised tablet physical quality in the form

of poorer weight uniformity and marred appearance. Mild punch sticking results in

the formation of a material film on the punch face after a certain number of

compaction runs and can be visualized by inspecting the tooling. However, mild

sticking may not cause observable physical defects on the tablet surface. Severe

punch sticking leads to more visually apparent surface defects and will require

termination of the tableting process for cleaning, resulting in a drop of

manufacturing efficiency. The adhesive interactions responsible for the sticking of

particles to the surfaces are predominantly molecular in nature, mainly by van der

20
Waals forces [60-63]. However, other interfacial phenomena such as contact melting

and metallic interactions could also contribute to punch sticking potential [64]. The

sticking tendency of a particular formulation is often not easily detectable during the

early stages of development as adhesion problems usually become apparent at the

later stages of development when prolonged compaction runs are needed for larger

batch production.

1.2.4 Factors influencing tablet physical quality

Tablet physical quality is affected by interparticulate bonds formed within the tablet

matrix during the consolidation/dwell phase and lost during the bond disruptive

events during the decompression/ejection phase. Factors that influence these

processes consequently impact tablet physical quality as well. Some of these factors

related to the presented study are discussed in the subsequent sections.

1.2.4.1 Deformation characteristics of tableting materials

The predominant deformational characteristics of the drug and excipients impact the

evolution of voids and extent of bond formation in the tablet microstructure. Plastic

deformation and fragmentation of particles are the two processes that control the

degree of interparticulate bonding within the tablet microstructure. Fragmentation

affects the number of interparticulate bonds while plastic deformation relates

primarily to the effective area of contact between particles [65, 66]. Nonetheless,

both deformation mechanisms have been reported to enhance tablet tensile strength

[67].

Tablet failure events (e.g. capping) are usually related to low degrees of plasticity in

the compact system. Formulations with inadequate proportion of ductile materials

are more susceptible to bond disruption events during decompression by elastic

recovery. Plastically deforming materials allow greater absorption of the total energy

21
of compaction and less energy to be stored elastically in the compact system. This

reduces tablet elastic expansion during decompression and maintains tablet

mechanical integrity. A higher degree of plastic deformation within the tablet matrix

could also reduce powder adherence to tool surfaces by the formation of stronger

cohesive forces holding particles onto the compact body and therefore, a much larger

adhesive bond breakage energy would be required to enable particle detachment

from the compact onto the punch surface.

1.2.4.2 Lubrication

During the tableting process, friction is generated at the interface between the

compact and tool surface. Lubricants are added to tablet formulations to reduce die-

wall friction. Solid lubricants are usually of small particle size, act by boundary

mechanism in which the lubricant particles adhere onto lager particle surfaces in the

formulation and to the compaction tooling. Magnesium stearate is the most

commonly used boundary lubricant for tableting in the pharmaceutical industry. The

concentration of magnesium stearate in tablet formulations and duration of blending

with the lubricant have been shown to affect tablet mechanical strength; high

magnesium stearate concentrations and prolonged blending may be detrimental to

tablet strength [68]. This latter effect has been attributed to the formation of a

lubricant film around particles which may interfere with particle-particle bonding,

resulting in a less cohesive and mechanically weaker tablet [69]. Material that

predominantly densify via plastic deformation are more sensitive than fragmenting

materials to the deleterious effects of magnesium stearate. Fragmentation allows the

creation of fresh surfaces and reduces lubricant coated surface area that interferes

with bond formation, thus the lubricant film effect by magnesium stearate is reduced

[70]. Lubricants with anti-adherent properties may also be used to reduce powder

22
sticking by the formation of a film which reduces contact between powder

components and tool surfaces [71, 72].

1.2.4.3 Compaction force

Compaction force is a major factor in the powder densification process. An increase

in the force applied on the powder bed causes a greater degree of deformation on the

particles, increases area of contact between particles and consequently strengthens

the compact in the process. A positive linear relationship between compaction force

and tablet tensile strength can be observed for most tablet formulations until a

plateau is reached, indicative of optimum densification of the tableting material. The

linear portion of a tensile strength-force curve can be described by Eq. (5).

𝑇𝑆 = 𝐶𝑝 . 𝑃 + 𝑏 (5)

where P is the peak compaction pressure, TS is the tablet tensile strength and Cp is

the compactibility coefficient (calculated from the slope of the linear region). Cp

quantifies the tabletability of a specific compact system, indicating the effectiveness

of the applied force in enhancing interparticulate bonding within the tablet

microstructure. Plastically deforming tableting materials generally exhibit high

tabletability [73]. In contrast, fragmenting and elastically deforming materials

provide greater resistance to densification at higher compaction pressures, leading

to reduced tabletability. While increases in compaction force allow greater degree

of deformation and bonding, it is often accompanied by increases in stored elastic

energy. Hence, tendency towards tablet failure increases at higher compaction

forces. Powder tendency to adhere to punch face is similarly believed to be sensitive

to compaction force level; sticking problems could possibly be mitigated with the

use of higher compaction loads during tableting [74, 75].

23
1.2.4.4 Tableting speed

Tableting speed determines the rate of punch displacement/force application on the

compact. Adjustments to the tableting speed often alters the compaction cycle profile

and parameters (consolidation, dwell and decompression times), as described earlier

in Section 1.1.4.1.2. This affects how particles deform and therefore, determines the

integrity of the compact formed. Consolidation and dwell times are related to rate of

particle rearrangement and stress relaxation, respectively. The decompression phase

determines the rate of tablet recovery that occurs in the die.

The influence of tableting speed on tablet physical properties is dependent on the

deformation behavior of the tableting material [76-86]. Tableting materials with

viscoelastic and plastic character are more sensitive to tableting speed; increases in

tableting speed generally cause a deterioration in tablet mechanical properties.

Tablet failure events are more likely to occur under accelerated tableting conditions.

This is attributed to the shift in the elastic-plastic equilibrium in the compact brought

about by changes to the compaction cycle. At higher tableting speeds, there is

reduced time for stress relaxation and plastic flow to take place. This increases the

amount of elastic energy released, in the form of tablet expansion. As a consequence,

bonding quality is negatively affected. Studies have also shown that increases in

tableting speed may cause the yield pressure of the tableting material to increase

correspondingly; resistance to densification increases at higher tableting speed [66].

Brittle materials are believed to possess reduced sensitivity to tableting speed, as

fragmentation process occurs primarily during the consolidation phase at lower force

levels. However, contradictory results have been reported in literature.

Microcrystalline cellulose and lactose tablet properties have been demonstrated to

be independent of tableting speed, while dicalcium phosphate, which was expected

24
to possess low strain rate sensitivity, showed increased tabletability at higher

tableting speeds. This was attributed to more extensive brittle fracturing of dicalcium

phosphate particles at higher speeds.

1.2.4.5 Tablet press and tooling technology

The compaction behavior of pharmaceutical materials could be significantly altered

by the tableting technology and compaction tooling used [42]. Tablets compacted

using manual presses usually exhibit superior physical quality than tablets from

rotary presses due to the differences in compaction kinematics. Compaction cycles

in the manual press are much longer and this facilitates greater degree of time-

dependent deformation (plastic flow) to take place, leading to better tablet physical

quality. Rotary presses are often fitted with innovative tableting technologies to

improve tablet quality or productivity. Multiple compaction stations, extension of

dwell time and advanced vacuum system for removal of air within the powder bed

are available press technologies that may enhance interparticulate bonding in the

tablet structure and mitigate tableting problems. Some of these technologies have

been discussed in Section 1.1.4.1.5.

Similarly, tablet compaction tooling can affect the physical quality of the tablets

produced. Modifications to tooling surface properties and metal composition has

been reported to alter the liability of tablet sticking. Roberts investigated the

influence of punch face surface quality and the effect of chrome plated punches on

material sticking and claimed that punch surface quality is an important factor as

surface roughness positively correlated with powder adhesion [87]. However, a

smoother punch face and chrome plating on the punch faces may not necessarily

correspond to a reduction in material sticking onto punches [87]. Modifications to

punch configuration may also impact compact deformation process during the

25
compaction cycle. The influence of geometrical changes to the tableting punch

configuration on the compaction process and tablet physical quality will be

discussed in later sections.

1.3 Tablet compaction tooling

Tablet compaction tooling and formulation optimization equally contribute in

successfully ameliorating a tableting program. Tooling must be designed and

engineered to produce tablets of acceptable quality and be able to withstand stresses

associated with the compaction process for a prolonged service life. Understanding

the basic physics of powder compaction in relation to tooling modifications

enhances the ability to produce quality tablets more efficiently and provides greater

knowledge to troubleshoot and identify potential pitfalls of a tableting program.

1.3.1 General concepts

A single set of tablet tooling typically consists of an upper punch, a lower punch

(Fig. 7), and a die. Internationally recognized standards for tablet tooling includes

the TSM (Tablet Specification Manual) and EU (European) standards [88, 89]. Both

standards identify the physical tool configuration for B- and D-type compaction

tools, the majority of the tool configurations used today. Standardization of tooling

configuration allows interchangeability between tablet presses. B-type tooling

possesses smaller configurational dimensions and allows the insertion of a greater

number of punches in a turret for a higher tablet production rate. The larger body of

the D-type tooling is favorable for producing large tablets. The TSM and EU

standard manuals provide useful technical information such as critical dimensions,

associated tolerances for standard tooling and valuable references for trouble

shooting and tool inspection.

26
Head flat

Head
radius Head

Neck

Barrel

Punch face with concavity

Figure 7. Schematics showing the lower (left) and upper (right) punches.

27
Innovations in tablet press and tooling technology over the past decades have

improved efficiency by allowing increased tableting speed, improving product yield

and quality along with improvements in the sanitary and safety aspects. Tableting

tools are engineered by tool makers to withstand higher stresses, last longer and

produce tablets of better quality. In some cases, standard tooling configurations may

not achieve the expected tableting performance, particularly when tableting

materials have problematic compaction characteristics. In such cases, minor

modifications to the tooling configurations may be necessary for optimal tableting

performance. Modifications to punch head and punch face configurations are some

design tweaks that could potentially improve tableting performance.

1.3.2 Punch head configuration

Punch head configuration is clearly one of the most reworked areas on the tablet

tooling. The head flat and head radius are design features of the punch head that can

potentially be modified by tooling manufacturers to prolong tooling life and improve

product quality (Fig. 8). Punch head flat refers to the flat portion of the punch head

that makes contact with the compaction roll and determines the dwell time of the

compaction cycle. The flat top on the punch head provides a space for constant strain

onto the compact in the die while punches are sliding over the compaction rolls.

Head flat

Head radius

Figure 8. Schematic showing the punch head configuration with the associated head
flat and head radius.

28
Modifications to the punch head flat specification affect the compaction parameters,

consequently impacting the tablet mechanical properties. The diameter of the punch

head flat, turret speed and turret pitch circle diameter determine the dwell time as

shown in Eq. (6).

60 × 𝐷ℎ𝑓
𝑇𝑑𝑤𝑒𝑙𝑙 = 1000 (𝜋 × 𝑉 × 𝐷𝑡𝑢𝑟𝑟𝑒𝑡
) (6)
𝑡𝑢𝑟𝑟𝑒𝑡

where Dhf and Dturret are the diameters of the head flat (if oval, the longest dimension)

and turret respectively, Vturret is the velocity of the turret in revolutions-per-min

(rpm). Dwell time has been pointed out as one of the key compaction parameters for

tablet quality [90]. Depending on the tableting material densification characteristics,

increases in dwell time could potentially result in mechanically stronger tablets due

to a favorable shift in the elastic-plastic equilibrium in the compact and this leads to

better energy utilization towards bond formation. In view of this effect, some tooling

manufacturers provide punches with a larger head flat area. Enlarging the head flat

transverse length leads to a longer flat surface for the rolls to transit across the punch

head, consequently increasing the dwell time and providing a longer ‘relaxation’

time period for the compact. However, extension of dwell time via head flat

modification is limited by punch body configuration. The initial contact of the

compaction roll to the punch head should be within the diameter of the neck (Fig. 7)

so as to provide structural support [89]. Otherwise, punch head fracturing could

occur with prolonged use. Thus, the extent to which the head flat can be expanded

is rather limited. The type of tableting tools (B- or D-type configuration) affects the

dwell time as well. Due to the ability of the larger barrel to accommodate a larger

punch head and head flat profile, the use of D-type tooling usually leads to longer

dwell time. Despite the known effect of dwell on compact deformation, information

29
on the influence of changing punch head profile on compaction process and tablet

physical quality is rather limited in literature.

Head radius refers to the radius of the curved surface on the top of the head which

blends the middle section of the head to the head flat, often referred to as the ‘domed’

head configuration (Fig. 8). The head radius makes initial contact with the

compaction roll and allows for a smoother transition of the punches into the

compaction cycle. The presence of a head radius reduces the impact stress when the

punch makes initial contact with the compaction roll. The impact of the roll on the

punch head at high speeds and compaction forces may cause work-hardening issues

which are detrimental to both the compaction rolls and punches. Punch head radius

theoretically affects the compaction cycle by influencing the rate of compaction

force application on the compact during the consolidation phase. A reduced head

radius leads to the formation of a curved punch head with a steeper slope which

could possibly cause a reduction in the consolidation duration and increase the strain

rate during the consolidation phase.

1.3.3 Punch face configuration

Punch face configuration determines the tablet face contour for shaped tablets.

Configurational modifications to the face design can lead to profound differences in

the tooling life, powder compaction process and the resultant tablet quality.

Optimization of punch face configuration is essential for a successful tableting

program.

During compaction, the punch face surfaces experience lateral stress that causes

expansion of the sides of the punch cup outward toward the die wall momentarily.

Excessive stresses developed during compaction beyond the maximum allowable

force level for a particular face configuration may lead to permanent distortion and

30
premature failure of the punch. The maximum allowable compaction force for a

particular punch face configuration is determined by the punch face configuration,

area and stress concentration factors. Some punch face configurations possess areas

that experience high levels of stress due to transitions in their geometric profile,

resulting in a lower maximum compaction force. Punch face configurations could

be modified to strengthen the punch. Reduction of the cup depth and increasing the

cup radius are two possible modifications. A deep concave cup configuration

facilitates development of high stress points on the cup edges while configurations

with a reduced depth aid in the distribution of the stresses on the punch face surface

and could potentially lengthen tooling life. Similarly, modification of punch face

edge could also potentially improve tableting performance. The flat-face bevel edge

(FFBE) tablet shape is a popular choice among tablet manufacturers due to the

availability of a top flat surface for embossing purpose (Fig. 9B). In comparison to

the flat-face plain (FFP) shape (Fig. 9A), the FFBE shaped tablets exhibit reduced

surface attrition. A simple modification of the bevel edge to a curved surface with a

radius (Fig. 9C) strengthens the punch and allows a higher compaction force to be

applied without causing punch tip damage and eliminates edge chipping on the tablet

surface.

In relation to powder compaction, modifications to punch face configuration affect

powder flow and packing in the die under stress conditions. During the tableting

process, differential movement of powder within the die occurs; central regions

experience greater movement than peripheral regions due to frictional forces along

die wall-compact interphase. Punch face configurations that lack concavity cause

uniform but restricted movement of particles along the flat surface of the punch face

and limit the powder movement along the die walls.

31
A

Figure 9. Schematics showing the FFP (A), FFBE (B) and FFRE (C) tablet designs
with difference in tablet edges shown. (FFP: flat-face plain, FFBE: flat-face bevel
edge, FFRE: flat-face radius edge)

The introduction of concavity to the punch face promotes in-die particle movement

to a greater degree due to the inherent greater volume to fill towards the center of

the compact [91]. David et al. demonstrated that the extent of powder movement

was dependent on the relative curvature of the punch face [92]. Increasing curvature

led to a corresponding increase in the lateral particle movement within the powder

body. Similar findings were revealed in a study by Brunel et al., in which a unique

X-ray imaging method was used to study powder movement [91]. FFP punch face

was shown to promote greater lateral powder movement, primarily at the top corners

of the tablet away from the die wall with corresponding minimal radial movement.

Radial powder movement was greater in convex curved tablets at regions adjacent

to die wall and along the edges of the punch surface.

32
Differential powder movement due to punch face curvature leads to variations in the

density distribution within the tablet and formation of non-homogenous tablet

internal character [93]. Density distribution pattern of uniaxially compacted flat-

faced tablets generally adhere to the classical Train explanation; high density regions

at corners and central regions of the tablet with concomitant stress relief within the

parts of the tablet adjacent to the moving punch face [94]. Introduction of a curved

concavity to the punch face increases variation in the density distribution producing

tablets with much more heterogeneous internal structure, leading to the creation of

density boundaries between regions of high and low densities which are susceptible

to fracture failure [95, 96]. This could consequently impact tablet physical quality

and their tendency to cap.

Modifications in the punch face configurations have also been reported to cause

changes to the net resolution of the resultant compaction force within the compact

and lead to heterogeneous stress distribution within the tablet. The use of a punch

face with a curvature causes the net resolution of the resultant compaction force to

be nearer to the moving punch face and towards the central regions of the tablet [97].

It is generally believed that the formation of non-uniform stress distributions due to

different punch face configurations is correlated to the tablet mechanical properties.

Kadiri and Michrafy studied the effect of punch face configuration on compaction

force transmission through the powder compact and demonstrated that increasing

punch depth (reducing curvature radius) created a distinct stress distribution pattern

with the potential to initiate crack formation and propagation [98].

Apart from being an influential factor on tablet mechanical properties, punch face

configuration also contributes to powder adherence during tablet compaction,

related to stress concentration factors on the punch face. Optimization of punch face

33
configuration via modifications to cup design and engraving style is often required

for formulations prone to punch sticking. Waimer et al. studied the influence of

tablet engravings with respect to powder adhesion using an instrumented upper

punch and claimed that concentration of shear stresses at the lateral faces of the

engraving during compaction and stress distribution within the tablet are important

contributors to strong powder adhesion [99]. Similar findings were reported in

another study in which embossment on the punch face increased powder adhesion

considerably, attributed to the probable concentration of shear stresses at the lateral

faces of the embossed logo [100]. Modification in the punch curvature from flat-face

configurations to concave configurations was also shown to reduce powder adhesion

potential, possibly due to variation in the stress distribution on the punch face.

1.4 Computational simulation of powder compaction

Computational modeling of powder compaction has generated growing interest as

direct measurement of particle movement and behavior during the compaction

process in the die is challenging. In view of this limitation, computational modeling

has been utilized as an effective tool for predicting compaction dynamics.

Computational modeling of powder compaction has typically been carried out by

two approaches: finite element modeling (FEM) and discrete element modeling

(DEM) methods [101].

1.4.1 Finite element modeling (FEM)

FEM treats the powder as a continuous material with elastic-plastic material property

throughout the continuum solid. Although a pharmaceutical powder mass is clearly

discontinuous at the particle level, this becomes irrelevant at the larger scale of

aggregation such as when it is compacted into a relatively dense compact in a die

during tableting. Therefore, the compaction behavior of pharmaceutical powders can

34
possibly be studied using principles of continuum mechanics [102]. The Drucker-

Prager Cap model is widely accepted to be a good constitutive model for modeling

powder compaction as it considers the densification and hardening character of

powders and captures the shearing phenomenon in powders during the different

stages of compaction [103-107]. Materials and compaction condition specific

parameters are required to simulate different compaction conditions. Sinha and

coworkers demonstrated the possibility of obtaining the necessary model input

parameters via simple compaction procedures [108]. The authors compared the

simulation of powder compaction using constant property and relative density

dependent property based Drucker-Prager Cap models and demonstrated that while

the relative density dependent property model proved to be more accurate, the

constant property model could be considered reasonably accurate. The use of a

relative density dependent property model may actually be disadvantageous due to

the requirement of high computational power and time-consuming calibration steps.

The Drucker-Prager Cap model has been used in FEM to study density and stress

distribution of a consolidated powder for identification of potential tablet failure

modes in compacted tablets. Wu et al. demonstrated that FEM is able to capture the

essential aspects of the behavior of pharmaceutical powders during compaction and

can predict possible mechanism of crack propagation during the decompression

phase from the examination of the evolution of stress distribution during unloading.

Localized intensive shear bands developed from the tablet edge to the central regions

of the tablets, indicative of failure patterns within the tablet structure [109]. The

influence of punch face configuration on density/stress distribution and the relation

with tablet mechanical properties has been investigated by a number of researchers

with the employment of FEM. Han et al. demonstrated that introduction of concavity

35
to the punch face caused large shear stresses at the top corner of the compacts, due

to radial elastic recovery of the material at the top central region upon removal of

the upper punch force [110]. The simulated results provided a possible explanation

for the more frequent occurrence of capping in concave tablets. In a similar study,

Kadiri et al. demonstrated that dense regions with strong shear stresses were formed

at the upper edges and the effect of punch face design on stress distribution became

more significant with changes in punch depth [98]. The influence on tablet thickness

and punch curvature on density distribution was further investigated by Diarra and

co-workers [95]. Flat-face plain and convex tablets compacted at equivalent force

levels exhibited different density distribution at the central region of the tablets and

increases in curvature radius promoted a smoother density gradient at the tablet

edges. From a tooling perspective, FEM enables accurate identification of stress

concentration factors on the punch face and this allows optimization of tooling

design by manufacturers [111].

1.4.2 Discrete element Modeling (DEM)

DEM models consider the powder, one particle at a time. The rules of the model

determine the extent of particle interaction. The particles move in the predetermined

simulation area, and collisions or contacts between them are analyzed by the

simulation [112-114]. Since every particle is handled separately, the number of

particles largely determines the computing time and outcome of the simulation. As

such, during development of the model, the properties of the particles, such as

density and yield strength, become inputs for the model.

Compaction induces very complex states of stress in the packing structure and

involves microscopic interactions between particles. FEM models are based on

continuum mechanics that operate largely at a bulk scale and focus on

36
phenomenological description of a compaction process. Bulk response of a compact

obtained in FEM may not be related to the microscopic information. In comparison,

DEM simulations allow detailed microscopic information of compacts to be

obtained. DEM allows the processes involved during compression and consolidation

- particle rearrangement and particle deformations - to be mimicked more precisely

with accurate particle properties. Sheng et al. successfully simulated uniaxial

compaction of powder particles using a DEM model and also showed that the DEM

results correlated well with experimental results [115]. Frenning et al. utilized a

combined finite/discrete element model to simulate compaction of particles and

showed that realistic systems comprising 1000 or more particles may be successfully

analyzed within a reasonable computation time [116]. In another study, Siiria et al.

used an improvised DEM model to calculate bond strength distribution in cylindrical

tablets and showed that the estimated tablet strength increased as a function of the

compaction force in the same way as the measured breaking strength did in the

tablets [117].

While DEM has been successfully applied to simulate particle packing, it has its own

set of limitations that requires further improvement. Theoretically, DEM allows

evaluation of an infinite number of particles. However, this may not be feasible in

practice. DEM calculates the interaction of each particle. A large number of particles

will require long unrealistic computation duration for a complete analysis of the

interactions. Thus, normally, the number of particles are kept to a maximum that

allows realistic computing times. Adjustments to particle and physical parameters

are usually needed to ensure that simulated particle movement could be fitted

properly to experimental bulk powder behavior [118].

37
1.5 Research gaps

Tablet compaction process has been extensively studied, numerous publications and

theories are readily available. There have been considerable literature reports on the

effect of formulation factors on tableting performance. In comparison, the influence

of process factors on the compaction process has been less studied. Among the

limited studies that investigated tableting process factors, research on tablet

compaction tooling is scarce in published literature.

Tablet compaction tooling exerts significant influence on the densification behavior

of particle and consequently impacts the tableting outcomes. However, research on

the specific impact of tooling configurational modifications on the compaction

process and resultant tablet properties has been limited. Existing research data on

powder compaction could not be extrapolated to predict the influence of tooling

modifications on compaction parameters and tablet properties, especially during

high speed tableting as compaction studies reported in literature were often

conducted under tableting conditions that were not representative of commercial

tablet production rates as encountered during actual tablet manufacture. In addition,

discrepancies in the results of tablet properties have been reported in the literature,

mainly due to the large number of varied methods used for compaction studies and

formulation dependent outcomes. These research gaps lead to a need to better

understand the influence of tooling configurational modifications on the tableting

process and quality of tablets produced. Fundamentally, the study of compaction

tooling design is necessary to elicit the compaction physics involved as this could

aid in making judicious choices of processing considerations and tool designs for a

successful tableting program.

38
1.5.1 Dwell effect on tablet quality

As previously described under Section 1.1.2, compact deformation during the dwell

phase differs from other stages of the compaction cycle. Compacts experience

greater plastic flow under constant strain, which could potentially impact tablet

strength. There are limited publications correlating the duration of dwell phase to

the resultant tablet physical quality. In a study by Vezin et al, it was reported that

dwell time extension could be less important than lag time (separation time between

precompaction and main compaction) prolongation in affecting tablet mechanical

strength, attributed to enhanced deaeration of the powder bed during the longer

separation time between the two distinct compaction events [29]. However,

contrasting results were reported in another study where dwell time was found to

exert a more pronounced influence than lag time in affecting tablet physical quality

[119].

The influence of dwell time on tablet physical quality in relation to different

deformation mechanisms of tableting materials has also been explored. Tye and co-

workers claimed that the impact of dwell phase on tabletability was dependent on

the predominant deformation mechanism of the tableting material [120]. Highly

ductile materials were unaffected by dwell time changes while materials with

viscoelastic nature were substantially affected. However, in another study when

pectin (brittle material) was compacted, significant reductions in peak force were

observed and tablets of superior physical quality were produced with prolongation

of dwell time [121]. The results indicated some degree of viscoelastic character in

the pectin compact despite its brittle nature.

The results from the various studies often did not concur and could be attributed to

differences in the amount of plastic and viscoelastic components in the tablet

39
formulations used. Dwell duration used in existing studies were often not

representative of high speed tablet compaction and may not provide an accurate

understanding of dwell influence on compact deformation process.

1.5.2 Punch head configurational modification

Modifications to the punch head configuration affects the compaction profile and

parameters with potential consequences on mechanical properties of tablets

produced. Alterations to the head flat and radius specifications could possibly

change the dwell time and consolidation strain rate during consolidation phase.

Sarkar et al. highlighted the complex relationship between dwell time and strain rate

during the consolidation phase and the impact on the resulting tablet properties

[122]. It was shown that the influence of strain rate on tablet strength and capping

tendency could predominate over dwell time, particularly when the compaction

force employed is above the minimum force required for particle deformation.

Higher strain rates reduced the time for particles to rearrange during the

consolidation phase, and had possibly contributed to more air entrapment. These

effects consequently impacted interparticulate bond formation negatively and led to

extensive tablet capping. However, extended dwell time at a low precompaction

force followed by application of a main compaction force improved the tablet

physical quality and alleviated the capping situation. It was explained that the

prolongation of dwell time during the precompaction stage allowed more time for

particle movement and air escape, resulting in improved conditions for compact

formation at the main compaction stage and led to greater densification and bonding

quality within the tablet microstructure.

While a considerable number of publications related to tableting is available, the

impact of geometrical changes to the head flat and radius on the tableting process

40
has not been well investigated [123]. Existing studies have largely focused on the

effect of compaction parameters on tablet properties. In most studies, the compaction

parameters were modified by adjustments to the compaction kinematics, e.g.

tableting speed, which resulted in uncontrolled concomitant changes to other

parameters of the compaction cycle and could not be extrapolated in predicting the

effect of punch head and radius modifications on the parameters and the resultant

tablet properties.

1.5.3 Punch face configurational modification

Punch face configuration is an important feature of the tableting punch that affects

tableting performance significantly, as described in Section 1.3.3. Modifications to

punch face configuration impart differences to powder movement during

densification, leading to variations in the density distribution within the tablet.

Formation of a non-homogenous tablet internal character may cause the tablet matrix

to be susceptible to fracture failure which could affect tablet physical quality. In

comparison to punch head configuration, the influence of punch face configurational

modification has been more widely studied. Stress/density conditions in the tablet

matrix due to different punch face specifications have been correlated to tablet

physical quality, particularly in terms of tablet failure and powder adherence

tendency.

Majority of the studies related to punch face configuration involved concave shaped

punches, with different degrees of curvature. Research on the influence of punch

face modifications with regard to flat-face tablets of different edge geometries has

been rather limited. Considerable differences in cup depth and other geometrical

specifications between flat-face and concave punch face configurations restricts

extrapolation of results from existing publications in predicting the effect of flat-face

41
punch edge geometrical modifications on tablet compaction.

42
CHAPTER 2

HYPOTHESES AND
OBJECTIVES

43
2. HYPOTHESES AND OBJECTIVES

This study was motivated by the need to provide a more comprehensive

understanding on the influence of tooling configurational modifications with respect

to tablet properties as this knowledge could potentially aid in making process and

formulation considerations for a successful tableting program. The findings could

also facilitate the design of better compaction tools as there would be a clearer

understanding of the impact of minor tool modifications to the quality of tablets

produced.

The lack of studies specific to tooling configuration was identified as a research gap

in this area of compaction research. Process and formulation conditions adopted in

many compaction studies were often not representative of actual manufacturing

conditions and many studies produced contradicting evidence on tablet properties.

These factors restricted the extrapolation of such data for predicting the effect of

tableting punch head and face modifications on the tableting process and physical

quality of the tablets to be produced.

Therefore, the main hypothesis of this work is that tableting punch head and face

configurational modifications of flat-face punches can alter compaction parameters

and the resultant tablet physical quality.

Investigations with the following objectives was conducted to test the research

hypothesis:

A. To investigate the influence of formulation variables on compact deformation

during dwell phase;

B. To investigate the influence of punch head configuration on compaction process

and tablet physical quality;

44
C. To investigate the influence of punch face configuration on compaction process

and tablet physical quality; and

D. To investigate the influence of punch face configuration on tablet material

adhesion during tablet compaction;

The research work done was organized into four parts:

Part 1. Investigation of the influence of deformation mechanism of tableting

materials during dwell, and the consequent effect on tablet mechanical properties

Part 2. Investigation of the effect of punch head modification on compaction

parameters and tablet physical quality

Part 3. Investigation of the effect of punch face modification on compaction

parameters and tablet physical quality

Part 4. Investigation of the effect of punch face modification on powder adhesion

tendency to the punch surface during powder compaction

45
CHAPTER 3

EXPERIMENTAL

46
3. EXPERIMENTAL

3.1 Part 1: Investigation of the influence of deformation mechanism of tableting

materials during dwell, and the consequent effect on tablet mechanical

properties

3.1.1 Rationale of study

Tableting outcomes are closely related to the predominant deformation mechanism

of the tableting material. Compact systems with different degrees of plastic, elastic

and fracture deformation may react to constant strain during dwell phase differently,

consequently impacting resultant tablet properties. Therefore, it was of interest to

investigate the influence of dwell on plastic flow in different compact systems. Three

types of tableting materials with contrasting densification behavior were used in this

study. Influence of dwell time extension on compact deformation and mechanical

properties of tablets compacted from different formulation systems were examined.

3.1.2 Choice of tableting materials

Microcrystalline cellulose (MCC; Avicel PH101, FMC Biopharma, PA, USA),

dicalcium phosphate (DCP; Emcompress, JRS Pharma, Rosenberg, Germany) and

paracetamol (APAP; Granules India Limited, Hyderabad, India) were chosen as

model materials in this study. Microcrystalline cellulose and dicalcium phosphate

have been studied extensively in literature, representative of materials that deform

plastically and brittle fracture respectively [11, 124]. Paracetamol is generally

considered to predominantly undergo elastic deformation, often used in studies as a

model drug in formulations prone to tablet failure [59, 122].

47
3.1.3 Characterization of tableting materials

3.1.3.1 Heckel analysis

Heckel analysis was utilized to study densification behavior of the tableting

materials. Heckel plot was constructed using a universal testing machine (Autograph

AG-10KNE, Shimadzu, Kyoto, Japan). The universal testing machine allows

compaction of a powder bed at high compaction pressures, under highly controlled

settings. It is fitted with a 100 kN load cell which controls the rate of load

application. The load cell applies a compaction load on the upper punch, which

transmits the force into the powder bed confined within the customized die

(positioned in the die plate). A 10 mm diameter flat-face punch, with a

corresponding 10 mm die was used for compact formation. Heckel plot was

constructed for each of the three types of tableting material, using an out-of-die

method. An accurately weighed amount (325 mg) of powder was manually fed into

the die. The compaction speed was fixed at 10 mm/min, programmed using the in-

built software Trapezium (Shimadzu, Kyoto, Japan). Tablets were formed at

compaction pressures ranging from 25 to 150 MPa. The compaction process was

programmed to allow immediate withdrawal of the load upon reaching the stipulated

maximal pressure, without dwell time. Post compaction tablet thickness was

measured using a digital micrometer (Series 293, Mitutoyo, Kawasaki, Japan).

Compressed densities of the compact at each specified compaction pressure were

calculated based on the measured compact dimensions. The true densities of the

tableting materials were determined using a helium pycnometer (Pentapycnometer,

Quantachrome Instruments, FL, USA). The intra-tablet porosity is related to the

compressed and true densities of the deformed particles. Porosity (ɛ) values of the

compacts were calculated using Eq. (7). The Heckel equation (Eq. (1)) was

employed to characterize densification behavior of the tableting materials.

48
𝐶𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑒𝑑 𝑑𝑒𝑛𝑠𝑖𝑡𝑦
ɛ=1− 𝑇𝑟𝑢𝑒 𝑑𝑒𝑛𝑠𝑖𝑡𝑦
(7)

The Heckel profiles of the materials were analyzed based on the evolution of the

Heckel curve with increasing compression pressure. A typical Heckel plot usually

consists of an initial curved region at the lower compression pressure regions

followed by a linear segment at higher pressures. The initial curved region which is

indicative of the fragmenting tendency of the particles were compared. Materials

with a steep initial curve at the lower compressive pressures may be suggestive of a

brittle nature. Slopes of the linear portion and onset pressure at which the linear slope

begins were also analyzed to provide a relative comparison of the deformation nature

of the materials at high pressures, in terms of elastic/plastic deformation. A high

slope of the linear section of the Heckel plot is representative of plastically

deforming material with a low yield pressure. Interception of the vertical axis from

the extrapolation of the linear segment is related to the apparent density of the

powder and particle packing before force application.

3.1.3.2 Particle size analysis

Particle size analysis was conducted on the tableting materials using a laser

diffraction particle sizer (Small Volume Module LS 230, Beckman Coulter, IN,

USA). The dx value corresponded to the Xth percentile under the cumulative

undersize distribution curve. The particle size expressed in terms of the median

particle size (d50) and the size distribution was calculated in terms of span using Eq.

(8).

𝑆𝑝𝑎𝑛 = (𝑑90 − 𝑑10 )⁄𝑑50 (8)

49
3.1.3.3 Hausner ratio

Bulk and tapped densities of the tableting materials were determined for the

calculation of the Hausner ratio. The material was passed through a sieve of aperture

size 1.7 mm into a 100 mL graduated cylinder. The top of the powder was gently

levelled with a spatula, powder adhering dusted off and the mass of powder in the

cylinder accurately weighed. The powder in the graduated cylinder was then

mechanically tapped to a constant volume using a tap density tester (Sotax TD2,

Sotax, Aesch, Switzerland). The final tapped volume was noted and used to calculate

tapped density. The Hausner ratio was calculated using Eq. (9).

𝑃𝑡𝑎𝑝𝑝𝑒𝑑
𝐻𝑎𝑢𝑠𝑛𝑒𝑟 𝑟𝑎𝑡𝑖𝑜 = (9)
𝑃𝑏𝑢𝑙𝑘

where Ptapped and Pbulk represent the tapped and bulk densities respectively. The

Hausner ratio indicates the flowability of a particular powder blend; larger values

indicate poorer flow characteristics [125].

3.1.3.4 Angle of repose

The angle of repose (AOR) was determined using a powder flowability tester

(Flowability tester BEP 2, Copley, Nottingham, UK). The powder was passed

through a sieve of aperture size 1.0 mm into a funnel equipped with a 10 mm nozzle

mounted 75 mm above a 100 mm diameter circular test platform. The powder was

allowed to fall gradually onto the platform to form a conical heap. The height of the

formed heap was measured using a digital height gauge (Absolute, Mitutoyo,

Kawasaki, Japan). Angle of repose was calculated as the angle between the slopes

of the conical heap relative to the horizontal base. Smaller values for the angle of

repose indicate powders with better flow characteristics [126].

50
3.1.4 Preparation of tableting blends

Since the objective of this study was to investigate the influence of formulation

variables on compact deformation during dwell, binary blends (APAP/MCC,

DCP/MCC and APAP/DCP) were prepared by mixing the respective tableting

material in predetermined proportions (percent proportions; 0/100, 25/75, 50/50,

75/25 and 100/0) in a polyethylene bag for 10 min. The final formulation blends

were prepared by mixing the binary blends with magnesium stearate 1 % w/w

(M125; Productos Metalest, Zaragoza, Spain) in the polyethylene bag for 1 min,

prior to tableting.

3.1.5 Tablet production

Tablets were produced from each formulation blend at compaction forces of 6 and 9

kN (compaction pressures, 76.4 and 114.6 MPa) using the universal testing machine

(Autograph AG-10KNE, Shimadzu, Kyoto, Japan). An accurately weighed amount

(325 mg) of powder was manually fed into the die. Compaction speed was kept

constant at 10 mm/min. Tablets were compacted at different dwell durations ranging

0-2 s for each tableting formulation.

3.1.6 Stress relaxation test

Viscoelastic behavior of the compacts during dwell period was evaluated from stress

relaxation tests. Stress relaxation test was conducted on each formulation blend

using the universal testing machine with a 10 mm diameter flat-face punch and

compaction speed of 10 mm/min. The compaction load cell was programmed to

pause at the position where it achieved the maximal stated force for a predetermined

duration, upper punch displacement was maintained during that period. The

reduction in force of the upper punch in relation with time was monitored during the

dwell period. The force measured from the load cell during this period is

51
representative of the resistance of the compressed material to densification. Plastic

flow due to viscoelastic deformation during constant strain (dwell) causes force

decay in the compact, reflected in the stress relaxation profile [25, 26].

The degree of viscoelastic compact deformation during dwell phase was analyzed

using different viscoelastic parameters calculated from the stress relaxation profile

for each formulation. The ratio of the maximum force (CSmax) at the beginning of

the dwell phase and the force measured by the upper punch after the defined time of

dwell (CS0) provides the viscoelastic parameter, relaxation quotient (RQ).

𝐶𝑆𝑚𝑎𝑥
RQ = (10)
𝐶𝑆0

RQ provides an indication of overall plastic flow during dwell time [28]. However,

it should be noted that RQ assesses the compaction process during dwell and does

not describe the compaction process preceding dwell phase. The slopes of the linear

sections of the stress relaxation profile provides valuable information related to the

rate of plastic flow during dwell phase. Viscoelastic slopes (VS) were calculated by

subjecting the stress relaxation profile of each formulation blend to linear regression

analysis with the identification of the linear sections (R2 > 0.99) of the profiles. This

method has been reported to be informative of the compact deformation process

during dwell [23].

3.1.7 Tablet characterization

3.1.7.1 Viscoelastic strain recovery

Viscoelastic strain recovery refers to the post compaction time-dependent tablet

expansion that occurs during storage period without being restrained by die wall

[127]. Tablet viscoelastic recovery and immediate elastic recovery are related [128].

It is generally believed that majority of the stored energy due to elastic deformation

52
of particles is released via immediate elastic recovery during decompression.

However, tablets produced from materials of viscoelastic character may also

undergo retarded elastic energy release during storage leading to dimensional

changes in the tablet during storage. The percent viscoelastic strain recovery was

calculated by the equation (Eq. (11)) proposed by Maganti and Celik [129].

𝐻72hr − 𝐻dcp
VER (%) = 𝐻dcp
× 100 % (11)

where tablet thickness immediately upon ejection (Hdcp) and tablet thickness 72

hours post compaction (H72hr) were measured by an out-of-die method using a digital

micrometer (Series 293; Mitutoyo, Kawasaki, Japan).

3.1.7.2 Tablet tensile strength

Mechanical quality of compacted tablets was determined from the calculation of

tablet tensile strength (Eq. (2)) which is indicative of interparticulate bonding

quality. Breaking force of tablets was obtained from diametrical compression testing

using a hardness tester (TBF 1000, Copley, Nottingham, UK). Tablet diameter and

thickness 72 hours post compaction were measured using a digital micrometer

(Series 293, Mitutoyo, Kawasaki, Japan).

3.1.8 Statistical analysis

Statistical analysis of the data was performed using PASW Statistics 18 Software

(SPSS Inc., Chicago, IL, USA). ANOVA was used for analyzing multiple groups of

data or statistical differences. Results with p-value of less than 0.05 were statistically

significant.

53
3.2 Part 2. Investigation of the effect of punch head modification on compaction

parameters and tablet physical quality

3.2.1 Rationale of study

Tableting punch head is a key component of the punch tool design, capable of

impacting tableting outcomes especially during high speed tableting. Modifications

to the punch head configuration could potentially affect dwell time along with other

associated compaction parameters. Thus, it was of interest to investigate the

influence of punch head flat and head radius configurational changes on the

compaction process and the mechanical properties of the tablets produced.

3.2.2 Choice of tableting materials

Paracetamol and lactose based formulations were used in this study. Both

formulations have been reported to be susceptible to tablet failure in the form of

capping, particularly at high tableting speeds. The elastic character of the

formulations could provide information pertaining to changes in plasticity/elasticity

balance in the tablets.

3.2.3 Preparation of tableting blends

Lactose granules (Tablettose 80, Meggle Pharma, Wasserburg, Germany) were used

as provided, without further treatment. Granules containing paracetamol (Rhodadap

Dense Powder, Rhodia Wuxi Pharmaceutical, Jiangsu, China) and potato starch

(Roquette, Lestram, France) (77.9:22.1) were produced using a high shear processor

(UltimaPro 10L; GEA Collette, Wommelgam, Belgium), based on a formulation

utilized by Sarkar et al in a study on capping tendency [122]. A 15.2 %, w/w starch

paste, which was used as a binder in the later stages, was prepared by adding potato

starch to water. Dry mixing of paracetamol powder and potato starch (19.9 %, w/w)

in the high shear processor was carried out for 2 min at an impeller speed of 450 rpm

54
prior to starch paste addition and the mixture wet massed for an additional 5 min

with the impeller and chopper speeds set at 450 rpm and 2800 rpm, respectively. The

moistened granules produced were passed through a cone mill with square-hole

screen of aperture size 6350 µm (Comil 197S, Idex-Quadro Engineering, Waterloo,

Canada) and impeller speed of 1240 rpm. Granules were dried in a fluidized bed

drier (Strea-1, GEA-Aeromatic, Bubendorf, Switzerland) with inlet air temperature

of 45°C and drying was terminated when product temperature reached 40°C. The

dried granules were re-granulated through the cone mill with round-hole screen of

aperture size 1143 µm and impeller speed 1350 rpm. Final tableting blends were

formed by mixing lactose and paracetamol-starch granules separately with 1 %, w/w

magnesium stearate (M125, Productos Metalest, Zaragoza, Spain) in a polyethylene

bag for 1 min, prior to tableting.

3.2.4 Characterization of tableting material

3.2.4.1 Particle size analysis

The particle size of paracetamol-starch and lactose granules was measured using the

laser diffraction particle sizer as described in Section 3.1.3.2.

3.2.4.2 Hausner ratio

The Hausner ratio of paracetamol-starch and lactose granules was obtained from the

measurement of bulk and tapped densities of the powder blends as described in

Section 3.1.3.3.

3.2.4.3 Angle of repose

The flow characteristics of the paracetamol-starch and lactose powders were

evaluated by measuring the angle of repose as described in Section 3.1.3.4.

55
3.2.4.4 Heckel analysis

The compression behavior of the paracetamol-starch and lactose granules was

studied using the Heckel plots as described in Section 3.1.3.1.

3.2.4.5 Stress relaxation test

The general viscoelastic behavior of the paracetamol-starch and lactose granules

were determined from the stress relaxation test as described in Section 3.1.6.

3.2.5 Tablet production

3.2.5.1 Tableting punch head configuration

Five different types of 10 mm flat-face bevel edge (FFBE) punches that could be

classified into two main punch head configurations – with and without a head flat –

were used for this study [130]. The specifications of the punch head configurations

are given in Fig. 10. The punches differed mainly in terms of their head flat and head

radius specifications. Three punches with head flats - extended head flat (EHF),

standard head flat (SHF) and reduced (half) head flat (RHF) - were used. The

remaining two punch head configurations completely lacked the head flat with one

having a standard radius head (SRH) and the other having a reduced radius head

(RRH). To accommodate the changes to the head flat and head radius specifications,

the punch head thickness was modified accordingly to keep the outside diameter of

the head (25.57 mm) and overall punch length (133.60 mm) consistent with tooling

standards. All other punch specifications adhered to EU B-type tooling standards.

56
A B C

HF HR HF HR HF HR

HT HT HT

HF 13.34 HF 10.41 HF 5.21


HR 16.00 HR 16.00 HR 18.00
HT 2.89 HT 3.18 HT 3.15

D E

HR HR
HT HT

HR 33.00 HR 27.43
HT 2.51 HT 3.08

Figure 10. Schematics showing the dimensions (in mm) of the EHF (A), SHF (B),
RHF (C), SRH (D) and RRH (E) punch head configurations in terms of their head
flat (HF), head radius (HR) and head thickness (HT) specifications. (EHF: extended
head flat, SHF: standard head flat, RHF: reduced head flat, SRH: standard radius
head, RRH: reduced radius head)

57
3.2.5.2 Rotary tableting

The rotary tablet press (R190FT; GEA-Courtoy, Halle, Belgium) used in this study

consists of a turret fitted with 14 tableting stations. A schematic diagram

representing the interior of the rotary tablet press is shown in Fig. 11. The rotary

press is capable of two compaction events – precompaction and main compaction.

The precompaction event consists of a compaction roll of diameter 150 mm with an

air compensator mounted over the precompaction roll. The main compaction event

is fitted with a larger compaction roll of diameter 240 mm, without an air

compensator system connected to the roll. The compaction roll control the extent of

upper/lower punch penetration into the die. Maximum upper punch displacement at

the main compaction event could be varied between 1.8-4.0 mm, while upper punch

displacement at the precompaction event was fixed at 3.25 mm. The lower punch

displacement varies with the amount of tableting material in the die and the desired

compaction force. Main compaction and precompaction forces were determined by

the set compact thickness (distance between the upper and lower punch tips) at the

respective events. A reduced set compact thickness value for a particular punch face

configuration would lead to application of a higher compaction load on the powder

bed.

Adjustment of the compaction parameters in the rotary press enabled deactivation of

the precompaction event, allowing both compaction rolls to mimic main compaction

event. This enabled the compaction of each tablet by single compaction cycles using

either the compaction roll of diameter 240 mm (D240) or the smaller compaction

roll of diameter 150 mm (D150), with the upper punch displacement fixed at 3.25

mm at both compaction events (Fig. 12).

58
Main compaction
stage

Ejection
stage
Precompaction
stage

Die filling
stage

Figure 11. Schematic showing the rotary press turret with positions of the die filling,
precompaction, main compaction and ejection stages.

(i) (ii)
A
Precompaction Main compaction
roll roll

B
Compaction force

C Time
Compaction force

Time
Figure 12. Schematic diagrams depicting the (i) motion of the punch at
precompaction event and (ii) motion of the punch at main compaction event (A).
Graphical representation of the force-time curve at the main compaction event, with
precompaction deactivated (B). Graphical representation of force-time curve at
precompaction event with main compaction deactivated (C).

59
When the effect of one compaction event was studied, the other compaction event

was deactivated by setting a high set compact thickness at precompaction or main

compaction roll so that the die fill passed the respective event without being

subjected to any compaction force. During the compaction of tablets using the larger

compact roll, a precompaction thickness value larger than the fill depth and a

negligible precompaction force limit was set to prevent any punch penetration into

the die at the precompaction event, led to a compaction force-time curve of only the

main compaction event as shown in Fig. 12B. Similarly, when compaction using

the smaller compaction roll was desired, a main compaction thickness value larger

than the fill depth was set, ensuring that punch penetration ended above the die or

die fill at the main compaction event, giving a force-time curve of only the

precompaction event (Fig. 12C). Tablets were produced at compaction forces

ranging between 6 and 9 kN (pressure range, 76.4-114.6 MPa), Tablet weight was

kept constant at 325 mg and tableting speed range 25-35 rpm (linear speed, 34-47

m/min) was used for compaction.

3.2.6 Compaction parameters

Compaction force-time profiles (Fig. 2B) were captured by the in-built software

CDAAS (Courtoy Data Acquisition and Analysis System). Contact time (Ttotal; time

for compaction and decompression excluding ejection time), dwell time (Tdwell; time

during which compaction force is above 90 % of its peak value), consolidation time

(Tcon; time from start of compression to dwell phase), decompression time (Tdec; time

from the end of the dwell phase to end of the compaction cycle) and ejection force

were recorded from the profiles. The rate of force application (RFA) represents the

strain rate during consolidation phase (Fig. 2), was calculated using MATLAB

(R2010a; The MathWorks, MA, USA) [122]. Gaussian distribution equation (Eq.

60
(12)) was used to fit the data using the least-square error algorithm (r2 close to 1). At

the inflexion point, the second-derivative of the equation is equal to 0. By solving

d2y/dx2 = 0, a tangent (Eq. (13)) was drawn at the inflexion point (x0, y0). The slope

of the tangent, m, calculated from Eq. (14) represents the RFA during the

consolidation phase.

𝑥−𝑏 2
𝑦 = 𝑎𝑒 −( 𝑐
)
(12)

2
𝑥0 −𝑏 −(𝑥0−𝑏)
𝑦 = −2𝑎 𝑐2
𝑒 𝑐 (𝑥 − 𝑥0) + 𝑦0 (13)

𝑥0 −𝑏 −(𝑥0−𝑏)2
𝑚 = −2𝑎 𝑒 𝑐 (14)
𝑐2

where a, b and c are constants.

3.2.7 Evaluation of compact deformation

The extent of compact deformation during the dwell phase was quantified by

calculating the area quotient (AQ) for the compaction profiles using Eq. (15). AQ

indicates the extent of compaction force decay during the dwell phase, caused by

stress relaxation of the compact under constant strain [34, 35].

𝐴2
AQ = 𝐴1
(15)

The dwell phase was identified from the compaction profile and divided into halves

as shown in Fig. 2B. The areas under the force time curve were calculated for each

half (A1 and A2) using the graphing software (v9.1, OriginLab, MA, USA). Both

areas were decreased by the rectangles under the minimum compaction force in the

dwell time. The ratio of the spatial areas (A2/A1) provides the AQ (Eq. (15)) which

is a sensitive and informative parameter that quantifies deformation behavior. Lower

AQ values are indicative of greater plastic flow during dwell phase.

61
3.2.8 Tablet characterization

3.2.8.1 Tablet elastic recovery

Tablet elastic recovery (ER) was calculated using Eq. (16) [128]. Tablet height after

decompression process (Hdcp) was measured by an out-of-die method using an

electronic micrometer (Series 293, Mitutoyo, Kawasaki, Japan) while the tablet

height at target compaction force (Hcp) was the set compaction thickness by adjusting

the position of the respective compaction rolls to achieve the required force level

with respect to the tablet weight.

𝐻dcp − 𝐻cp
ER (%) = 𝐻cp
× 100 % (16)

3.2.8.2 Tablet porosity

Porosity of the tablets was calculated using Eq. (7). Densities of compacts at each

specified compaction force were calculated using the tablet dimensions upon tablet

ejection. The true density of the granules was determined using a helium pycnometer

(Pentapycnometer, Quantachrome Instruments, FL, USA).

3.2.8.3 Tablet tensile strength

Breaking force of tablets was measured using a hardness tester (TBF 1000; Copley,

Nottingham, UK). Tensile strength measurement of paracetamol-starch and lactose

tablets were calculated using Eq. (2). Tablet diameter and thickness 24 hours post

compaction were measured using a digital micrometer (Series 293; Mitutoyo,

Kawasaki, Japan). Tabletability (Eq. (5)) was quantified using the compactibility

coefficient (Cp) which is a dimensionless compaction parameter calculated from the

slope of the tensile strength-compaction force plots [15]. Tabletability describes the

effectiveness of the applied force in increasing tablet tensile strength and provides

an insight into the compaction efficiency of the tooling.

62
3.2.8.4 Capping index

Capping index was calculated using Eq. (4) that was proposed by Akseli et al, as

described in Section 1.2.2 [59]. Due to difficulties in batching product output from

a rapid tableting process run, batching of samples were based on a fixed production

duration (15 mins). The quantity of tablets produced and the number of tablets that

capped off the press during tablet production and hardness testing (TBF 1000,

Copley, Nottingham, UK) were recorded as inputs for Eq. (4).

3.2.9 Computational simulation of tablet compaction

Tablet compaction was simulated using Finite Element Modeling (FEM). The

Drucker-Prager Cap model was implemented in the FEM software (Abaqus, Ansys,

PA, USA). The use of Drucker-Prager Cap model for numerical studies on powder

compaction has been discussed under Section 1.4.1. The yield function was defined

by three surfaces represented in Fig. 13: shear failure surface Fs, elliptical surface

(cap surface) Fc and the transition surface Ft [103]. The evolution of the cap surface

was described with the hardening function Pb which is the position of the cap on

hydrostatic pressure axis for each density state. For a uniaxial cylindrical die

compaction test, the hydrostatic pressure stress (p) and Mises equivalent stress (q)

were expressed as:

1
𝑝 = − 3 ( 𝜎𝑧 + 2𝜎𝑟 ) (17)

𝑞 = |𝜎𝑧 − 𝜎𝑟 | (18)

where σz and σr are the axial and radial stresses, respectively. Six parameters were

required to define the yield surface of the Drucker-Prager Cap model (β, d, pa, R, pb

and a). In order to identify these parameters, diametrical and uniaxial compression

(Fig. 14) tests were carried out [98].

63
Mises equivalent stress q
Transition surface, Ft

Shear failure, Fs α(d + patan β)

d + pa tan β
Cap, Fc
β

d R(d + pa tan β)
pa pb

Equivalent hydrostatic pressure stress p

Figure 13. Drucker-Prager Cap model with associated parameters.

Figure 14. Schematic showing the uniaxial compression test for material property
calibration.

64
The friction angle (β) and cohesion (d) are needed to define the Drucker-Prager Cap

shear failure surface; the cap eccentricity parameter (R) and evolution (pa) are

required to define the cap surface, and pb as a function of the volumetric plastic strain

required to define the cap hardening/softening law; a was required to define the

transition surface

Material parameter identification for the Drucker-Prager Cap model:

Cohesion and internal friction angle were expressed as:

𝑓 𝑓
𝜎𝑐 𝜎𝑑 (√13−2)
𝑑= 𝑓 𝑓 (19)
𝜎𝑐 + 2𝜎𝑑

𝑓
3(𝜎𝑐 +𝑑)
𝛽 = tan−1 [ 𝑓 ] (20)
𝜎𝑐

where σfd is the radial tensile strength of tablets measured from the diametrical

compression test, and σfc is the axial tensile strength of compacts measured from the

uniaxial compression test. Radial and axial tensile strengths of the tablets were

determined using Eq. (21) and Eq. (22) respectively.

𝑓 2𝐹
𝜎𝑑 = 𝜋𝐷𝑡
(21)

𝑓 4 𝐹𝑦
𝜎𝑐 = 𝜋𝐷 2
(22)

where F is the breaking force obtained from the diametrical compression test and Fy

is the axial crushing force obtained from the uniaxial compression test, using the

universal testing machine. D is the tablet diameter and t is the tablet thickness. a is

a small number (typically 0.01 – 0.05), used to define a smooth transition surface

between the shear failure surface and the cap, was set to 0.02 [103]. Material

parameter R, that controls the shape of the cap, is given as:

65
2 (1+𝛼−𝛼/ cos 𝛽)2
𝑅= √ 3𝑞𝐴
(𝑝𝐴 − 𝑝𝑎 ) (23)

Consequently, the evolution parameter pa and pb were obtained from Eq. (24) and

Eq. (25) respectively.

[ 3𝑞𝐴 +4𝑑 tan 𝛽(1+𝛼−𝛼/ cos 𝛽)2 ]


𝑝𝑎 = − 4 [ (1+𝛼−𝛼/ cos 𝛽) tan 𝛽 ]2
+

2 + 24𝑑𝑞 (1+𝛼−𝛼/ cos 𝛽)2 tan 𝛽+8(3𝑝 𝑞 +2𝑞 2 (1+𝛼−𝛼/ cos 𝛽) tan 𝛽 ]2
√9𝑞𝐴 𝐴 𝐴 𝐴 𝐴 )[

4 [ (1+𝛼−𝛼/ cos 𝛽) tan 𝛽 ]2


(24)

𝑝𝑏 = 𝑝𝑎 (1 + 𝑅 tan 𝛽) + 𝑅𝑑 (25)

3.2.10 Statistical analysis

Statistical analysis of the data was performed using PASW Statistics 18 Software

(SPSS Inc., Chicago, IL, USA). ANOVA was used for analyzing multiple groups of

data or statistical differences. Results with p-value of less than 0.05 were statistically

significant.

66
3.3 Part 3. Investigation of the effect of punch face modification on compaction

parameters and tablet physical quality

3.3.1 Rationale of study

Tableting performance of punches may however differ with different punch face

configurations. The edges of cylindrical tablets are prone to damage as the edges,

produced by the right-angle meeting two large surfaces (cylindrical and flat

surfaces), tend to contain more loosely pack particles. Thus, friability of these tablets

is higher. The solution adopted by tool manufacturers to overcome this deficit is to

provide bevel edge flat punches. Modification of the angled bevel edge geometry to

a curved radial surface leads to the formation of the flat-face radius edge punch face

configuration, a relatively new design proposed as an alternative to bevel edge

punches. Investigation of the influence of such edge geometry modifications of the

flat-face punches on compaction process and tablet properties could provide useful

information for tablet and tooling manufacturers.

3.3.2 Preparation of tableting blends

Lactose and paracetamol-starch granules described in Part 2 of the study, were used

in this part of the study as well. Tableting blends were prepared following the

procedure described under Section 3.2.3.

3.3.3 Tablet production

3.3.3.1 Tableting punch face configurations

Two types of flat-face punch face configurations with different edge geometries,

with standard punch head, were used for this part of the study (Fig. 15 and Table 1)

[131]. The flat-face bevel edge (FFBE) face configuration consists of an angled

straight edge between the flat portion and the land of the punch tip, with the bevel at

an angle of 30° (Fig. 15A).

67
A

0.4
8.22
10.0 30°

0.4
7.09
10.0
R2.50

Figure 15. Schematics showing the dimensions (in mm) of the FFBE (A) and FFRE
(B) punch face configurations. (FFBE: flat-face bevel edge, FFRE: flat-face radius
edge)

Table 1. Specifications of FFBE and FFRE punch face configurations (FFBE: flat-
face bevel edge, FFRE: flat-face radius edge)

Punch Cup Cup Cup Working


face depth volume area length
3 2
configuration (mm) (mm ) (mm ) (mm)

FFBE 0.4 26.03 81.47 133.2


FFRE 0.4 25.10 80.75 133.2

68
The flat-face radius edge (FFRE) face configuration (Fig. 15B) consists of a

curvature with a radius (R) of 2.50 mm between the flat portion and the land. In

order to ensure that the radius does not exceed the 30° bevel used for the FFBE

configuration, the flat area for FFRE configuration was reduced accordingly, causing

the cup area to be reduced marginally. The punch face specifications are stated in

Table 1. All other punch specifications adhered to EU B-type tooling standards.

3.3.3.2 Rotary tablet production

Tablets were compacted using single or double compaction cycles. In single

compaction cycles, only the main compaction event (compaction roll of diameter

240 mm) was used, with the precompaction event deactivated by setting a high

precompaction set compact thickness and low precompaction force limit, which

disabled the air compensator mounted over the precompaction roll and led to a large

compaction stroke distance between the top and bottom punches so that the die fill

passed the event without being subjected to any compaction force. This method was

described earlier under Section 3.2.5.2. The main compaction force was determined

by the set compact thickness (ST main) at main compression roll. A smaller STmain

value for a particular punch face configuration led to greater compaction force on

the compact. In the double compaction cycles, both main and precompaction events

were used. Two different tableting modes (with and without the air compensator

operating) were utilized at the precompaction event (Fig. 16-17). When the air

compensator was disabled (mode1), the precompaction event acted like the main

compaction event (without floating compaction roll) as shown in Fig. 16, with the

precompaction force determined by the set compaction thickness (STpre) at

precompaction event. The precompaction force was set to 3 kN.

69
A (i) (ii)
PC roll MC roll

B
Compaction force

Main compaction
curve

Precompaction
curve

Time

Figure 16. Schematic diagrams depicting the motion of the punch at precompaction
(PC) event (i) with air compensator deactivated (mode1) and (ii) motion of the punch
at main compaction (MC) event (A). Graphical representation of force-time curve at
precompaction event with air compensator deactivated (mode1) and the force time-
time curve at the main compaction event (B).

A (i) (ii)
PC roll
MC roll

B
Roll displacement
Compaction force

Main compaction
Roll displacement curve
curve

Time

Figure 17. Schematic diagrams depicting the motion of the punch at precompaction
event (i) with air compensator activated (mode2) and (ii) motion of the punch at main
compaction event (A). Graphical representation of force-time curve at
precompaction event with air compensator activated (mode2) with the associated roll
displacement curve and the force-time curve at the main compaction event (B).

70
In the other precompaction mode (mode2), the air compensator was activated by

preseting a force limit (3 kN). When the precompaction force exceeded the force

limit, the air compensator floated the roll, causing it to displace upwards (Fig. 17),

and thereby maintaining a relatively constant applied force during the dwell phase,

with the dwell time extended accordingly.

3.3.4 Compaction energy

The amount of work done by the punches during compaction was calculated from

the compaction force-punch separation profiles for each flat-face punch

configuration [84]. Punch separation was approximated by calculating the vertical

punch displacement (z) at time t using Eq. (26).

1
z = [(𝑟1 + 𝑟2 )2 − (𝑟3 sin 𝑤𝑡 − 𝑥2 ) 2 ]2 (26)

where r1 and r2 are the radii of the compaction roll and punch head curvature

respectively, w is the angular velocity of the turret with a radius of r3, and x2 is the

radius of the punch head flat, described in Section 1.1.3.3.3 [28]. Net work of

compaction (Wnet) and work of expansion (Wexp) were obtained from the calculation

of specific areas under the force-punch separation curve using the graphing software

(v9.1, OriginLab, MA, USA) [80]. The ratio of work of expansion and compaction

provides the compaction quotient (Eq. (27)) which quantifies the change in elastic

recovery relative to the change in the degree of interparticulate bonding.


𝑊𝑒𝑥𝑝
𝐶𝑜𝑚𝑝𝑎𝑐𝑡𝑖𝑜𝑛 𝑞𝑢𝑜𝑡𝑖𝑒𝑛𝑡 = 𝑊𝑛𝑒𝑡
(27)

71
3.3.5 Tablet characterization

3.3.5.1 Tablet elastic recovery

Elastic recovery of the tablets produced from the different punch face geometries

were calculated as described in Section 3.2.8.1.

3.3.5.2 Tablet porosity

Porosities of tablets produced from the different punch face geometries were

calculated as described in Section 3.2.8.2.

3.3.5.3 Tablet tensile strength

Tensile strength of tablets produced from the different punch face geometries were

calculated as described in Section 3.2.8.3.

3.3.5.4 Capping index

Capping tendency of tablets produced from the different punch face geometries were

evaluated as described in Section 3.2.8.4.

3.3.6 Computational simulation of tablet compaction

Discrete Element Modelling (DEM) was used as a tool for examining the effect of

punch face configuration on particle compaction within the tablet in this part of the

study. DEM is capable of incorporating particle physical displacement into

simulation, likely to occur with punch concavity modifications. FEM, which was

utilized in Part 2 of the study for studying the influence of dwell time on compact

stress state, treats the simulated powder as a continuum mass, less suitable for

particulate study. The software EDEM 2.4 (DEM Solutions Ltd., Edinburgh, UK)

was used for the simulations. Particles were simulated as spheres of diameter 0.5

mm, where the contact forces were calculated using the Hysteretic Spring contact

model in EDEM for its ability to simulate elastic and plastic deformation behavior

72
[132]. The material properties used for the simulated paracetamol-starch

particles/tableting tools are as follows: density = 1347/7667 kg/m3, shear modulus =

1.0/79.3 GPa, yield pressure = 166/1830 MPa and Poisson’s ratio = 0.3/0.3. For

simplicity, the wall properties of the tooling and particles’ surface were set the same

in the simulations. Coefficients of restitution and static friction were set at 0.1 and

0.3, respectively. In order to acquire reasonable computation times, the resistance

of the particles to compaction was reduced to 0.001 times the actual resistance

(computational time = 30 ms). This reduced the interaction forces and friction forces

between the particles. However, the relative densities and particle deformation

behavior remained the same as if the original forces had been used. This approach

has been employed in a number of previous studies and found to accurately capture

the dynamics of compaction process [117].

Particle deformation behavior is significantly affected by particle size. To ensure

that the deformation behavior of simulated particles was stable and reliable, the ratio

of compaction load to total particle surface area was kept constant. Compaction

pressures for simulation purpose were calculated using experimental and simulated

particle sizes. Simulated tablet thickness at target compaction pressure and rate of

force application (RFA) were recorded as validation parameters.

3.3.6.1 Compact stress analysis

Stress distribution within the simulated tablet matrix at the target compaction force

was approximated by the simulations. Reduction of the acquisition time to construct

the stress data was enabled by using a grid bin group consisting of a suitable number

of bins at the tablet (Y-Z) cross-section. The average particle stress for each bin was

calculated. Stress distribution data was exported to a graphical software (v9.1,

OriginLab, Northampton, MA, USA) to construct contour plots representing relative

stress distribution. A calibration bar which relates the scale to the apparent stress is

73
added to allow simple understanding and comparison of the different stress

conditions in the tablets.

3.3.7 Statistical analysis

Statistical analysis of the data was performed using PASW Statistics 18 Software

(SPSS Inc., Chicago, IL, USA). ANOVA was used for analyzing multiple groups of

data or statistical differences. Results with p-value of less than 0.05 were statistically

significant.

74
3.4 Part 4. Investigation of the effect of punch face modification on powder

adhesion tendency to the punch surface during powder compaction

3.4.1 Rationale of study

Punch face modification could possibly alter stress concentration factors at the

punch face-compact boundary and impact powder adhesion tendency during powder

compaction. Investigation of the influence of the different face configurations of flat-

face punches on the extent of powder sticking during tablet compaction could

provide valuable information for tablet and tooling manufacturers.

3.4.2 Choice of tableting materials

Ibuprofen (Hubei Granules-Biocause Pharmaceutical, Hubei, China) was used as a

model active ingredient in this study. Ibuprofen is a popular choice in powder

adhesion studies, due to its poor compaction properties and high tendency for

adhesion to compaction tools [62]. The high adhesion propensity of ibuprofen

particles has been attributed to high particle surface energy and relatively lower

melting point in comparison to other APIs such as paracetamol. Ibuprofen particle

adherence to punch surfaces is reportedly sensitive to compaction speed and force

[63]. Compressible lactose granules (Tablettose 80, Meggle Pharma, Wasserburg,

Germany) was used as the filler based on its adequate flow properties, superior

compactibility and minimal contribution to punch adhesion.

3.4.3 Preparation of tableting blends

A direct compaction tableting blend, consisting of ibuprofen (70 %, w/w), lactose

(28.5 to 29 %, w/w) and magnesium stearate (1-1.5 %, w/w) (M125; Productos

Metalest, Zaragoza, Spain), was prepared. Ibuprofen and lactose mixture were

premixed manually in a polyethylene bag for 10 min. The required amount of

75
magnesium stearate was added and the formulation blended for a further 1 min, prior

to tableting.

3.4.4 Characterization of tableting blends

3.4.4.1 Particle size analysis

The particle size and size distribution of ibuprofen and lactose particles were

evaluated as described in Section 3.1.3.2.

3.4.4.2 Angle of repose measurement

The angle of repose of ibuprofen-lactose powder blend was measured as described

in Section 3.1.3.4.

3.4.5 Tablet production

3.4.5.1 Punch face configuration

Three types of 10 mm flat-face punches with different face configurations were used

for tableting in this study. Two of the experimented punch face configurations, flat-

face radius edge (FFRE) and flat-face bevel edge (FFBE) were used for tablet

compaction in Section 3.3.3.1. The FFRE configuration consists of 7.09 mm flat

circle and a curvature with a radius (R) of 2.50 mm between the flat circle and the

land. The flat-face bevel edge punch face consists of 8.22 mm flat circle and bevel

straight edge inclined at an angle of 30º between the flat circle and the land (Fig.

15). The cup volume and area for edged geometries have been reported in Table 1.

The remaining punch face configuration, flat-face plain (FFP) lacks a concavity and

cup volume with punch face area 78.5 mm2.

3.4.5.2 Tablet production using manual press

Two compaction methods were employed for tablet production. The first

compaction method involved a semi-automated manual press (NP-RD10A, Natoli,

76
St Charles, MO, USA). The manual press allows uniaxial compaction of tablets, with

a stationary lower punch. A photograph of the semi-automated manual press is

shown in Fig. 18. The hand press consists of a hydraulic system which is used to

control the vertical upper punch movement to achieve the desired compaction force,

monitored via the force sensor connected to the lower punch. Once the desired input

compaction force is reached, the upper punch withdraws automatically. Tableting

material (325 mg) was weighed out and transferred manually into the die for each

tablet preparation. Batches of tablets amounting to 30 or 50 tablets were produced

at compaction forces ranging between 5 and 10 kN. Triplicates were obtained for

each tableting batch at the various compaction forces and magnesium stearate

concentration levels. The use of the manual press for tablet compaction provides

information on the effect of uniaxial compaction and time dependent compaction

factors on powder adhesion.

Hydraulic cylinder

Ejection
lever
Upper punch
assembly Manual
pump

Force and fill


depth meter

Figure 18. Manual press used for tablet production.

77
3.4.5.3 Tablet production using compaction simulator

The second compaction method utilized an instrumented compaction simulator

(STYL’ONE EVOLUTION; MedelPharm, Beynost, France) for biaxial powder

compaction. A photograph of the tooling setup and the force feeder is shown in Fig.

19.

Upper
punch
assembly

Lower punch
and die
assembly

Force
feeder

Take-off
force sensor

Figure 19. Tooling setup (A) and the force feeder with the take-off force sensor
(B) in the compaction simulator.

78
The compaction simulator consists of a single station with upper and lower punches

capable of vertical displacement. Force and displacement sensors fitted at both the

upper and lower punch positions allowed capturing of force data and punch

movement during the time course of each compaction cycle. The compaction

simulator was programmed to compact the powder using single compaction cycle.

A feeder system was utilized to transfer the tableting material into the die during

tablet production. Batches of 50 tablets were produced at compaction forces ranging

between 5 and 10 kN. The use of the compaction simulator allowed the compaction

cycle to be shortened to milliseconds range, representative of rotary tablet

compaction.

3.4.6 Assessment of powder adherence

3.4.6.1 Quantification of powder adherence

Powder adhesion during compaction using the manual press was quantified via

spectroscopy analysis of the amount of ibuprofen adhered to punch face [87]. After

each batch of tablet production, the upper and lower punches were carefully

dismantled from the press. Powder adhered to the peripheral surfaces of the punch

tip was carefully removed using a standardized cleaning procedure, with minimal

disturbance to powder adherence state on the punch face. The punch tip was

immersed in 5 mL of 96 %, v/v ethanol to allow dissolution of ibuprofen adhered to

the punch face. The receiving solution was then analyzed by ultraviolet-visible

spectroscopy at 264 nm using a diode array spectrophotometer (U-5100, Hitachi

High-Tech Science Corporation, Tokyo, Japan). Absorption was converted to

concentration using a separately constructed standard curve. Overall, the linear range

was obtained at 0.2-0.8 absorbance, corresponding to 145-600 µg/mL concentration

of ibuprofen. The amount of ibuprofen adhered onto each punch face was recorded

79
as amount of ibuprofen adhered per unit area (µg/mm2) to remove confounding

impact due to differences in surface area. The amount of adhered residual powder

on the upper punch face during tablet production using the compaction simulator

was analyzed using spectroscopy procedure as described above. Powder adherence

on the lower punches were evaluated in terms specific force parameters measured,

discussed in the subsequent sections.

3.4.6.2 Force parameters related to powder adherence

Powder adhesion during tablet production using the compaction simulator was based

on measurement of ejection force and take-off force at the lower punch for each

tablet [133]. Ejection force is a measure of the force required to overcome the

residual die wall pressure and the frictional conditions at the interphase between the

tablet surface and the die wall during the compact travel to the die surface (Fig. 20A).

Ejection force may provide valuable information related to material-die wall

sticking.

Take-off force is the force required to detach the formed tablet off the lower punch

face after the tablet has been ejected from the die, a parameter that evaluates shear

stress during desorption of tablet from the surface of the lower punch (Fig. 20B)

[75]. Take-off force is generally considered as the most direct method to quantify

powder adhesion on the lower punch. Take-off force was measured by the take-off

piezoelectric sensor, positioned on the scraper bar located on the feeder as shown in

Fig. 19B, while the ejection force was determined by the lower punch force sensor.

Ejection and take-off force values for each tablet, over the time course of tablet

production, were individually captured by the in-built software ANALIS

(MedelPharm, Beynost, France).

80
A

Fr Tablet Fr

Fe

Lower Die
punch

Ft

B Fa

Figure 20. Schematics showing the stress factors associated with tablet ejection (A)
and take-off (B). (Fe: ejection force, Ft: take-off force, Fa: adhesive force, Fr:
compact-die wall force)

3.4.7 Statistical analysis

Statistical analysis of the data was performed using PASW Statistics 18 Software

(SPSS Inc., Chicago, IL, USA). ANOVA was used for analyzing multiple groups of

data or statistical differences. Results with p-value of less than 0.05 were statistically

significant.

81
CHAPTER 4

RESULTS AND
DISCUSSION

82
4. RESULTS AND DISCUSSION

4.1 Part 1: Investigation of the influence of deformation mechanism of tableting

materials during dwell and the consequent effect on tablet mechanical

properties

4.1.1 Physical characteristics of tableting materials

Particle size and flow properties of the tableting materials (MCC, DCP and APAP)

are presented in Table 2. The particle size of MCC and APAP were significantly

smaller than DCP particles, as shown by the d50 values. DCP powder exhibited

superior flowability with low Hausner ratio and angle of repose (AOR) values.

Table 2. Physical characteristics of microcrystalline cellulose (MCC), dicalcium


phosphate (DCP) and paracetamol (APAP) particles

Tableting d50 Span Hausner AOR


material (µm) ratio (°)

MCC 56.3 (4.1) 1.31 (0.07) 1.55 (0.02) 43.0 (0.7)


DCP 164.6 (3.8) 1.13 (0.08) 1.19 (0.02) 30.6 (1.4)
APAP 48.2 (4.1) 1.16 (0.02) 1.65 (0.05) 54.9 (0.4)

Parenthesized values represent the respective standard deviation.

4.1.2 Compression behavior of tableting materials

Heckel analysis was utilized to identify the predominant deformation behavior of

the tableting materials (Fig. 21). Heckel plots provide information about the

compression behavior of substances used and allow some empirical classification of

them into basic densification mechanisms. The intercept of the vertical axis, via

extrapolation of the linear region is related to the apparent density of the powder.

While it is difficult to accurately distinguish between elastic and plastic deformation

83
components from the slopes of the linear portion of the Heckel plots, the steepness

of the slopes provides a relative comparison of the deformation nature of the

materials at high pressures. Linear section with a high slope value represents

predominant plasticity during densification. Brittle fracturing and elastically

deforming materials generally exhibit low linear slopes.

2.5

2
Ln (1/e)

1.5

0.5

0
0 50 100 150
Compression pressure (MPa)

Figure 21. Heckel plots for microcrystalline cellulose (♦), dicalcium phosphate (■)
and paracetamol (●) particles

Heckel curves for APAP, DCP and MCC are present in Fig. 21. Densification

behavior for the tableting materials largely conformed to the expected trends.

Compression of DCP showed an initial curve region, indicative of a fast volume

reduction process dominated by particle rearrangement and fragmentation. The

subsequent long quasi-linear portion with a less steep slope indicated prevailing

plastic flow, in combination with elastic deformation. After initial brittle fracturing

of the DCP particles at low pressures, the stress necessary to cause further fracture

of the smaller particles increased. As a result, the mechanism of volume reduction

84
evolved from fragmentation to elastic deformation, reflected in the reduced rate of

porosity change in the linear region. Extrapolation of linear portion suggested the

highest degree of packing in the die for the DCP powder. Compression behavior of

MCC particles was distinctively different as shown by the Heckel plot with linear

curve and steep slope. The much greater slope of the Heckel plot for MCC suggested

a higher degree of plasticity and easier densification at the higher applied pressures.

Plastic densification behavior of the MCC particles has been attributed to the

numerous slip planes and dislocations present in the structure, with hydrogen

bonding between deformed particles [134]. Heckel curve of APAP particles showed

an initial region of reducing compact porosity with increasing compression pressure,

subsequent linear region began at a relatively higher compression pressure (delayed

onset) with the lowest slope. This demonstrated a predominance of elastic

deformation in the APAP particles. The high resistance of the APAP particles to

deformation has been attributed to energetically unstable movement of molecular

slip planes under external mechanical stress, resulting in reduced compliance to

shear stress and rate of densification [135-137].

4.1.3 Compact deformation during dwell

Deformation of material under compressive force during the dwell phase differed

from other stages of the compaction cycle, Heckel analysis may not be suitable for

dwell phase deformation characterization. Therefore, stress relaxation curves and

parameters related to plastic flow were used in the evaluation of viscoelastic

deformation of the compacts during dwell. Measured force observed from the stress

relaxation profile is a direct consequence of the resistance exerted by the material

under compaction, dependent on the physical properties of the substances which may

undergo time-dependent volumetric changes. The process of compaction force

85
decay within the compact during dwell could be divided into two distinct phases.

During the consolidation stage, rearrangement of particles causes the formation of a

compact comprising certain number of pore sites. The first phase begins immediately

after the compact has reached maximum intended compaction force. During this

phase, particles tend to undergo further deformational changes and move into void

interparticulate spaces in the compact. Part of the stored elastic energy in the

compact could be released at this stage, characterized by a steep drop in force. The

second phase is a retarded deformational response of the particles, in which the

particles deformed further in a plastically manner but at a lower rate, shown by the

reduced slope within the stress profile. The onset of the second phase is indicative

of a compact microstructural state with minimal available free sites. Resistance to

movement of deformed particles increases due to high internal friction caused by

close and constrained contact between particles, consequently reducing the

possibility of plastic flow at this stage [25]. Viscoelastic slopes for the first (VS1)

and second (VS2) phases were used in conjunction with the relaxation quotient (RQ)

values for analysis of the stress relaxation profiles.

The changes to the stress relaxation process with different levels of

elasticity/plasticity were analyzed in the APAP/MCC (Fig. 22 and Table 3) and

DCP/MCC (Fig. 23 and Table 4) compact systems. Increasing the amount of MCC

in the APAP/MCC compact system increased the overall extent of force reduction

and the rate of relaxation (Fig. 22). As expected, the viscoelastic slope of the second

phase was lower than the first phase as evidenced by the VS values in Table 3. The

enhancement in structural relaxation was largely attributed to the basic deformation

mechanism of the two types of tableting materials.

86
9

8.8 F1
8.6
Force (kN)

8.4
F2
8.2
F3
8

7.8 F4
F5
7.6
0 0.5 1 1.5 2
Time (s)

Figure 22. Stress relaxation profiles of APAP/MCC compacts of different


proportions. (APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100)

Table 3. Viscoelastic parameters obtained for APAP/MCC compacts of different


proportions (APAP/MCC; F1: 100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100)

Formulation Compaction RQ VS1 VS2


force (kN) (kN/ms) (kN/ms)

F1 6.0 1.03 0.28 0.04


F2 6.0 1.10 0.93 0.08
F3 6.0 1.14 1.20 0.12
F4 6.0 1.18 1.55 0.13
F5 6.0 1.21 1.73 0.16

F1 9.0 1.02 0.32 0.03


F2 9.0 1.08 1.09 0.11
F3 9.0 1.12 1.52 0.17
F4 9.0 1.15 1.80 0.20
F5 9.0 1.16 1.95 0.20

87
9

8.8

8.6
Force (kN)

F6
8.4

8.2 F7
8
F8
7.8 F9
7.6
0 0.5 1 1.5 2
Time (s)

Figure 23. Stress relaxation profiles of DCP/MCC compacts of different


proportions. (DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75)

Table 4. Viscoelastic parameters obtained for DCP/MCC compacts of different


proportions (DCP/MCC; F6: 100/0, F7: 75/25, F8: 50/50, F9: 25/75)

Formulation Compaction RQ VS1 VS2


force (kN) (kN/ms) (kN/ms)

F6 6.0 1.06 0.56 0.06


F7 6.0 1.12 1.10 0.08
F8 6.0 1.16 1.40 0.07
F9 6.0 1.20 1.59 0.10

F6 9.0 1.05 0.76 0.08


F7 9.0 1.11 1.37 0.10
F8 9.0 1.13 1.69 0.16
F9 9.0 1.16 1.92 0.17

88
APAP particles, majority of which were elastically deformed at the end of the

compaction cycle, reacted to the punch stop by an instantaneous elastic response by

trying to recover their original shape volume and had impeded particle movement

during dwell. Presence of MCC particles in the system brought about a more pliable

compact and allowed irreversible deformation at the beginning of the relaxation

process, thus storing much lower elasticity in the compact and hence reduced the

tendency of volume expansion during dwell. This favorably allowed further

deformational changes to the particles and facilitated movement of deformed

particles into interstitial spaces, resulting in increased plasticity.

While the introduction of plastically deforming component into the compact system

increased the overall structural relaxation in the APAP/MCC compacts, the degree

of stress relief was observed to be disproportional to MCC content, as clearly shown

by the stress relaxation profiles (Fig. 22). Initial addition of MCC into the compact

system (APAP/MCC, F2) led to a force decay of about 10 %. However, the extent

of structural relaxation was not consistent with subsequent increases in MCC

concentration; marginal differences in stress decay between APAP/MCC compacts

containing 75-100 % MCC (APAP/MCC, F4 and F5). The results could be explained

by porosity dependent secondary process that occurred in the compact

microstructure. Plastically deforming substances lead to remarkably lower porosities

in the compact than other materials with different deformation mechanisms, at

equivalent compaction loads. As a consequence of reduced internal porosity and

strain hardening characteristic of highly plastic MCC particles, resistance to

particulate displacement into interstitial spaces increased due to higher cohesiveness

and friction between particles [25]. Plastic flow was negatively affected as a result,

as reflected in the APAP/MCC stress profiles.

89
Replacement of elastically deforming APAP with brittle fracturing DCP particles,

exhibited a similar trend to the APAP/MCC systems, albeit with some differences.

In comparison to the APAP/MCC compacts (Tablet 3), presence of DCP particles

allowed a greater total stress decay (higher RQ values), at equivalent concentrations

of MCC particles in the system (Table 4). This was attributed to the reduced

deformational resistance of DCP particles under compaction load which imparted

lesser amount of elastic energy into the compact system than APAP particles. The

greater structural relaxation could have also been aided by the deformation behavior

of DCP which allowed particles to be in closer proximity with formation of a more

densely compacted coherent microstructure. This probably facilitated uniform

energy distribution throughout the compact, thus minimized localized high stress

concentrations and reduced the amount of elasticity in the compact. The enhanced

structural relaxation of DCP/MCC compacts was again observed in the viscoelastic

slopes of the first phase. VS1 values of DCP/MCC compacts (Table 4) were higher

than that of corresponding APAP/MCC compacts (Table 3). Interestingly, the

viscoelastic slopes for the subsequent phase (VS2), was noticeably lower when

compared to equivalent APAP/MCC systems, indicating reduced availability of

interstitial spaces possibly due to rapid movement of fractured particles during the

earlier phase coupled with increased interparticulate friction at the subsequent phase

in the DCP/MCC compact system.

The introduction of DCP to the compact system containing elastically deforming

APAP enhanced the structural relaxation of the compact; elastic energy decreased

via plastic flow (Fig. 24 and Table 5).

90
9

8.8
F10
8.6 F11
F12
Force (kN)

8.4

8.2

7.8

7.6
0 0.5 1 1.5 2
Time (s)

Figure 24. Stress relaxation profiles of APAP/DCP compacts of different


proportions. (APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75)

Table 5. Viscoelastic parameters obtained for APAP/DCP compacts of different


proportions (APAP/DCP; F10: 75/25, F11: 50/50, F12: 25/75)

Formulation Compaction RQ VS1 VS2


force (kN) (kN/ms) (kN/ms)

F10 6.0 1.04 0.37 0.05


F11 6.0 1.05 0.43 0.05
F12 6.0 1.06 0.54 0.04

F10 9.0 1.03 0.45 0.06


F11 9.0 1.04 0.64 0.07
F12 9.0 1.05 0.64 0.08

91
The total amount of stress relaxation increased marginally as the concentration of

DCP increased (Fig. 24). As expected, the extent of stress relief was much reduced

in comparison to the APAP/MCC and DCP/MCC compact systems. This was

attributed to the substantial elasticity in the system by the deformation of APAP and

fractured DCP particles during the consolidation stage. As the DCP particle size

decreased due to fragmentation, yield pressure and stress necessary to cause

deformation of the particles increased and contributed to high elasticity. In addition,

the internal area of contact due to reduction in particle size within the tablet

increased. This contributed to the cohesiveness and friction within the tablet

microstructure, restricting particulate movement during dwell. These combined

effects limited structural relaxation to a large extent in the APAP/DCP compact

system, in comparison to compact systems containing MCC particles.

4.1.4 Effect of formulation variables on tablet properties

The changes to the mechanical properties of tablets containing MCC (APAP/MCC,

DCP/MCC) with respect to formulation changes are depicted in Tables 6-9. Tablet

strength is a function of the bonding created by material deformation and bond

disruptive events that occur upon punch withdrawal. Tablet tensile strength

increased significantly with corresponding increases in MCC concentration in the

system, as observed in APAP/MCC (Tables 6-7) and DCP/MCC (Tables 8-9)

compacts. The relatively lower yield pressure of MCC particles allowed high degree

of plastic deformation which greatly increased the number and area of contact points

between particles within the tablet microstructure, a prerequisite for interparticulate

bonding. The capability of MCC to form strong hydrogen bonds also contributed

greatly towards the improved compact tensile strength [124].

92
Table 6. Effect of dwell time on stress decay and mechanical properties of
APAP/MCC tablets of different proportions, compacted at 6 kN (APAP/MCC; F1:
100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F1 6.0 0 1.00 0.08 (0.01) 0.24 (0.06)


100 1.00 0.06 (0.03) 0.11 (0.10)
500 1.02 0.08 (0.01) 0.20 (0.06)
1000 1.02 0.08 (0.01) 0.29 (0.06)
2000 1.03 0.08 (0.09) 0.26 (0.05)

F2 6.0 0 1.00 0.26 (0.01) 0.35 (0.11)


100 1.01 0.28 (0.03) 0.37 (0.10)
500 1.06 0.33 (0.02) 0.39 (0.03)
1000 1.08 0.31 (0.01) 0.36(0.10)
2000 1.10 0.36 (0.01) 0.33 (0.10)

F3 6.0 0 1.00 0.89 (0.06) 0.44 (0.07)


100 1.02 1.16 (0.09) 0.38 (0.03)
500 1.08 1.17 (0.05) 0.36 (0.20)
1000 1.11 1.26 (0.05) 0.42 (0.20)
2000 1.14 1.38 (0.03) 0.47 (0.08)

F4 6.0 0 1.00 2.00 (0.06) 0.42 (0.06)


100 1.02 2.16 (0.02) 0.48 (0.11)
500 1.10 2.13 (0.03) 0.59 (0.05)
1000 1.14 2.14 (0.1) 0.59 (0.09)
2000 1.18 2.55 (0.04) 0.50 (0.20)

F5 6.0 0 1.00 3.38 (0.10) 0.64 (0.16)


100 1.02 3.79 (0.06) 0.77 (0.07)
500 1.12 3.80 (0.07) 0.63 (0.07)
1000 1.16 3.68 (0.10) 0.58 (0.12)
2000 1.21 4.27 (0.20) 0.65 (0.30)

Parenthesized values represent the respective standard deviation.

93
Table 7. Effect of dwell time on stress decay and mechanical properties of
APAP/MCC tablets of different proportions, compacted at 9 kN (APAP/MCC; F1:
100/0, F2: 75/25, F3: 50/50, F4: 25/75, F5: 0/100)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F1 9.0 0 1.00 0.14 (0.02) 0.11 (0.07)


100 1.00 0.13 (0.01) 0.12 (0.01)
500 1.01 0.13 (0.01) 0.16 (0.02)
1000 1.02 0.13 (0.02) 0.11 (0.03)
2000 1.02 0.13 (0.09) 0.26 (0.01)

F2 9.0 0 1.00 0.52 (0.03) 0.18 (0.12)


100 1.01 0.51 (0.01) 0.27 (0.10)
500 1.05 0.53 (0.01) 0.30 (0.13)
1000 1.06 0.52 (0.03) 0.18 (0.06)
2000 1.08 0.53 (0.02) 0.26 (0.07)

F3 9.0 0 1.00 1.30 (0.09) 0.38 (0.13)


100 1.01 1.54 (0.10) 0.28 (0.20)
500 1.07 1.57 (0.01) 0.40 (0.16)
1000 1.09 1.56 (0.01) 0.40 (0.13)
2000 1.12 1.72 (0.11) 0.41 (0.13)

F4 9.0 0 1.00 3.09 (0.06) 0.38 (0.12)


100 1.02 3.28 (0.05) 0.33 (0.13)
500 1.08 3.22 (0.11) 0.38 (0.10)
1000 1.12 3.23 (0.05) 0.33 (0.04)
2000 1.15 3.60 (0.11) 0.38 (0.13)

F5 9.0 0 1.00 5.50 (0.08) 0.52 (0.11)


100 1.02 5.44 (0.09) 0.60 (0.05)
500 1.09 5.60 (0.06) 0.57 (0.12)
1000 1.13 5.58 (0.05) 0.60 (0.09)
2000 1.16 5.61 (0.06) 0.50 (0.03)

Parenthesized values represent the respective standard deviation.

94
Table 8. Effect of dwell time on stress decay and mechanical properties of
DCP/MCC tablets of different proportions, compacted at 6 kN (DCP/MCC; F6:
100/0, F7: 75/25, F8: 50/50, F9: 25/75)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F6 6 0 1.00 0.11 (0.01) 0.26 (0.04)


100 1.01 0.11 (0.01) 0.32 (0.06)
500 1.04 0.13 (0.01) 0.26 (0.11)
1000 1.05 0.10 (0.00) 0.28 (0.15)
2000 1.06 0.12 (0.01) 0.30 (0.09)

F7 6 0 1.00 0.41 (0.02) 0.43 (0.05)


100 1.01 0.48 (0.03) 0.37 (0.12)
500 1.07 0.51 (0.02) 0.38 (0.10)
1000 1.09 0.50 (0.02) 0.43 (0.23)
2000 1.12 0.52 (0.04) 0.42 (0.11)

F8 6 0 1.00 1.15 (0.11) 0.48 (0.22)


100 1.02 1.38 (0.03) 0.61 (0.20)
500 1.09 1.36 (0.05) 0.62 (0.18)
1000 1.13 1.37 (0.04) 0.70 (0.14)
2000 1.16 1.53 (0.04) 0.47 (0.20)

F9 6 0 1.00 2.29 (0.08) 0.53 (0.22)


100 1.02 2.50 (0.05) 0.66 (0.11)
500 1.10 2.54 (0.04) 0.61 (0.13)
1000 1.14 2.48 (0.10) 0.57 (0.13)
2000 1.20 2.76 (0.07) 0.68 (0.04)

Parenthesized values represent the respective standard deviation.

95
Table 9. Effect of dwell time on stress decay and mechanical properties of
DCP/MCC tablets of different proportions, compacted at 9 kN (DCP/MCC; F6:
100/0, F7: 75/25, F8: 50/50, F9: 25/75)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F6 9 0 1.00 0.20 (0.01) 0.25 (0.05)


100 1.01 0.19 (0.01) 0.24 (0.13)
500 1.03 0.19 (0.01) 0.25 (0.06)
1000 1.05 0.21 (0.01) 0.23 (0.03)
2000 1.05 0.21 (0.00) 0.22 (0.06)

F7 9 0 1.00 0.76 (0.05) 0.38 (0.08)


100 1.01 0.80 (0.06) 0.27 (0.13)
500 1.06 0.87 (0.08) 0.36 (0.16)
1000 1.08 0.86 (0.02) 0.39 (0.08)
2000 1.11 0.87 (0.04) 0.37 (0.11)

F8 9 0 1.00 2.04 (0.04) 0.40 (0.07)


100 1.01 2.15 (0.02) 0.47 (0.03)
500 1.07 2.12 (0.05) 0.67 (0.07)
1000 1.10 2.12 (0.03) 0.64 (0.13)
2000 1.13 2.41 (0.10) 0.44 (0.07)

F9 9 0 1.00 3.40 (0.09) 0.44 (0.14)


100 1.01 3.61 (0.10) 0.43 (0.14)
500 1.08 3.63 (0.04) 0.37 (0.03)
1000 1.12 3.60 (0.06) 0.51 (0.32)
2000 1.16 4.05 (0.09) 0.60 (0.06)

Parenthesized values represent the respective standard deviation.

96
Enhanced densification of powder bed led to a less restrained transmission of the

applied compaction energy and allowed greater dissipation of the built-up elastic

energy in the compact. This reduced the degree of elastic expansion of compact

during decompression which contributed to preserving the bonds formed during the

compaction cycle. Comparing the compacts formed by the two MCC containing

systems, the presence of a fragmenting material in the compact helped to form

mechanically stronger tablets at equivalent MCC concentrations, largely attributed

to the deformation mechanism of DCP particles. Fragmentation of DCP particles

with subsequent percolation of broken particles combined with occupation of voids

by plastically deformed MCC particles led to the optimal state of force utilization

and enhanced interparticulate bonding quality. Deformation of DCP particles

possibly facilitated greater dissipation of energy within the compact and reduced the

level of residual elasticity in the compact, as explained earlier.

Influence of DCP on tablet strength in the presence of APAP particles was rather

marginal, as shown by the tensile strength values in Tables 10-11. Initial introduction

of DCP (APAP/DCP, F10) to the system containing paracetamol only (F1) caused

noticeable increases in tablet strength (p < 0.05), attributed to creation of new

bonding surface and reduced elasticity conferred by the deformation of DCP

particles, as explained earlier. Interestingly, subsequent increases in the proportion

of DCP particles in the compact system (APAP/DCP, F11-F12) resulted in

insignificant changes to tablet tensile strength. This could be explained by the

differences in the bonding strength and utilization of compaction energy. Despite the

increase in effective contact area for bonding, the strength of the bonds formed was

relatively weak. The amount of elasticity from the deformation of APAP particles

was substantial, causing extensive rupture of the particle-particle bonds.

97
Table 10. Effect of dwell time on stress decay and mechanical properties of
APAP/DCP tablets of different proportions, compacted at 6 kN (APAP/DCP; F10:
75/25, F11: 50/50, F12: 25/75)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F10 6 0 1.00 0.10 (0.01) 0.30 (0.04))


100 1.01 0.09 (0.00) 0.21 (0.08)
500 1.02 0.10 (0.01) 0.29 (0.10)
1000 1.03 0.10 (0.01) 0.26 (0.10)
2000 1.04 0.10 (0.00) 0.26 (0.05)

F11 6 0 1.00 0.11 (0.00) 0.19 (0.09)


100 1.00 0.11 (0.00) 0.23 (0.07)
500 1.03 0.11 (0.01) 0.33 (0.09)
1000 1.04 0.11 (0.01) 0.23 (0.08)
2000 1.05 0.11 (0.01) 0.20 (0.09)

F12 6 0 1.00 0.11 (0.00) 0.27 (0.05)


100 1.01 0.10 (0.01) 0.19 (0.02)
500 1.03 0.10 (0.01) 0.29 (0.08)
1000 1.04 0.10 (0.00) 0.24 (0.12)
2000 1.06 0.10 (0.01) 0.32 (0.15)

Parenthesized values represent the respective standard deviation.

98
Table 11. Effect of dwell time on stress decay and mechanical properties of
APAP/DCP tablets of different proportions, compacted at 9 kN (APAP/DCP; F10:
75/25, F11: 50/50, F12: 25/75)

Formulation Compaction Dwell RQ Tensile VER


force time strength
(kN) (ms) (MPa) (%)

F10 9 0 1.00 0.15 (0.01) 0.14 (0.09)


100 1.00 0.16 (0.01) 0.29 (0.07)
500 1.02 0.16 (0.01) 0.17 (0.04)
1000 1.03 0.16 (0.01) 0.19 (0.07)
2000 1.03 0.15 (0.01) 0.18 (0.11)

F11 9 0 1.00 0.18 (0.01) 0.26 (0.11)


100 1.01 0.17 (0.00) 0.22 (0.09)
500 1.03 0.17 (0.01) 0.24 (0.09)
1000 1.04 0.17 (0.01) 0.28 (0.04)
2000 1.04 0.17 (0.01) 0.25 (0.11)

F12 9 0 1.00 0.17 (0.00) 0.29 (0.04)


100 1.01 0.18 (0.01) 0.25 (0.09)
500 1.03 0.17 (0.01) 0.30 (0.10)
1000 1.04 0.17 (0.00) 0.27 (0.03)
2000 1.05 0.17 (0.01) 0.27 (0.08)

Parenthesized values represent the respective standard deviation.

99
The effect of formulation variation on post compaction tablet strain recovery was

studied by comparing the percent viscoelastic recovery of tablets. The total tablet

elastic expansion may be defined as the sum of elastic recovery that occurs axially

immediately upon removal of the upper punch while still being constrained by the

die wall, and post compaction viscoelastic recovery that occurs during storage.

Tablets with MCC component showed distinct time-dependent viscoelastic recovery

(Tables 6-9). The percent viscoelastic recovery showed a clear trend in relation to

the proportion of MCC in the compact system; viscoelastic recovery generally

increased at higher concentrations of MCC in the system (p < 0.05). Plastic

deformation of MCC under applied constant pressure neutralized most of the stored

elastic energy and maintained high degree of bonding, preventing immediate release

of elastic energy upon punch withdrawal and facilitated viscoelastic strain recovery.

The use of a higher compaction force (9 kN) caused noticeable reductions in post-

compaction tablet expansion in both the APAP/MCC (Tables 6-7, F2-F5) and

DCP/MCC (Tables 8-9, F7-F9) compacts (p < 0.05). As the compaction load

increased, the degree of particulate deformation increased, with concomitant

increases in accumulated elasticity in the compact. However, the increase in stored

elastic energy was probably released as immediate tablet expansion (elastic

recovery) and resulted in lower energy store for viscoelastic tablet expansion [128].

In contrast to the post compaction recovery behavior of compacts containing MCC,

the physical specifications of DCP/APAP compacts remained relatively stable

throughout the storage period with consistent VER values observed (Table 10-11).

4.1.5 Effect of dwell on tablet properties

The effect of dwell time extension was observable in compact systems containing

MCC as one of the components; mechanical quality of tablets generally improved

100
with longer exposure to constant strain (dwell phase). The effect was more

noticeable in APAP/MCC (Tables 6-7) and DCP/MCC (Tables 8-9) compact

systems with 25-75 % MCC content (p < 0.05). Clearly, the increase in dwell time

caused significant improvements to the tablet mechanical quality by facilitating the

neutralization of stored elastic energy by plastic deformation during constant strain.

In consequence, reduced elastic energy contained in the compacts before the start of

punch withdrawal during the decompression phase ensured that strain recovery was

also reduced with less bond disruptive events and strengthened the tablets. These

results provided further evidence that the reaction of compacted solid to stress relief

was related to the densification mechanism of the solid.

Interestingly, it was noticed that increases in plastic flow along the time course of

constant strain did not lead to improvements in tablet strength in a proportional

manner, particularly at higher concentrations of MCC in the system (Tables 6-9, F2-

4, F7-9). Enhanced plastic flow (increasing RQ values) during dwell time 100-1000

ms very often did not cause significant changes to tablet tensile strength. In contrast,

appreciable increases in tablet strength was observed after the end of the second

phase (dwell time 2000 ms), emphasizing the time dependency of bond

consolidation in compacts showing viscoelastic character. Continuous particulate

deformation and movement during the initial phase of the relaxation process could

not provide sufficient time for building up of new interparticulate bonds. The

particulate system was in a state of flux, and without permanence, interparticulate

bonds could not be successfully formed. During the second phase of stress

relaxation, the rate of compact deformation slowed down with much reduced

movement of particles and this scenario provided conditions for greater

opportunities for fusion along with bond reinforcement. Tablet viscoelastic recovery

101
was observed to be largely independent of changes in dwell time, despite observable

changes in extent of plastic flow as indicated by the RQ values. Compact

composition and the applied compaction force were concluded to be more influential

factors in impacting viscoelastic recovery. Dwell time extension possibly impacted

immediate tablet recovery to a greater extent than time-dependent viscoelastic

recovery.

4.1.6 Summary

In this study, the influence of different deformation mechanisms of tableting

materials on plastic flow during the dwell phase and the consequent effect on tablet

mechanical properties were investigated. Plastic flow during dwell was observed in

the formulation systems investigated, at different magnitudes and rates of stress

decay. The process of stress relaxation within the compact during the time course of

constant strain showed two distinct phases. The first phase was related to particulate

deformation and movement into interstitial voids while the second phase was related

to reduced rate of plastic flow contributed by resistance to particulate movement due

to high internal friction between particles.

Presence of highly plastic material (MCC) in the compact system increased the

overall extent of force reduction and the rate of relaxation significantly. However,

the degree of stress relief during the different phases was observed to adopt a non-

linear relation with MCC content, attributed to porosity-dependent particulate

deformation. Extensive plastic deformation of MCC particles led to low internal

porosity and increased resistance to particulate movement which negatively affected

the rate of structural relaxation. The effect of dwell time extension was observable

in compact systems containing MCC; mechanical strength of tablets generally

increased with longer exposure to constant strain (dwell phase). Enhanced plastic

102
flow during the first phase of stress decay very often did not translate into tablet

strength enhancement. In contrast, appreciable increase in tablet strength was

observed after the end of second phase, emphasizing the time dependency of bond

consolidation in compacts showing viscoelastic character. The results demonstrated

the influence of formulation variation on dwell sensitivity and the limitations of

dwell extensions in different compact systems. The findings could be useful for

formulation optimization for the subsequent experiments involving dwell

modification.

103
4.2 Part 2. Investigation of the effect of punch head modification on compaction

parameters and tablet physical quality

4.2.1 Physical characteristics of tableting materials

Particle size and flow properties of the paracetamol-starch and lactose granules are

presented in Table 12. The differences in the median particle size for both types of

granules were marginal with paracetamol-starch granules exhibiting a wider size

distribution as shown by the larger span. Flow properties of both powders,

represented by the Hausner ratio and angle of repose could be considered

comparable; both powders possessed acceptable flowability suitable for high speed

tableting.

Table 12. Physical characteristics of paracetamol-starch and lactose granules

Tableting d50 Span Hausner AOR


material (µm) ratio (°)

Paracetamol -
starch 174.2 (3.80) 2.86 (0.02) 1.27 (0.01) 34.9 (1.6)
Lactose 168.3 (13.4) 1.76 (0.16) 1.28 (0.01) 33.5 (0.6)

Parenthesized values represent the respective standard deviation.

4.2.2 Compression behavior of tableting materials

Two types of tableting materials (paracetamol-starch and lactose) that differed in

terms of deformation mechanism were used in this study. The Heckel slope and the

extent of stress relaxation were used as non-specific test parameters to provide some

general description on the deformation mechanism of the tableting materials during

the compaction cycle and the period of constant strain as shown in Fig. 25.

104
A
2.05

1.9

1.75
ln (1/e)

1.6

1.45

1.3

1.15

1
0 20 40 60 80 100 120 140 160
Compression pressure (MPa)

B
9

8.9
Force (kN)

8.8

8.7

8.6

8.5

8.4
0 100 200 300 400 500
Time (ms)

Figure 25. Heckel plots (A) and stress relaxation curves (B) for lactose (  ) and
paracetamol-starch (  ) granules.

105
The Heckel plots for both materials showed an initial curve region, indicating

particle rearrangement and different degrees of brittle fracture, followed by long

quasi-linear portions pointing to prevailing plastic flow, in combination with elastic

deformation. While it is difficult to accurately distinguish between elastic and plastic

deformation components from the slopes of the linear portion of the Heckel plots,

the steepness of the slopes provides a relative comparison of the deformation nature

of the materials at high pressures. The steeper slope of the Heckel plot for lactose

(Fig. 25A) suggested a greater degree of plasticity and easier densification at the

applied pressures. In comparison to the tableting materials used in Part 1,

paracetamol-starch and lactose granules exhibited an overall deformation

mechanism intermediate of APAP and MCC particles (Fig. 21). The stress relaxation

curves (Fig. 25B) characterized the time-dependent viscoelastic behaviors of the

tableting materials. The rate of stress decay of the tested materials resembled MCC

containing compact systems used in Part 1. It was noted that, in comparison to

paracetamol-starch compacts, the force recorded for the lactose compacts decreased

more quickly under constant strain at the compaction forces used, particularly for

the initial 100 ms of the stress profile. This suggested that lactose compacts

presented a more pronounced viscoelastic character leading to greater structural

relaxation than paracetamol-starch compacts, attributed to the development of

reduced elasticity in the lactose compacts.

4.2.3 Influence of punch head configuration on compaction profile

The compaction parameters and profiles obtained for the different configurations of

punch head at various compaction forces are presented in Table 13 and Fig. 26. The

changes in the compaction cycle in relation to the punch head configurations were

found to be similar for the compaction of both lactose and paracetamol-starch

106
tablets, thus only the parameters and compaction profiles obtained from the

compaction of lactose tablets are presented. Results on the compaction parameters

were found to differ greatly between punch heads with and without a physical head

flat. Therefore, it was necessary for the results to be discussed in accordance to these

two classifications of punch head configurations.

Table 13. Effect of punch head configuration on compaction parameters obtained


using the larger compaction roll (D240)

Punch Compaction RFA Tcon Tdwell Tdec Ttotal


head force
configuration (kN) (daN/ms) (ms) (ms) (ms) (ms)

EHF 6.0 26.9 (0.5) 31.8 (0.2) 27.2 (0.2) 19.0 (0.1) 78.0 (0.0)
SHF 26.5 (0.6) 33.5 (0.5) 22.8 (0.1) 17.0 (0.8) 74.0 (0.4)
RHF 26.1 (0.3) 33.7 (0.4) 16.9 (0.2) 14.3 (1.8) 64.7 (0.9)
SRH 26.0 (0.4) 36.8 (0.7) 12.3 (0.2) 12.0 (0.5) 61.0 (1.0)
RRH 26.5 (0.4) 33.8 (0.2) 11.6 (0.2) 11.0 (0.0) 56.0 (0.4)

EHF 7.5 31.5 (0.2) 37.0 (0.1) 26.8 (0.2) 20.0 (1.0) 85.0 (1.3)
SHF 31.6 (0.4) 39.0 (0.1) 23.3 (0.3) 17.5 (0.5) 80.0 (0.0)
RHF 30.8 (0.5) 38.7 (0.5) 17.2 (0.1) 15.7 (0.4) 69.9 (0.8)
SRH 30.8 (0.2) 39.8 (1.3) 12.5 (0.4) 13.0 (0.4) 65.3 (0.9)
RRH 31.0 (0.4) 37.5 (0.5) 12.0 (0.3) 12.3 (0.2) 62.0 (0.0)

EHF 9.0 37.2 (0.5) 37.0 (0.1) 27.7 (0.4) 18.7 (0.4) 84.0 (0.0)
SHF 37.7 (0.5) 39.0 (1.0) 22.8 (0.1) 18.0 (0.0) 80.7 (1.0)
RHF 37.0 (0.6) 38.0 (0.1) 17.2 (0.1) 14.7 (0.4) 70.0 (0.0)
SRH 37.1 (0.2) 40.0 (0.1) 12.9 (0.1) 14.0 (0.0) 66.0 (0.8)
RRH 37.8 (0.4) 40.3 (0.3) 12.6 (0.2) 14.0 (0.0) 66.0 (0.6)

Parenthesized values represent the respective standard deviation.

107
A
10000

SHF
Compaction force (N)

8000
RHF

6000

4000

2000

0
0 10 20 30 40 50 60 70 80 90
Time (ms)
B
10000

SRH
Compaction force (N)

8000
RRH

6000

4000

2000

0
0 10 20 30 40 50 60 70 80 90
Time (ms)

Figure 26. Comparison of compaction profiles (6 and 9 kN) using the larger
compaction roll (D240) between SHF and RHF (A), SRH and RRH (B) head
configurations. (SHF: standard head flat, RHF: reduced head flat, SRH: standard
radius head, RRH: reduced radius head)

108
The RFA values for punches with head flat (EHF, SHF and RHF) were rather

similar, indicating marginal impact of the head radius curvature on the consolidation

phase (Table 13). A head curvature with a steeper slope (smaller radius) was

predicted to accelerate the consolidation phase and cause the dwell limit to be

reached earlier. However, despite the difference in the head radii between the SHF

(HR 16 mm) and RHF (HR 18 mm), the consolidation times for the two head

configurations did not vary significantly. EHF punches displayed shorter

consolidation time than SHF head configuration (p < 0.05) although both head

configurations shared common head radius, indicative of a quicker densification of

compact during the consolidation phase for the EHF punch head configuration.

Comparison of the compaction profiles of SHF and RHF head configurations (Fig.

26A) showed that the force-time curves at the early consolidation phase for the two

punches were almost identical, regardless of the compaction force used with peak

compaction force reached at similar time points. In contrast, alterations to the head

flat dimensions had a more pronounced effect on the compaction profiles. The dwell

time for the punches with a head flat increased proportionally with the head flat

diameter, with the EHF (HF 13.34 mm) configuration showing the longest dwell

time, followed by SHF (HF 10.41 mm) and RHF (HF 5.21 mm). This was expected,

as the time required for the compaction roll to transit across the flat portion of the

punch head has been reported to be independent of the required input force.

In contrast to the consolidation phase, the decompression phase was affected by the

punch head configuration. Decompression times were significantly different (p <

0.05), with the RHF head configuration having the most rapid release rate of

compaction energy during decompression and the EHF head configuration allowed

a more sustained release of compaction energy with a prolonged decompression time

109
(Table 13). However, the changes to the decompression phase was likely a result of

the dimensional modifications to the head flat, rather than changes to the head

curvature.

The punch head curvature was noticed to have a more pronounced effect on the

compaction parameters in the absence of a physical head flat. The results obtained

for the punches without head flat (SRH and RRH) showed slightly differing trends

(Table 13). While the RFA values were similar between them, there were noticeable

differences in the consolidation, dwell and decompression times. Due to the smaller

head radius, RRH (HR 27.43 mm) head configuration showed significantly shorter

consolidation, dwell and decompression durations than the SRH (HR 33.00 mm), at

lower compaction force levels of 6-7.5 kN. In consequence, total contact time

between the punch head and compaction roll for the RRH head configuration, was

lower by about 3-5 ms, largely contributed by the changes in the consolidation time.

However, the differences in the parameters were generally negligible at the higher

compaction force of 9 kN, with both head configurations sharing almost identical

compaction profiles as shown in Fig. 26B.

4.2.4. Influence of compaction roll configuration on compaction profile

The size of the compaction roll affects tablet compaction significantly, roll diameter

controls the rate and duration of force application as the punches diametrically

converge into the die. In view of this, the changes to compaction parameters and

profile due to reduction in compaction roll diameter were also recorded (Table 14

and Fig. 27). The use of a smaller roll caused the compaction profiles to undergo

significant changes, particularly at higher compaction forces. In comparison to the

compaction profiles from the larger roll, the smaller compaction roll accelerated

force application (increased RFA), reduced the consolidation and total contact times

110
for all the punch head configurations (p < 0.05). This considerably narrowed the

compaction cycle (Fig. 27A). The punch vertical displacement, which determines

the extent of punch penetration, was not affected by the change in roll dimensions.

Interestingly, amongst the head configurations with head flat, the compaction

profiles of EHF head configuration were least affected by the reduction in roll size

(Fig. 27B). This suggested that increasing the head flat diameter could possibly

counter the impact of roll size change. Punch head configurations with an extended

head flat could be useful to tablet manufacturers with the intention to preserve the

compaction profile when tablet presses fitted with a smaller compaction rolls are

utilized.

Table 14. Effect of punch head configuration on compaction parameters obtained


using the smaller compaction roll (D150)

Punch Compaction RFA Tcon Tdwell Tdec Ttotal


head force
configuration (kN) (daN/ms) (ms) (ms) (ms) (ms)

EHF 6.0 29.9 (0.4) 31.0 (0.4) 30.3 (0.4) 17.3 (0.9) 78.7 (0.5)
SHF 30.3 (0.2) 32.0 (0.0) 25.6 (0.1) 15.3 (0.4) 72.0 (0.8)
RHF 28.7 (0.8) 33.3 (0.8) 17.7 (0.2) 14.7 (0.4) 65.7 (0.9)
SRH 28.8 (0.2) 33.7 (1.3) 11.3 (0.2) 13.0 (0.4) 58.0 (1.6)
RRH 29.0 (0.2) 32.7 (0.8) 10.6 (0.1) 11.8 (0.2) 55.3 (0.9)

EHF 7.5 35.3 (0.2) 31.8 (0.7) 29.8 (0.7) 18.5 (0.5) 81.0 (1.0)
SHF 35.3 (0.2) 32.7 (0.9) 26.2 (0.1) 16.7 (1.3) 75.3 (0.9)
RHF 35.1 (0.6) 32.7 (0.3) 17.7 (0.5) 13.7 (1.2) 64.7 (0.4)
SRH 35.0 (0.4) 33.5 (0.5) 12.0 (0.0) 13.5 (0.5) 59.0 (1.0)
RRH 35.5 (0.3) 35.5 (1.0) 11.3 (0.2) 12.3 (0.2) 59.0 (1.0)

EHF 9.0 43.8 (0.8) 30.7 (0.4) 30.7 (0.5) 17.7 (1.2) 82.0 (1.1)
SHF 44.2 (0.5) 31.9 (0.5) 26.5 (0.1) 16.7 (0.4) 75.0 (0.3)
RHF 43.6 (0.4) 32.3 (0.5) 17.7 (0.4) 16.0 (0.8) 65.3 (0.9)
SRH 42.0 (0.3) 34.5 (1.0) 11.9 (0.1) 13.8 (0.7) 60.0 (0.0)
RRH 42.8 (0.2) 35.0 (1.1) 11.7 (0.1) 14.0 (0.4) 60.7 (0.9)

Parenthesized values represent the respective standard deviation.

111
A
10000
Compression force (N)

8000

6000

4000

2000

0
0 10 20 30 40 50 60 70 80 90
Time (ms)
B
10000

8000
Compaction force (N)

6000

4000

2000

0
0 10 20 30 40 50 60 70 80 90
Time (ms)

Figure 27. Comparison of compaction profiles (6 and 9 kN) between compaction


rolls of diameters 240 (─) and 150 (---) mm using the SHF (A) and EHF (B) head
configurations. (EHF: extended head flat, SHF: standard head flat)

112
Nonetheless, the consolidation phase was clearly independent of the changes in the

punch head radius even with a smaller roll as evidenced by the largely indifferent

RFA values and consolidation times of the SHF and RHF head configurations (Table

14). The results also indicated that the effects on dwell and decompression times

differed between the classes of punch head configurations. The values for dwell time

were found to increase significantly for the head configurations with a flattened head

geometry at all force levels studied; Tdwell increased by approximately 3 ms and 0.5

ms for the EHF/SHF and RHF head configurations respectively (p < 0.05) (Tables

13-14). Despite the reduction in roll size, the mounting surface of the compaction

rolls of different sizes onto the head flat was maintained thus theoretically caused

minimal changes in the time for the rolls to transit across the flatten surface of the

punch head. However, the reduced roll size caused an accelerated rise in the

compaction force during consolidation phase, causing the force limit for dwell phase

to be reached earlier and this resulted in increased dwell time (Table 14). In

comparison, SRH and RRH punch head configurations experienced a small

reduction in dwell times of about 1 ms, indicating the effects of not possessing a

physical flat top on the punch head; lack of head flat caused the roll to traverse at a

faster rate across the head surface. The EHF and SHF head configurations

experienced a reduction in the decompression time, similar to the effect on

consolidation phase, while the other head configurations were generally unaffected

when a smaller compaction roll was used.

The findings suggested that the head flat and compaction roll specifications are

influential factors that could cause observable changes to the compaction process.

The head radius of the punch head was predicted to impact the loading and unloading

rate of the consolidation and decompression phase respectively; larger head radius

113
should result in lower loading and unloading rates. The results, however, suggested

that the effect of the head radius on the compaction process was generally marginal,

if any. This is possibly due to the greater impact of the roll curvature in translating

the horizontal punch movement into axial compaction force than the much limited

punch head radius. Nonetheless, the punch radius curvature could play an important

role in increasing the tooling life of the punches. The curved surface of the punch

head makes initial contact with the compaction roll during rotary movement. A

larger head radius (reduced curvature) could possibly aid in reducing the initial stress

exerted on the punch head upon contact and potentially reduce the tendency to head

fracturing.

4.2.5 Influence of punch head configuration on tablet properties

The changes to the mechanical properties of lactose tablets with respect to the

different punch head configurations are depicted in Fig. 28. The results showed that

there was good agreement between the differences observed in the tablet mechanical

properties and the changes in the compaction profiles. Tablet tensile strength showed

a positive relationship with dwell time. The tablets compacted using the EHF head

configuration possessed the highest tensile strength, followed by the SHF and RHF

head configurations, while the tablets prepared using the SRH and RRH punches

displayed much lower tensile strengths (Fig. 28A) at equivalent compaction forces

(p < 0.05). It was observed that tabletability decreased in accordance with the change

of head configuration from EHF (Cp = 0.097) to RRH (Cp = 0.080), suggesting

reduced conversion of compaction energy into plastic energy. The amount of plastic

flow due to stress relaxation during the period of constant strain was evaluated by

the calculation of the ratio of spatial areas under the force-time profiles.

114
A
0.6

Tensile strength (MPa)


0.5

0.4

0.3

0.2

RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF
0.1
6 7.5 9
Compaction force (kN)
B
1.4

1.2

1
AQ

0.8

0.6
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF

0.4
6 7.5 9
Compaction force (kN)

Figure 28. Effect of punch head configuration on lactose tablet tensile strength (A)
and area quotient (B) at the experimental compaction forces for compaction roll of
diameter 240 mm (D240). (EHF: extended head flat, SHF: standard head flat, RHF:
reduced head flat, SRH: standard radius head, RRH: reduced radius head)

115
Apparently, the changes in dwell time had contributed to differences in the stress

relaxation process for the compacts that were prepared by the different punch head

configurations, as evidenced by the changes in AQ values (Fig. 28B). The ratio of

partial areas after and before the middle of dwell time indicates the extent of

irreversible deformation of the material under load at constant punch displacement.

During the dwell phase, the volume is nearly constant, resulting in a decrease in

force at the punch face in contact with the tablet surface, as a result of stress

relaxation. The AQ values are interpreted inversely as the relaxation quotient (RQ)

in Part 1, lower values indicate greater conversion of elasticity to plastic flow.

Materials with greater viscoelastic character would deform more readily during the

dwell phase causing reduction in AQ values. Stress relaxation during dwell

decreased at higher compaction forces, a trend similar to data on relaxation quotient

in Part 1. The extent of stress relaxation is largely influenced by the amount of

elasticity in the compact at the beginning of the dwell phase; greater elastic

deformation with increasing compaction load at the end of the consolidation phase

led to reduced plastic yielding during dwell phase.

By comparing the AQ values for the punch head configurations with a head flat (Fig.

28B), it could be seen that the EHF head configuration allowed the greatest amount

of plastic flow (lowest AQ value) during the dwell phase followed by SHF and RHF

head configurations. Clearly, the increase in dwell time by even a small extent

facilitated the conversion of stored elastic energy to plastic deformation during

constant strain and led to significant improvements to the tablet mechanical quality.

In consequence, reduced elastic energy contained in the compacts before the start of

punch withdrawal during the decompression phase ensured that axial strain recovery

was also reduced; decreasing tablet elastic expansion with increasing dwell time

116
(Table 15). The reduction in decompression time from EHF to RHF head

configurations (Tables 13-14) could also have contributed to tablet elastic recovery

by allowing accelerated release of stored elastic energy, facilitating porosity

expansion. Porosity expansion during decompression phase causes disruption of

interparticulate bonds formed during the consolidation and dwell phases, with

formation of microcracks which, if propagated, could translate to lower compact

mechanical strength during the diametrical compression test [14]. The plasticizing

action of increased periods of stress relaxation coupled with slower release of stored

energy reduced the porosity of the resultant tablets upon ejection, consequently

increased interparticle contact area. Thus, the lower porosity of tablets produced

using EHF head configuration (Table 15) suggested more interparticle bonds

preserved as compared to tablets produced from SHF and RHF punches, and

translated to better tablet mechanical strengths. In contrast, higher resistance to stress

relaxation was noted for the SRH and RRH head configurations, as evidenced by the

substantial increase in the AQ values (Fig. 28B). Despite the slightly longer dwell

time and lower strain rate by the SRH head configuration at lower force levels (Table

13), differences in the physical quality of the tablets and extent of plastic flow during

dwell for the SRH and RRH punches were, however, insignificant. The sharp

increase in AQ values could be attributed to the lack of a physical head flat that

resulted in the compacts experiencing extremely shortened periods of constant

strain, if any, during the compaction cycle. Thus, the compacts experienced a state

of continual deformation during compaction, possibly with constant increases in

elasticity without the opportunity for stress relaxation. These unfavorable conditions

possibly promoted the build-up of an abundance of elastically deformed sites within

the compacts.

117
Table 15. Effect of punch head configuration on lactose tablet mechanical properties
at the experimental compaction forces for compaction roll of diameter 240 mm
(D240) (EHF: extended head flat, SHF: standard head flat, RHF: reduced head flat,
SRH: standard radius head, RRH: reduced radius head)

Punch Compaction ER Porosity Ejection


head force force
configuration (kN) (%) (%) (N)

EHF 6.0 5.9 (0.2) 18.96 (0.11) 242 (10)


SHF 7.7 (0.3) 19.17 (0.01) 233 (13)
RHF 8.4 (0.2) 19.40 (0.05) 182 (11)
SRH 9.2 (0.1) 20.00 (0.09) 181 (12)
RRH 9.3 (0.1) 19.93 (0.07) 178 (10)

EHF 7.5 6.4 (0.5) 17.76 (0.08) 333 (10)


SHF 7.9 (0.3) 17.95 (0.11) 329 (12)
RHF 8.7 (0.2) 18.33 (0.08) 317 (12)
SRH 9.6 (0.2) 18.77 (0.02) 288 (7)
RRH 9.4 (0.2) 18.70 (0.16) 285 (9)

EHF 9.0 8.1 (0.4) 16.00 (0.10) 403 (27)


SHF 9.6 (0.3) 16.19 (0.17) 418 (14)
RHF 10.2 (0.1) 16.51 (0.03) 344 (16)
SRH 11.5 (0.3) 16.60 (0.09) 348 (18)
RRH 11.9 (0.4) 16.80 (0.11) 340 (14)

Parenthesized values represent the respective standard deviation.

Rapid elastic recovery induced widespread microcracks with the consequence of

reduced tablet tensile strengths, as shown by the increase in the elastic recovery and

porosity values. The elastic response of the powder during compaction in accordance

with changing head configurations from EHF to RRH punches was also reflected in

the ejection force recorded (Table 15). Compacts prepared from punches with head

flat generally exhibited higher ejection force values than the punches with radial

head configurations (p < 0.05). Ejection force is a function of residual die wall

pressure and friction at the interphase between the compact and die wall surface.

Increased strain recovery for tablets produced with radial head punches caused

compacts to have greater axial recovery and by the Poisson effect, contracted

118
radially to a greater extent, thus leading to lower residual die wall pressure and

correspondingly, a lower ejection force [138-140].

The influence of the punch head configuration on the mechanical properties of

tablets compacted using the more elastically deforming paracetamol-starch granules

is presented in Fig. 29 and Table 16. Paracetamol-starch tablets generally

experienced reduced tensile strength (Fig. 29A) and greater elastic recovery (Table

16) compared to corresponding lactose tablets at equivalent compaction forces. This

was attributed to the excessive elastic deformation of paracetamol-starch granules

as supported by the Heckel plots and reduced tendency towards plastic flow during

dwell as illustrated by the stress relaxation curves (Fig. 25). Nonetheless, the benefit

of dwell time extension on paracetamol-starch tablet mechanical properties was

evident (Fig. 29A), a similar trend as exhibited by the lactose tablets (Fig. 28A),

albeit with some limitations. At compaction forces up to 7.5 kN, the tensile strength

predictably decreased from EHF to RRH, with negligible differences observed

between the tablets produced from SRH and RRH head configurations. However, it

was noticed that the tablets produced from EHF and SHF punches were equivalent

in terms of physical strength while the differences in tensile strength of tablets from

the other punches were insignificant at 9 kN. Tablet failure in the form of capping

was observed in the paracetamol-starch tablets. Analysis of the results on capping

index showed a similar outcome with capping tendency generally decreasing with

increasing dwell time (Table 16). The improved resistance to tablet failure with

increasing dwell time was observed again which reinforced the earlier discussion

that the reduction of stored elastic energy, indicated by the percent elastic recovery,

had been advantageous. Enhanced plastic yield led to decreased tablet porosity and

increased inter-particle contacts.

119
A
0.6
Tensile strength (MPa)

0.5

0.4

0.3

0.2
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF
0.1
6 7.5 9
Compaction force (kN)
B
1.4

1.2

1
AQ

0.8

0.6
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
SHF

EHF
SHF

EHF
SHF
EHF

0.4
6 7.5 9
Compaction force (kN)

Figure 29. Effect of punch head configuration on paracetamol-starch tablet tensile


strength (A) and area quotient (B) at the experimental compaction forces for
compaction roll of diameter 240 mm (D240). (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head, RRH:
reduced radius head)

120
Table 16. Effect of punch head configuration on paracetamol-starch tablet
mechanical properties at the experimental compaction forces for compaction roll of
diameter 240 mm (D240) (EHF: extended head flat, SHF: standard head flat, RHF:
reduced head flat, SRH: standard radius head, RRH: reduced radius head)

Punch Compaction ER Capping Porosity Ejection


head force index force
configuration (kN) (%) (%) (N)

EHF 6.0 11.3 (0.4) 0.8 17.66 (0.04) 94 (5)


SHF 12.6 (0.2) 1.1 18.18 (0.15) 97 (3)
RHF 12.9 (0.3) 1.3 18.65 (0.14) 85 (4)
SRH 13.4 (0.4) 1.8 19.00 (0.13) 87 (4)
RRH 13.8 (0.3) 1.7 19.16 (0.23) 89 (4)

EHF 7.5 13.0 (0.2) 1.1 16.30 (0.10) 109 (2)


SHF 13.9 (0.3) 1.5 16.70 (0.13) 108 (3)
RHF 14.7 (0.4) 1.8 17.96 (0.07) 107 (5)
SRH 15.4 (0.5) 2.0 17.33 (0.13) 106 (3)
RRH 16.0 (0.1) 2.2 17.33 (0.06) 103 (7)

EHF 9.0 14.4 (0.6) 2.0 15.84 (0.18) 118 (10)


SHF 16.4 (0.4) 2.1 16.18 (0.22) 116 (3)
RHF 17.3 (0.5) 2.7 16.38 (0.26) 114 (11)
SRH 17.5 (0.4) 2.6 16.82 (0.16) 101 (2)
RRH 17.9 (0.7) 2.7 16.83 (0.08) 104 (2)

Parenthesized values represent the respective standard deviation.

At higher compaction forces, the effectiveness of dwell extension on capping

tendency was reduced. While tablets produced using the EHF punches showed the

least capping tendency, the differences between the capping indices of EHF and SHF

narrowed as the compaction force increased while the differences between SHF and

RHF head configurations widened. At compaction force 9 kN, the punches could be

grouped into two distinct groups by the mechanical quality of the tablets produced -

EHF/SHF and RHF/SRH/RRH. These results suggested that dwell extension may

not be as beneficial, particularly at higher compaction forces and for materials that

undergo high degree of elastic deformation. The results obtained were similar to the

outcome observed in Part 1, in which increases in dwell time did not contribute to

121
the physical strength of the tablets, attributed to insufficient time for bond

reinforcement after plastic yielding during dwell.

The AQ values for paracetamol-starch tablets (Fig. 29B) were higher than that of

lactose tablets (Fig. 28B) at equivalent compaction forces, indicating greater

resistance to plastic flow attributed to the presence of high elasticity at the end of the

consolidation phase. The AQ values as expectedly adhered to a positive relation with

compaction force, AQ increased at higher force levels. In a trend parallel to that of

lactose tablets, plastic flow during dwell was greatest in the tablets compacted using

EHF punches; AQ values increased from EHF to RHF with a significant rise in

values for the SRH and RRH punch head configurations at all compaction forces. In

relation to the tablet mechanical property data, the enhanced stress relaxation with

increased dwell time was not entirely reflected by the tensile strength and capping

index data. Tablets compacted using EHF head configuration experienced greater

stress relaxation and lower elastic expansion than tablets from the SHF head

configuration at all forces.

The enhanced plastic yielding, however, did not lead to an improved mechanical

quality of the tablets at the highest compaction force of 9 kN. A similar observation

was also made when tablets from RHF and SRH/RRH head configurations were

compared. These results could again be related to the data in Part 1, in which

consistent increases in overall stress decay experienced by the compact with

increasing dwell time was not translated into enhancement in the tablet strength.

4.2.6 Finite element analysis

FEM is an effective tool to simulate compaction mechanics of powder. In this study,

FEM was utilized to investigate the transmission of the applied load by the bevel

shaped punches and the reaction of the compacted solid to dwell time changes. The

122
mechanical behavior of the powder was modeled using the Drucker-Prager Cap

model, using material properties (β = 72.41, R = 0.622, a = 0.02, d = 0.5, Pb = 69.6)

calibrated from experiments with the instrumented die described in Section 3.2.9,

using the paracetamol-starch granules. A constant material property model was used,

as it consumed a smaller amount of computational time [108]. Symmetric

compaction was simulated, upper and lower punches moved at the same time and in

opposite directions, with compaction force fixed at 9 kN.

The stress distribution within the powder compact was produced by contour

mapping of the stress experienced by every element in the simulated compact. The

relative stress distribution at the end of different periods of dwell representative of

head flat change, before the decompression process, is shown in Fig. 30. In all cases,

the load of compaction imposed by the punches was not completely transmitted

within the compact and led to development of stress gradients throughout the height

of the simulated tablet. This heterogeneity of the stress distribution was principally

due to the friction between the tools and powder, and the influence of punch face

surface on the resolution of load through the compact. The effective stress at the top

half of the compact was observed to be generally lower than the bottom half of the

compressed powder bed indicating that the powder near the top generally relaxed

more than that at the bottom half of the compact, consistent with the results reported

in existing studies [98, 103].

Examination of the stress distribution revealed development of maximum stresses in

the regions adjacent to the angled surface of the punch face. The degree of stress was

almost constant along the regions in contact with the flat portion of the punch face.

123
A B

C D

S, mises (relative)
E
9.7
8.9
8.1
7.3
6.5
5.7
5.0
4.2
3.4
2.5
1.8
1.0

Figure 30. Stress distribution contours of simulated paracetamol-starch tablets


compacted at dwell times representative of RRH (A), SRH (B), RHF (C), SHF (D)
and EHF (E) punch head configurations.

Stress intensity rose when approaching the angled surface of the punch face followed

by the development of an intensive shear band that ran from the top edge towards

the mid-center of the compact. Formation of shear bands has been suggested to

facilitate crack propagation during the decompression phase during which a part of

the tablet is free to dilate axially and other parts affected by the die wall effect,

contributing to capping [109]. This clearly showed the great influence of the punch

form on stress distribution in convex tablets.

124
Stress distribution within the compact, which corresponded to the development of

elastic energy, was observed to react significantly to changes in dwell duration.

Strain deformation during the consolidation phase led to the accumulation of stored

energy. The introduction of shortened dwell phases, representative of RRH and SRH

punch head configurations, allowed certain regions of the compacts to undergo

minimal stress relief (Fig. 30A-B). Extensive stress gradients continued to exist

between the central regions and the regions adjacent to the punch face. Further

extension of the dwell phase led to significant change in the stress distribution and

promoted reduction in the intensity of the stress gradient and high stress regions,

particularly in the central regions of the compact. This indicated different rates of

stress relief in different regions of the compact. Plastic flow under constant strain in

the compact was observed to initiate from the central regions of the compact, with

the shear band from the peripheral regions still preserved. The total compact stress

at punch face level was distributed more equally at extended dwell times, thus

reduced local shear intensity experienced by the particles in the compact and reduced

stress gradients between the different regions in the compacts. At the end of the

dwell phase allowed by the EHF head geometry, it could be seen that the regions of

high accumulated stress were localized at the regions opposite the beveled surface

of the punch face. Clearly, the extension of time for viscoelastic deformation reduced

the overall amount of stored energy in the compact.

While the numerical data proved valuable in studying the stress state in the

compacts, the presented simulation model was unable to differentiate between the

different deformation mechanisms e.g. plastic and brittle fracturing, of the tableting

materials. More localized bands may occur in some types of materials which could

not be predicted by the simulation. The powder mass in the simulation was treated

125
as a continuum mass with uniform bulk density throughout the compact, which is

often not the case after die filling, particularly during high speed tableting. Powder

compaction studies have shown that material properties of powders can vary

substantially during the compaction process itself due to changes in the relative

densities within the powder bed. Numerical predictions of force transmission could

also be enhanced by considering the inter-particle cohesive strength, which is not

straight forward and further investigations will be required to the implication of such

particle properties used in the simulation system [109]. Nonetheless, the presented

FEM model was sufficient for the investigation of dwell phase compact deformation.

4.2.7 Influence of compaction roll configuration on tablet properties

The changes to the mechanical properties of lactose tablets with respect to the

smaller compaction roll size are depicted in Fig. 31 and Table 17. In comparison to

corresponding tablets compacted using the bigger roll (Fig. 28 and Table 15) at

equivalent compaction forces, the use of the smaller compaction roll caused the

tensile strengths (Fig. 31A) and elastic recovery (Table 17) of the tablets to decrease

and increase significantly (p < 0.05), respectively, indicative of increased elastic

content within the compact. Increased strain recovery resulted in greater tablet

porosity expansion (bond disruptive event) and consequently led to mechanically

weaker resultant tablets. The effect of the change in compaction roll size was also

noticed in the ejection forces; reduction in roll size led to reduced ejection force,

explained by the increased post compaction axial recovery during decompression

phase. Nonetheless, the overall benefit of dwell time extension even with reduced

roll dimension was again evident.

126
A 0.6

Tensile strength (MPa)


0.5

0.4

0.3

0.2
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF
0.1
6 7.5 9
Compaction force (kN)
B 1.4

1.2
AQ

0.8

0.6
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF

0.4
6 7.5 9
Compaction force (kN)

Figure 31. Effect of punch head configuration on lactose tablet tensile strength (A)
and area quotient (B) at the experimental compaction forces for compaction roll of
diameter 150 mm (D150). (EHF: extended head flat, SHF: standard head flat, RHF:
reduced head flat, SRH: standard radius head, RRH: reduced radius head)

127
Table 17. Effect of punch head configuration on lactose tablet mechanical properties
at the experimental compaction forces for compaction roll of diameter 150 mm
(D150) (EHF: extended head flat, SHF: standard head flat, RHF: reduced head flat,
SRH: standard radius head, RRH: reduced radius head)

Punch Compaction ER Porosity Ejection


head force force
configuration (kN) (%) (%) (N)

EHF 6.0 6.1 (0.1) 20.47 (0.08) 175 (12)


SHF 8.1 (0.2) 20.58 (0.09) 183 (7)
RHF 8.6 (0.1) 20.89 (0.11) 147 (5)
SRH 8.9 (0.1) 21.13 (0.12) 130 (10)
RRH 9.1 (0.3) 21.32 (0.11) 141 (17)

EHF 7.5 7.2 (0.1) 18.18 (0.07) 230 (11)


SHF 9.4 (0.2) 18.39 (0.08) 241 (7)
RHF 9.8 (0.2) 19.10 (0.16) 222 (9)
SRH 10.3 (0.1) 18.94 (0.09) 206 (5)
RRH 10.9 (0.3) 19.06 (0.06) 209 (10)

EHF 9.0 9.8 (0.2) 16.92 (0.09) 316 (6)


SHF 11.9 (0.3) 17.12 (0.03) 327 (12)
RHF 12.1 (0.4) 17.21 (0.07) 280 (7)
SRH 13.1 (0.3) 17.70 (0.16) 236 (15)
RRH 13.3 (0.4) 17.84 (0.10) 242 (15)

Parenthesized values represent the respective standard deviation.

Tablet mechanical quality improved from RRH to EHF head configuration at all

experimented force levels with EHF punches still possessing the highest tabletability

properties. While the shift to a smaller roll size led to weaker tablets, the impact on

compact deformation during dwell phase was observed to be more positive (Fig.

31B). The AQ values for the punches with flat head geometries decreased noticeably

due to increased dwell time (p < 0.05). This suggested that the particles underwent

enhanced stress relaxation within the confines of the die cavity when a smaller roll

was used for compaction purpose. EHF and SHF punches experienced increased

plastic yielding as evidenced by the greater reduction in the AQ values than RHF

punch head configuration due to the greater increase in dwell time for EHF and SHF

128
punches than RHF punches (Table. 17). The contrasting outcome on tablet

mechanical properties despite greater plasticity development during dwell phase is

attributed to the changes in the densification during consolidation phase. The use of

a smaller roll for tablet compaction caused the strain rate to increase dramatically

during the consolidation phase, increased elastic content in the materials under stress

and contributed to increased air entrapment.

By comparing the data obtained from lactose and paracetamol-starch tablets, it could

be seen that the use of a smaller compaction roll altered the mechanical properties

of the paracetamol-starch tablets similarly but with some differences. The

mechanical quality of the paracetamol-starch tablets, as expected, deteriorated with

the use of a smaller roll and this could be assigned to the rapid elastic expansion

upon punch withdrawal (Fig. 32A). Interestingly, the positive impact of dwell time

extension on tablet mechanical strength also declined with a smaller roll.

While the tablets compressed using EHF punches continued to exhibit the highest

mechanical quality, tensile strength values of the tablets produced from SHF and

RHF punches were statistically indifferent but of higher values than those compacted

using SRH/RRH punches at compaction forces of 6-7.5 kN. At compaction force 9

kN, tensile strength of tablets produced using the EHF and SHF punches were

equivalent, despite observable differences in the extent of elastic recovery between

the two punches (Table 18), and only marginal differences existed between the

tablets from RHF and SRH/RRH punches, a similar trend observed in paracetamol-

starch tablets compacted using the bigger roll.

129
A 0.4
Tesnile strength (MPa)

0.3

0.2 RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF
0.1
6 7.5 9
Compaction force (kN)

B 1.4

1.2

1
AQ

0.8

0.6
RRH

RRH

RRH
SRH

SRH

SRH
RHF

RHF

RHF
EHF
SHF

EHF
SHF

EHF
SHF

0.4
6 7.5 9
Compaction force (kN)

Figure 32. Effect of punch head configuration on paracetamol-starch tablet tensile


strength (A) and area quotient (B) at the experimental compaction forces for
compaction roll of diameter 150 mm (D150). (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat, SRH: standard radius head, RRH:
reduced radius head)

130
Table 18. Effect of punch head configuration on paracetamol-starch tablet
mechanical properties at the experimental compaction forces for compaction roll of
diameter 150 mm (D150). (EHF: extended head flat, SHF: standard head flat, RHF:
reduced head flat, SRH: standard radius head, RRH: reduced radius head)

Punch Compaction ER Capping Porosity Ejection


head force index force
configuration (kN) (%) (%) (N)

EHF 6.0 11.4 (0.3) 1.2 18.30 (0.24) 66 (6)


SHF 13.5 (0.2) 1.3 18.94 (0.13) 64 (10)
RHF 13.6 (0.3) 1.8 18.99 (0.11) 76 (13)
SRH 14.3 (0.4) 2.2 19.67 (0.18) 64 (8)
RRH 14.1 (0.3) 2.2 19.41 (0.08) 67 (11)

EHF 7.5 13.5 (0.6) 1.5 17.20 (0.26) 108 (3)


SHF 15.6 (0.2) 1.8 17.36 (0.25) 86 (12)
RHF 16.5 (0.3) 2.1 17.98 (0.20) 82 (10)
SRH 17.8 (0.8) 2.5 18.24 (0.26) 68 (8)
RRH 17.1 (0.5) 2.7 18.25 (0.17) 69 (9)

EHF 9.0 16.5 (0.2) 2.2 16.28 (0.27) 111 (9)


SHF 18.1 (0.3) 2.3 16.48 (0.08) 97 (13)
RHF 20.0 (0.3) 2.8 16.73 (0.15) 89 (10)
SRH 19.5 (0.1) 2.9 17.00 (0.05) 87 (4)
RRH 19.8 (0.1) 3.0 17.08 (0.10) 90 (2)

Parenthesized values represent the respective standard deviation.

Capping indices increased for all punch configurations at equivalent compaction

forces with the smaller roll and this can be attributed to increased elasticity (Table.

18). However, the inter-punch trend for capping tendency differed from the data

obtained using the bigger compaction roll. The differences in capping indices for the

tablets from EHF and SHF were generally smaller when compared to the capping

index values obtained from the larger roll. A distinct difference in capping index

values between RHF and SRH/RRH punch head configurations existed until

compaction force 7.5 kN. At 9 kN, the differences in capping index between RHF

and SRH/RRH punches could be considered marginal. This consistent plateauing of

the tablet mechanical properties at higher force levels of the paracetamol-starch

131
tablets compacted by either of the roll sizes suggested the presence of different levels

of elasticity thresholds in the compact system under the experimented tableting

conditions.

4.2.8 Effect of tableting speed

Manipulation of the rotary tableting speed affects the time of contact between the

compaction roll and punch head, impacting the compaction cycle as a result.

Tableting speed and dwell time follow an inverse relation; slowing the turret speed

increases the dwell time as it takes more time for the roll to transit across the head

flat. In view of this, the relationship between tableting speed, punch head geometry

and the resultant tablet properties was also evaluated (Tables 19-20). Lactose tablets

were produced using the larger compaction roll with punch configurations having

head flats at a higher turret speed of 35 rpm (47 m/min). The increased tableting

speed narrowed the differences in compaction profiles, with increased RFA

accompanied by a drastic reduction in dwell time by about 30 % (Table 19). The

force-time curves during dwell time became more symmetrical, resulting in

significant increases in the AQ values indicative of lesser overall stress relaxation

during dwell time (p < 0.05) (Table 20). As a result, the tablets compressed using

the higher turret speed displayed reduced tensile strengths. Contrastingly, tablet

elastic recovery decreased at equivalent compaction forces. This could be explained

by the higher compact porosities, shown by the increased set compaction thickness

values obtained at the higher tableting speed. As the punch vertical velocity

increased, the rate of load application provided less time for powder consolidation

and air exhaustion, resulting in compacts of increased porosities [78, 81]. These

effects reduced the degree of interparticulate bonding in the tablets and therefore,

tabletability decreased at higher tableting speeds.

132
Table 19. Effect of punch head configuration on compaction parameters of lactose
tablets compacted at turret speed 35 rpm (47 m/min) (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat)

Punch Compaction RFA Tcon Tdwell Tdec Ttotal


head force
configuration (kN) (daN/ms) (ms) (ms) (ms) (ms)

EHF 6.0 36.2 (0.2) 24.1 (0.1) 18.5 (0.1) 14.7 (0.5) 58.0 (1.6)
SHF 35.8 (0.2) 25.0 (0.4) 15.7 (0.2) 13.2 (0.6) 53.3 (0.9)
RHF 36.1 (0.3) 26.1 (0.6) 12.1 (0.1) 10.3 (0.9) 48.0 (0.6)

EHF 9.0 49.5 (0.6) 26.1 (0.2) 19.2 (0.2) 14.3 (0.2) 60.0 (0.1)
SHF 50.9 (0.3) 27.2 (1.2) 15.5 (0.1) 13.2 (0.2) 56.0 (1.6)
RHF 51.8 (0.1) 26.3 (0.6) 11.6 (0.1) 10.8 (0.6) 49.6 (0.9)

Parenthesized values represent the respective standard deviation.

Table 20. Effect of punch head configuration on mechanical properties of lactose


tablets compacted at turret speed 35 rpm (47 m/min) (EHF: extended head flat, SHF:
standard head flat, RHF: reduced head flat)

Punch Compaction Tensile ER Porosity Ejection AQ


head force strength force
configuration (kN) (MPa) (%) (%) (N)

EHF 6.0 0.23 (0.02) 5.4 (0.3) 20.58 (0.17) 179 (11) 0.79
SHF 0.21 (0.02) 7.3 (0.3) 20.83 (0.03) 162 (7) 0.82
RHF 0.19 (0.01) 7.9 (0.2) 21.06 (0.21) 165 (12) 0.88

EHF 9.0 0.48 (0.02) 8.1 (0.2) 17.02 (0.08) 287 (6) 0.91
SHF 0.42 (0.01) 9.7 (0.2) 17.25 (0.08) 293 (13) 0.96
RHF 0.38 (0.01) 10.4 (0.2) 17.39 (0.15) 281 (8) 1.02

Parenthesized values represent the respective standard deviation.

133
Nonetheless, the trend of enhanced plastic yielding with increasing head flat

diameter was similar for turret speeds of 25 (Fig. 28) and 35 (Table 20) rpm. Tablets

compacted using the EHF punch head configuration continued to exhibit the best

mechanical quality with highest tensile strengths and least tablet expansion,

followed by SHF and RHF head configurations (Table 20).

4.2.9 Summary

In this study, the influence of tableting punch head geometry, in relation to

compaction roll profile, on tablet mechanical properties using lactose and

paracetamol-starch granules was investigated. The use of punches with head

geometry capable of increasing the dwell time was advantageous as the increased

dwell time often improved the mechanical strength of tablets and reduced capping

tendencies. Analysis of the lactose and paracetamol-starch tablet properties revealed

tablets produced from punches with a flat head profile consistently displayed better

mechanical quality than tablets produced from punches that lacked a head flat. The

presence of a head flat on the punch head appeared to be crucially important to allow

a state of constant strain for the compacts to undergo stress relief. Tablet tensile

strength showed a positive relationship with dwell time indicating that an increase

in dwell time by even a small extent caused significant improvements to the tablet

mechanical quality. Prolongation of dwell time was advantageous as it allowed

greater plastic flow during dwell phase and consequently reduced tablet expansion

during decompression, demonstrated using experimental and computational data. In

contrast, the effect of the punch head radius modifications on the compaction

parameters was considered marginal at the experimented force levels.

The positive impact of dwell time extension was also noticeable at a higher tableting

speed. The use of a smaller compaction roll caused the strain rate to increase

134
dramatically, facilitated air entrapment in the tablets and thus caused the tensile

strength and elastic recovery of the tablets to decrease and increase, respectively.

The increase in dwell time by the smaller roll enhanced stress relaxation in the

compacts but was insufficient to counteract the rise in elasticity in the compact due

to increased strain rate during consolidation stage. The findings could be used to

provide a comprehensive understanding on the impact of punch head and

compaction roll modifications on tablet properties and aid tablet/tooling/press

manufacturers to make processing and designing considerations.

135
4.3 Part 3. Investigation of the effect of punch face modification on compaction

parameters and tablet physical quality

4.3.1 Influence of punch face edge geometry on compaction parameters

The compaction parameters obtained for the different punch face edge geometries at

main compaction event were studied. The punch face edge modification was found

to cause negligible effect on the force-time profiles, with the consolidation (35 ms),

dwell (23 ms) and decompression (20 ms) times showing almost identical values for

both configurations. The punch face edge modification was however found to have

noticeable effect on the set main compaction thickness (STmain), related to the tablet

thickness at the specified force level and in-die porosity of the compact formed

(Table 21). In comparison to the FFBE geometry, the STmain values for FFRE

geometry were consistently lower at the investigated compaction forces, indicating

deeper penetration for the FFRE punch in the die cavity during the compaction cycle

to achieve a desired compaction load. This caused the volume, and consequently the

in-die porosity, of the compact formed by the radius edge punch geometry to be

relatively lower than that of the beveled edge geometry at equivalent compaction

forces, despite the FFRE punch face configuration having a smaller cup volume

(25.10 mm3) than the FFBE configuration (26.03 mm3). This suggested that changes

observed in the packing efficiency and resistance to densification of the powder

particles were due to the alterations in punch face edge geometry.

During a compaction cycle, there are differential movements of powder within the

die, largely initiated from the punch face governed by frictional conditions between

the compact and tooling surface. Central regions of the compact show relatively

greater movement than at the peripheral regions in contact with the die walls [91].

136
Table 21. Effect of punch face edge geometry on set compaction thickness at the
main compaction event

Punch Compaction STmain


face force (mm)
configuration (kN) Lactose Paracetamol
-starch

FFBE 6.0 2.51 2.80


7.0 2.44 2.72
8.0 2.39 2.65
9.0 2.32 2.60

FFRE 6.0 2.47 2.73


7.0 2.38 2.65
8.0 2.32 2.59
9.0 2.24 2.53

The powder mass in the die cavity during consolidation is believed to be in a

fluidized state with air attempting to escape through the clearance between the punch

tips and die wall, thus powder movement and migration of air pockets within the

powder mass are largely believed to be inter-related. From the punch penetration

data (Table 21), it is reasonable to suggest that the radius edge geometry facilitated

the smooth transition of powder particles along the moving punch face and provided

lesser resistance to powder movement during the consolidation phase, allowing

increased radial movement of powder particles [92]. Axial movement of particles at

the die wall/compact interphase relative to the center of the compact was possibly

enhanced due to accelerated particle movement along the radius punch face. The

favorable powder movement by the curved edges of FFRE geometry could also have

facilitated migration of air from the central regions of the powder mass to the

peripheral regions adjacent to the die wall. This consequently allowed greater

powder densification and possibly aided in reducing variation in density distribution

137
within the compact. In contrast, the sharp angled edges and extended flat surface of

the FFBE geometry likely provided greater resistance to movement of powder

particles along the punch face surface thus affecting the particle rearrangement

during the consolidation stage of the compaction process. The radial powder

movement from the peripheral regions of the compact to the center of the compact

was relatively reduced and impeded the extent of axial powder movement along the

die walls, subsequently resisted the flow of air through the powder mass in the

confined cavity of the die. The packing efficiency of the radius edge geometry was

further demonstrated by the widening difference in ST main values between the two

geometries with increasing compaction force during production of lactose tablets

(Table 21). At higher compaction forces, the rate of compaction force application

rises dramatically, providing lesser time for powder movement and air exhaustion.

However, with the curved surface of the FFRE geometry, this negative effect was

reduced by a greater extent than by the angled edges of the FFBE geometry.

The results obtained for the paracetamol-starch tablets showed that the design

modification in the punch face edge affected the compaction parameters of the

paracetamol-starch tablets similarly, with some differences. The STmain values for the

paracetamol-starch granules were greater than the values observed for the

compaction of lactose granules (Table 21) under equivalent compaction conditions.

While the trend of lower STmain values for the FFRE geometry was maintained

(Table 21), the differences between the STmain values for the two edged punches

during paracetamol-starch tablet production were generally consistent at increasing

compaction force. Both observations could be attributed to the elastically deforming

nature of the paracetamol-starch particles [78]. At higher compaction forces, the rate

of force application increased considerably and probably exceeded the rate at which

138
the particles could react. This caused the paracetamol-starch granules to experience

much more elastic deformation at higher forces and increased resistance to

densification, altering the in-die movement of the particles.

4.3.2 Influence of punch face edge geometry on tablet properties

The changes to the mechanical properties of lactose tablets produced using single

compaction cycles with respect to the different punch face edge geometries were

studied. The FFRE tablets consistently displayed higher tensile strength than FFBE

tablets at the investigated compaction forces as shown in Fig 33. Tabletability

decreased in accordance with the change of punch face configuration from FFRE

(Cp = 0.102) to FFBE (Cp = 0.090). By comparing the extent of tablet elastic recovery

(Table 22), it could be seen that the radius edge geometry reduced elasticity in the

compact, prior to punch withdrawal during the decompression phase. This ensured

that the final tablet porosities (Table 22), measured upon ejection, followed a trend

parallel to STmain values, with the FFRE tablets displaying lower tablet porosities

than the FFBE tablets.

It should be noted that the compaction profiles for radius and bevel punch face

configurations were nearly identical, without observable changes to the symmetry

of the curve during the dwell phase. This showed that the differences in tablet

expansion was contributed significantly by elasticity development during the

consolidation phase. Apparently, the different extent of punch penetration by the two

types of punches during tableting led to changes to the tablet internal character as

previously described and contributed to differences in the tablet physical quality.

The FFRE configuration, with enhanced particle movement and air removal,

probably reduced variation in density distribution within the compact accompanied

by greater degree of densification causing the particles to be packed closer together.

139
0.8
Tensile strength (MPa)

0.6

0.4

0.2

0
6 7 8 9
Compaction force (kN)

Figure 33. Tensile strength of lactose tablets prepared using single compaction cycle
by FFBE (■) and FFRE (□) punch face configurations.

Table 22. Properties of lactose tablets prepared using single compaction cycle by
FFBE and FFRE punch face configurations

Punch Compaction ER Porosity Compaction Ejection


face force quotient force
configuration (kN) (%) (%) (N)

FFBE 6.0 7.1 (0.2) 19.80 (0.04) 0.077 212 (1.4)


7.0 7.6 (0.1) 18.61 (0.25) 0.108 321 (1.2)
8.0 8.5 (0.3) 17.76 (0.18) 0.130 346 (1.4)
9.0 9.4 (0.1) 16.35 (0.05) 0.126 405 (3.1)

FFRE 6.0 6.8 (0.1) 19.02 (0.04) 0.070 224 (1.0)


7.0 7.0 (0.2) 18.13 (0.13) 0.082 330 (3.6)
8.0 7.7 (0.3) 17.16 (0.10) 0.092 354 (1.6)
9.0 8.7 (0.2) 15.46 (0.21) 0.102 420 (2.5)

Parenthesized values represent the respective standard deviation.

140
In comparison, beveled edges introduced more heterogeneity in the compact internal

structure that probably facilitated development of much more stored elastic energy

in the tablet, as evidenced from the tablet expansion data. The elastic response of the

powder during compaction in accordance with changing edge geometries was also

reflected in the ejection force recorded, which is a function of residual die wall

pressure and friction conditions in the die (Table 22). Compacts prepared from FFRE

punch generally exhibited higher ejection force values than the compacts produced

from the punches with the bevel edge (p < 0.05). Increased strain recovery for FFBE

tablets caused compacts to have greater axial recovery and contracted radially to a

greater extent, thus leading to lower residual die wall pressure and correspondingly,

a lower ejection force. The differences in compaction efficiency between the edge

geometries was further demonstrated by the force-punch displacement curves.

Radius edge punches consistently exhibited greater area under the curve, indicating

a greater work of compaction and reduced energy utilization for expansion (Fig. 34).

10000
Compaction force (N)

8000

6000

4000

2000

0
0 2 4 6 8
Punch separation (mm)

Figure 34. Compaction force-punch separation curves for the FFBE (●) and FFRE
(○) punch face configurations.

141
The influence of the punch face edge geometry on the mechanical properties of

tablets compacted using paracetamol-starch granules at main compaction event was

also studied. Paracetamol-starch tablets generally experienced reduced tensile

strength (Fig. 35) and greater elastic recovery accompanied by tablet failure in the

form of capping (Table 23), compared to corresponding lactose tablets at equivalent

compaction forces. This was attributed to the excessive elastic deformation of

paracetamol-starch granules as previously described under Section 4.2.2. However,

the benefit of punch face edge modification on tablet mechanical properties was

evident, a similar trend as exhibited by the lactose tablets, albeit with some

limitations.

FFBE tablets generally displayed a higher propensity to cap than FFRE tablets as

evidenced by the higher capping index values (Table 23). Capping incidences during

high speed rotary production was observed to be more frequent when beveled

punches were used. The greater capping severity of the FFBE tablets was also

observed during the hardness testing during which double capping (capping of top

and bottom of tablet crowns) occurred, particularly at higher compaction forces. The

improved resistance to tablet failure with face edge modification again reinforced

the earlier discussion that the favorable particle densification state and reduced

internal density differences within the compact formed by FFRE geometry had been

advantageous. Extensive elastic expansion in FFBE tablets promoted propagation of

fissures or cracks along the density gradient boundaries of the high pressure regions,

often situated near the moving punch faces and the central body of the compact,

during the unloading phase, amplifying the capping tendency in FFBE tablets.

142
0.5

0.4
Tensile strength (MPa)

0.3

0.2

0.1

0
6 7 8 9
Compression force (kN)

Figure 35. Tensile strength of paracetamol-starch tablets prepared using single


compaction cycle by FFBE (■) and FFRE (□) punch face configurations.

Table 23. Properties of paracetamol-starch tablets prepared using single compaction


cycle by FFBE and FFRE punch face configurations

Punch Compaction Capping ER Porosity Compaction Ejection


face force index quotient force
configuration (kN) (%) (%) (N)

FFBE 6.0 1.53 13.1 (0.4) 19.03 (0.31) 0.164 87 (1.6)


7.0 1.82 13.7 (0.2) 18.08 (0.30) 0.171 88 (2.9)
8.0 2.27 15.9 (0.1) 17.12 (0.07) 0.186 96 (1.4)
9.0 2.69 17.0 (0.4) 15.85 (0.07) 0.198 105 (2.3)

FFRE 6.0 0.98 11.8 (0.4) 18.45 (0.13) 0.140 93 (2.6)


7.0 1.36 12.9 (0.4) 17.39 (0.17) 0.155 98 (2.3)
8.0 1.81 15.1 (0.2) 16.28 (0.13) 0.166 104 (3.9)
9.0 2.79 15.7 (0.3) 14.60 (0.14) 0.175 103 (1.1)

Parenthesized values represent the respective standard deviation.

143
The positive effect of the curved edges in alleviating the tendency of the tablets to

cap tapered off at the highest compaction force of 9 kN with both tablets showing

comparable capping indices indicating the limitation of edge modification. The

outcomes concurred with the results reported in Part 2, in which dwell time extension

had limited effect on tablet properties at higher compaction forces, attributed to

development of high elasticity in the tablet structure.

4.3.3 Discrete Element analysis

DEM is an effective tool to simulate compaction mechanics of granular particles.

Setting up of detailed models can be time consuming and it can also be limited by

computational power, which restricts the number of particles and the size of the

system that can be processed. In this study, approximately 3700 spherical particles

were simulated to form a theoretical 325 mg tablet. This ensured the completion of

a simulation run within a feasible computing timeframe. A constant compaction load

to total particle surface area ratio was maintained. The compaction process was

simulated using three types of punch face configurations – FFP, FFRE and FFBE.

Computational data of powder compaction from the FFP face configuration was

used as a reference.

4.3.3.1 Comparison between simulations and experiments

Usability of DEM simulations to predict the performance of real-world biaxial

compaction process was validated by comparing experimental measurements and

computed responses derived from the simulations [141]. Tablet thickness at target

compaction force and rate of compaction force application (RFA) were chosen as

the simulation variables for the comparison (Fig. 36).

Tablet thickness. As shown in Fig. 36A, qualitatively the effects of punch edge

geometry on tablet height were similar in both experimental results and simulations.

144
A
4.8
Simulation tablet thickness (mm)

4.6

4.4

4.2
3.4 3.6 3.8 4
Experimental tablet thickness (mm)

B
800
Simulation RFA (N/ms)

700

600

500

400
150 200 250 300 350
Experimental RFA (N/ms)

Figure 36. Comparison between simulation and experimental tablet thickness (A)
and RFA values (B) for FFP (●), FFRE (♦) and FFBE (▲) tablets.

145
FFP tablets consistently exhibited the lowest simulated thickness, followed by FFRE

and FFBE tablets (p < 0.05). The simulated thickness values for FFP, FFRE and

FFBE tablets correlated positively with their corresponding experimental results

respectively. One-to-one quantitative comparisons showed that the experimental

values were lower than the simulation values. This could be attributed to the larger

particle size used in simulation and reduced particle rearrangement during the

compaction process.

Rate of Force Application. In both simulations and experiments, simulated RFA

was shown to increase with increasing compaction force for all three types of flat

punch edge geometries. It was also noted that simulated RFA values (Fig. 36B)

correlated positively with experimental RFA values (R2 = 0.99). At any one

compaction force, simulated RFA did not differ significantly between the punches,

agreeing with experimental results (p > 0.05). Quantitatively, experimental RFA

values were lower than simulation values at all investigated compaction forces.

Generally, the trends observed for the mentioned variables showed good qualitative

agreement with experimental values. Quantitative comparison between the two

methods showed some differences as can be expected due to limitations of the

simulation model (larger and spherically shaped particles, approximated contact

parameters, neglected effects of air movement or frictional effects among others).

These comparisons however indicated that the numerical calculations with the

current simulation parameters could accurately capture essential compaction

dynamics, and simulation calculations took place within a feasible computational

time. DEM proved to be a suitable tool for predicting trends with regard to powder

compaction.

146
4.3.3.2 Effect of punch face edge geometry on tablet stress distribution

DEM was employed to investigate how the compaction force would be transmitted

within a powder bed during biaxial compaction with flat-face punches of different

configurations. Computational stress analysis of the tablet cross-sectional area was

prepared for examination. A grid bin group consisting of 96 bins was formed at the

tablet (Y-Z) cross-section as shown in Fig. 37(i). The mean apparent particle stress

of FFP tablet at the lowest compaction force was used as reference value for the

calibration bar. This allowed meaningful comparison between the different types of

tablets by the use of a common stress scale.

As expected, the force of compaction imposed by the upper and lower punches was

not uniformly transmitted within the three tablet types. The differences in the stress

distribution plots were more noticeable at the higher compaction forces. Fig.

37A(ii)-(iii) show the computed relative stress distribution present in FFP tablets

prepared using compaction forces from 8-9 kN. Stress distributions were observed

to be fairly symmetrical with higher shear surfaces created at the central regions of

the tablets and at the interface between the die wall and particles. This form of

symmetrical stress distribution is consistent for a biaxial compaction. The lack of

perfect symmetry may reflect the variation in initial particle arrangement caused by

imperfect die filling and random movement of particles during compaction. Shear

friction formed at the regions adjacent to the die wall indicated higher resistance to

particle movement during the compaction process at those regions. As the

compaction force increased, force components acting on the particles were

accommodated by an increase in local stresses directed towards the central regions

of the tablets, producing loci of high stress. The findings were in general agreement

with those reported in other studies [94, 109].

147
A B C

(i)

(ii)

(iii)

Figure 37. (i) Compaction of particles with grid bin group in the vertical Y-Z cross-section obtained for FFP (A), FFRE (B) and FFBE (C) tablets.
Fig. 37 (ii-iii) Stress distribution in the vertical Y-Z cross-sections obtained for FFP (A), FFRE (B) and FFBE (C) tablets at compaction force (ii)
8.0 and (iii) 9.0 kN. The X and Y axes of the contour plots represent the distance (mm) from the edge and bottom of the tablet respectively.

148
The introduction of a cup to the punch face configuration was shown to cause

significant changes in the stress contour plots. Fig. 37(B-C) shows the computed

relative stress distribution in the FFRE and FFBE tablets with the corners of the plots

representing the regions adjacent to the radial/bevel edges of the punches. The

variation in stress distribution within the FFRE tablet (Fig. 37B(ii)-(iii)) appeared

much greater and less homogenous than in the corresponding FFP tablet (Fig.

37A(ii)-(iii)). The upper half of the tablets generally experienced more stress with

the formation of a high stress core near to the symmetry axis. This indicated that the

radius edge tool significantly increased interaction between particles and punch

surface, thus consequently changing the distribution of compaction load during

compaction. Resolution of increasing compressive force exerted by the FFRE

geometry on the particles interacting with the curved edge of the punch surface

showed that the resultant forces acted towards the center of the tablet at an increased

rate compared to FFP punches. This caused the stress zones at the tablet core to

increase more in intensity and area than in corresponding FFP tablets. This condition

occurred despite the greater tablet volume measured for the FFRE tablets at

equivalent compaction forces. The compressive stress loci occurred in a similar

manner as in the FFP tablet, primarily towards the core regions. However, when a

higher compaction force was used, regions around the circumference of the die wall

experienced reduced stress relative to the central regions due to decreased particulate

interaction possibly caused by increased radial movement of particles. This resulted

in a steep stress gradient between the tablet edges and core with large amounts of

stored energy within the tablet itself.

149
From comparisons of the FFRE contour plots (Fig. 37B(ii)-(iii)) with the FFBE plots

(Fig. 37C(ii)-(iii)), it was apparent that the change from a curved edge to an angled

straight slope caused stress contour of the central regions of the tablet to change

significantly. Horizontal component of forces acting on particles interacting with the

bevel edges allowed a more intensive transfer of shear in the direction of the internal

volume of the tablet with reduced shear surfaces formed at the interface between the

particles and the punch surface. This resulted in formation of shear bands running

from the top edges towards the central regions with subsequent dissipation of the

forces acting between the die wall and particles. The results for the FFBE specimens

showed that the highest stress regions tend to be concentrated at the symmetry axis

with the maximum stress achieved at the central regions shown to be higher than in

corresponding FFP or FFRE tablets. This occurred despite the reduced state of

densification achieved in FFBE tablets. These observations implied that bevel edges

direct force dissemination in a manner that facilitate the development of much more

stored energy in the tablet.

Stress analysis results provided further support to the experimental data on tablet

properties. Curved edge surface of FFRE punches facilitated a more favorable

distribution of the compaction load among the particles compared to the FFBE

punches, thus allowing particles to attain a lower average stress within the tablet

volume and to be compressed to a greater extent to achieve the target compaction

force. High stress regions for the FFRE tablet also seemed to be reduced in intensity

and area. In comparison, the presence of a bevel edge elevated the non-uniformity

of load distribution, leading to greater stress anisotropy in the tablet microstructure.

Compressive forces were directed towards the central regions of the tablet

facilitating localization of stress and created intensive stress gradient in the regions

150
between the constrained and unconstrained tablet body. Upon removal of the applied

force, the release of a higher amount of stored elastic energy caused extensive elastic

recovery in FFBE tablets. This facilitated the disruption of bonds between particles

and propagation of fissures or cracks along the stress gradient boundaries in the

FFBE tablets during the unloading phase. Particle deformation and interparticulate

bonding were reduced as a consequence, thus directly reducing compact mechanical

strength and amplifying the capping tendency in FFBE tablets.

4.3.4 Effect of tableting speed

The relationship between tableting speed, punch face edge geometry and the

resultant tablet properties was also evaluated (Table 24). The increased tableting

speed narrowed the force-time profiles for both edged punches similarly, with

reductions in consolidation (26 ms), dwell (16 ms) and decompression (14 ms) times.

The higher tableting speed caused the STmain values to increase for both punches,

indicating reduced punch penetration and higher in-die compact porosities at the

respective compaction forces. This could be explained by the greater resistance to

compaction by particles at the increased rate of punch displacement, explained

earlier under Section 4.2.8. Nonetheless, the deeper punch penetration by the radius

edge punch was well preserved. Interestingly, the differences in the STmain values

between the two punch face configurations increased at the higher tableting speed.

This could again be explained by the better particle flow and favorable force

transmission by the radius edges. The accelerated compaction process (increased

rate of force application) restricted particulate movement and reduced time available

for particle rearrangement and air escape through the powder mass.

151
Table 24. Properties of lactose tablets prepared using single compaction cycle by FFBE and FFRE punch face configurations at turret speed 35 rpm
(47 m/min)

Punch Compaction STmain Tensile ER Porosity Compaction Ejection


face force strength quotient force
configuration (kN) (mm) (MPa) (%) (%) (N)

FFBE 7.0 2.47 0.27 (0.03) 7.1 (0.1) 19.90 (0.06) 0.125 200 (4.6)
9.0 2.34 0.45 (0.01) 9.0 (0.3) 17.63 (0.17) 0.148 320 (0.8)

FFRE 7.0 2.40 0.32 (0.02) 6.2 (0.2) 19.43 (0.03) 0.083 209 (4.6)
9.0 2.25 0.50 (0.02) 8.3 (0.2) 16.72 (0.41) 0.110 339 (5.0)

Parenthesized values represent the respective standard deviation.

152
The ability of the curved surface of the FFRE punch face to allow particle movement

with reduced slide resistance led to better tableting performance under the less

unfavourable tablet formation conditions. In addition, the radius edge’s ability to

reduce the generation of localized stress regions probably aided particle movements

during the consolidation phase.

Tablets compressed at the higher tableting speed displayed poorer tablet physical

quality. However, tablet expansion was reduced in contrast to tablets compacted at

the lower turret speed 25 rpm (Table 22), indicating that the amount of stored elastic

strain was possibly reduced when tablets were prepared with increased compaction

velocity. Thus, the reduction in tablet mechanical strength was attributed to the

higher compact porosities and the resultant reduced interparticulate bonding in the

tablets, rather than due to disruption of formed bonds during decompression. When

compared to the results obtained at turret speed 25 rpm (Table 22), the compaction

efficiency expectedly decreased for both punch face configurations at the higher

tableting speed, as indicated by the values for the compaction quotient (Table 24).

Interestingly, between the two punch face edge geometries, the compaction quotient

for radius edge punch was least affected by the increase in tableting speed. This

suggested that the modification of the punch face curvature could possibly counter

the adverse impact of high tableting speed and preserve compaction efficiency of

compaction tooling to some extent.

4.3.5 Effect of precompaction

The effect of precompaction to the mechanical properties of paracetamol-starch

tablets with respect to the different punch face edge geometries are depicted in Table

25. A precompaction cycle (Tcon: 30 ms, Tdwell: 20 ms, Tdec: 20 ms) with a peak force

of 3 kN was introduced prior to the application of the main compaction forces.

153
Table 25. Properties of paracetamol-starch tablets prepared using double compaction (mode1) cycle by FFBE and FFRE punch face configurations

Punch Compaction STpre STmain Tensile Capping ER Porosity Compaction Ejection


face force strength index quotient force
configuration (kN) (mm) (mm) (MPa) (%) (%) (N)

FFBE 6.0 3.21 2.76 0.42 (0.01) 0.78 11.1 (0.5) 17.31 (0.17) 0.118 124 (1.4)
7.0 3.21 2.69 0.48 (0.02) 0.92 11.8 (0.3) 16.41 (0.08) 0.144 135 (2.1)
8.0 3.21 2.63 0.52 (0.02) 1.20 14.2 (0.7) 15.89 (0.11) 0.162 148 (1.8)
9.0 3.21 2.58 0.56 (0.03) 1.59 14.6 (0.5) 15.37 (0.04) 0.173 147 (2.0)

FFRE 6.0 3.11 2.70 0.47 (0.01) 0.29 9.6 (0.2) 17.01 (0.09) 0.090 117 (2.4)
7.0 3.11 2.62 0.53 (0.02) 0.27 10.7 (0.2) 15.99 (0.21) 0.095 141 (0.5)
8.0 3.11 2.55 0.60 (0.01) 0.53 11.4 (0.3) 15.27 (0.09) 0.113 143 (1.4)
9.0 3.11 2.50 0.64 (0.02) 0.93 12.7 (0.3) 14.54 (0.19) 0.123 141 (6.0)

Parenthesized values represent the respective standard deviation.

154
Main compaction cycle was unaffected by the precompaction cycle; consolidation,

dwell and decompression times remained largely indifferent. STmain values however,

decreased further for both punch face configurations (Table 25), indicating a positive

effect on powder densification. Introduction of precompaction led to greater

reduction in porosities of the compacts in the die aided by removal of significant

amount of air in the powder bed at precompaction event which reduced resistance to

powder movement during the main compaction stage. Nonetheless, the punch

penetration to achieve the desired load was still different between the two edge

geometries; FFRE punches continued to display deeper punch penetration at both

precompaction (STpre 3.11 mm) and main compaction events. Interestingly, the

differences between the bevel and radius punches’ ST main values widened at

increasing compaction force, a trend that was not observed in paracetamol-starch

tablets compacted using single compaction cycles. This indicated that the positive

effect of the radius edge on powder packing was amplified when double compaction

cycle was utilized.

The positive influence of precompaction was also reflected by the improvement in

tablet physical quality for both tablet designs. Tablet tensile strength significantly

increased with corresponding reduction in elastic recovery, porosity and capping

tendency (Table 25), probably associated with the neutralization of elastic energy

during the precompaction phase which resulted in lesser tablet recovery and

interparticulate bond disruption. Despite the pronounced effect of precompaction on

tablet properties, the differences in tablet quality between the FFRE and FFBE

tablets were still evident; FFRE tablets displayed better tablet physical quality than

the FFBE tablets. This again was attributed to the deeper punch penetration enabled

by the FFRE geometry, at both pre and main compaction.

155
The FFRE punch face’s ability to disseminate elastic energy in tablets and

consequently reduce capping tendency was well demonstrated by the increased

disparity in the capping index values between the two punch face configurations

(compared to capping index values from single compaction), even at the highest

compaction force of 9 kN as shown in Table 25. The total compact stress at punch

face level was distributed more equally due to the growing interparticle contacts

within the compact formed by the radius punch face, thus reduced the local shear

stress experienced by the particles in the compact and minimized the tendency to

cap. Overall, the findings demonstrated the agonistic effect of precompaction on the

tableting performance of radius edged punches and could provide an alternative in

resolving tablet failure issues.

4.3.6 Effect of roll displacement

The effect of precompaction on the compaction parameters and mechanical

properties of paracetamol-starch tablets with respect to the different edge geometries

was further investigated with activation of the floating roll technology. The air

compensator was activated at the precompaction stage by fixing the precompaction

force (3 kN) and set precompaction thickness (STpre 2.82 mm). This caused the

displacement of the roll and extended the dwell phase (Tdwell: 28 ms) at

precompaction event. Due to the ability of the FFBE face configuration to achieve

the desired compaction load with reduced punch travel, FFBE punches experienced

greater displacement as compared to the FFRE punches, as shown in Fig. 38. The

dwell times for both punch configurations (evaluated from the displacement-time

curve) remained relatively similar despite the different degrees of displacement.

156
0.4

Roll displacement (mm)


0.3

0.2

0.1

0
0 20 40 60 80
Time (ms)

Figure 38. Roll displacement-time curves for the FFBE (—) and FFRE (---) punch
face configurations with the activation of the air compensator (mode2) during double
compaction cycle.

The extended dwell time facilitated further removal of air from the powder bed at

precompaction and consequently reduced in-die compact porosities at the

precompaction stage (evaluated from the addition of STpre and roll displacement).

This subsequently allowed easier particle consolidation and lower in-die porosities

at the main compaction phase, reflected by the reduced STmain values (Table 26). The

FFRE punch continued to show greater punch travel with lower ST main values then

the FFBE punch. Tablet physical properties were significantly improved with the

extension of dwell phase at precompaction with the FFRE tablets particularly

experiencing a substantial increase in tablet strength (Table 26). The results were

again attributed to the efficient particle packing by the radius edge geometry and

aided by greater interparticulate bond reinforcement and plastic flow within the

compact during the extended contact time [45, 119].

157
Table 26. Properties of paracetamol-starch tablets prepared using double compaction (mode2) cycle by FFBE and FFRE punch face configurations

Punch Compaction STpre STmain Tensile ER Porosity Compaction Capping Ejection


face force strength quotient index force
configuration (kN) (mm) (mm) (MPa) (%) (%) (N)

FFBE 7.0 2.82 2.69 0.55 (0.07) 11.2 (0.3) 16.02 (0.12) 0.132 0.20 137 (2.4)
9.0 2.82 2.57 0.81 (0.05) 13.1 (0.1) 13.84 (0.06) 0.157 0 146 (3.0)

FFRE 7.0 2.82 2.60 0.76 (0.02) 9.7 (0.2) 14.47 (0.10) 0.086 0 133 (3.0)
9.0 2.82 2.47 0.99 (0.03) 11.8 (0.2) 12.87 (0.06) 0.106 0 144 (2.1)

Parenthesized values represent the respective standard deviation.

158
However, it was found that the tablets produced using both FFBE and FFRE punches

could be considered equivalent in terms of capping tendency, despite obvious

differences in the states of elastic strain condition in both types of tablets as indicated

by the significantly different elastic recovery values (Table 26). The results

demonstrated the considerable influence of dwell time extension in effectively

counteracting the elastic strain built-up in the compact. Nonetheless, the value of the

punch face edge curvature at improving the tableting performance is clear. FFRE

tablets compacted at equal compaction forces to FFBE tablets generally displayed

better physical properties, attributed to deeper punch penetration and enhanced

compaction efficiency of the FFRE punches as previously mentioned. This

advantage of the FFRE configuration can also be extrapolated when tableting at

equal thickness is involved, as used by most tablet manufacturers. Tablets

compacted by the FFRE punch configuration would be able to attain equal thickness,

as tablets formed using the FFBE punch face, at relatively lower compaction forces.

This allows the punch tip to experience lesser stress during tablet production and

could possibly also lengthen the tooling lifespan.

4.3.7 Summary

The use of flat-face punches with a radius edge face configuration was advantageous

in tablet production as it allowed greater powder densification which often improved

the mechanical strength of tablets and reduced capping tendencies. Analysis of the

lactose and paracetamol-starch tablet mechanical properties revealed that tablets

produced using the radius edge punches consistently displayed better mechanical

quality than tablets from the bevel edge punches. The presence of the radius

curvature on the punch face edge was able to allow deeper punch penetration in the

die cavity during the compaction cycle, causing the compact structure to experience

159
enhanced powder consolidation which is necessary for interparticulate bond

formation and efficient dissemination of compaction load through the compact. This

latter effect allowed better utilization of the compaction force energy and reduced

local elastic strain energy developments, thus consequently reduced tablet expansion

during decompression. Stress analysis results provided further support to the

experimental data on tablet properties. Curved edge surface of FFRE punches

facilitated a more favorable distribution of the compaction load through the

particulate system as compared to the FFBE punches, thus allowing particles to

attain a lower average stress within the tablet volume and to be compressed to a

greater extent to achieve the target compaction force. High stress regions for the

FFRE tablet also seemed to be reduced in intensity and area. In comparison, the

presence of a bevel edge elevated the non-uniformity of load distribution, leading to

formation of localized shear stresses in certain regions and contributed to

development of elasticity within the compact.

The positive impact of the radius face configuration on compaction was also

noticeable at a higher turret speed. The application of a precompaction force and

dwell phase extension (via roll displacement) at the precompaction event amplified

the tableting performance of radius face punches to a greater extent when compared

to bevel face punches, due to enhanced stress relaxation of particles under

compaction by the radius face at precompaction stage.

This study has provided a better understanding on the implications of edge

modifications on flat-face punches on the tableting process. Tooling manufacturers

have shown that the radius edge face configuration is an upgrade over the bevel edge

face configuration in terms of tooling strength. The findings from this study have

proven that edge modification could also be beneficial in overcoming compression

160
challenges such as capping and could provide an alternative to the commonly used

bevel shaped punch face configuration.

161
4.4 Part 4. Investigation of the effect of punch face modification on powder

adhesion tendency to the punch surface during powder compaction

4.4.1 Physical characteristics of tableting material

The particle size properties of lactose granules are presented in Section 4.2.1. The

ibuprofen particles had a median particle size 130 ± 6.2 µm, with span 2.32 ± 0.05.

In comparison to the paracetamol-starch and lactose granules used in Part 2 & 3 of

the study, the flowabilty of ibuprofen-lactose blend was poorer (Hausner ratio 1.35

± 0.03, AOR 40.8 ± 1.3°).

4.4.2 Powder adhesion model

Tablets were produced from three different punch face configurations using the

manual press. Based on visual observation of residual powder distribution on the

punch faces at the end of the tableting cycle (30/50 tablets), it could be concluded

that powder adhesion in the current investigation adopted the adhesion model

proposed in literature [61]. The extent of powder adhesion found could be attributed

to various factors impacting the interrelationships between API-tooling surface

adhesive forces, API-excipient adhesive forces and API-API cohesive forces. When

the forces between the powder particles (API-API and API-excipient) exceed that of

the adhesive forces formed between the particles and the punch surface, sticking will

not occur as the stronger cohesion within the compact prevents powder transfer to

the tool surface. Conversely, when the adhesion of powder to punch surface is

stronger than the cohesive strength of the powder components in the compact,

powder sticking will occur. Powder adhesion to punches could be classified into two

basic categories - first and second-degree adhesion. First degree powder adhesion

was characterized by the formation of a relatively uniform monolayer on the punch

face as shown in Fig. 39A.

162
A B

Figure 39. FFP punch face showing first-degree (A) and second-degree (B) powder
adhesion.

Non-uniform powder adhesion with uneven powder masses on the punch face was

associated with second-degree adhesion (Fig. 39B). First-degree powder adhesion is

considered as a minor form of punch sticking, often referred to as “filming”. Filming

occured due to equivalent API-excipient adhesive and API-API cohesive forces,

with corresponding higher API-punch surface adhesive interactions. Formation of

film on the punch face led to dull grainy appearance of the tablet. Filming accelerated

thickening of the API layer on the tooling surface with continuation of the tableting

process, eventually leading to second-degree powder adhesion scenario. This

occurred when the cohesive forces between the API-API particles were greater than

that of the adhesive forces between the API and other tablet components

(excipients), causing the API particles from the tablet body to detach and bond with

the adhered API on the punch surface.

4.4.3 Influence of punch face configuration

The sticking quantification results were corrected for surface area of the punch face

geometries. Examination of the powder adhesion results indicated that the FFP

punches exhibited the highest propensity towards punch sticking at equivalent

compaction forces (Fig. 40).

163
A 70
Powder adherence (µg/mm2)
60

50

40

30

20
FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE
10
FFP
FFP

FFP
FFP

FFP
FFP
0
5 7.5 10
Compaction force (kN)

B 70
Powder adherence (µg/mm2)

60

50

40

30

20
FFRE

FFBE
FFBE
FFBE
FFRE

FFBE

FFRE
FFRE

FFBE
FFBE
FFRE
FFRE

10
FFP
FFP

FFP
FFP

FFP
FFP

0
5 7.5 10
Compaction force (kN)

Figure 40. Effect of punch face configuration on powder adherence on the upper (■)
and lower (□) punch faces after compacting 30 (A) and 50 (B) tablets using an
ibuprofen-lactose formulation containing 1 %, w/w magnesium stearate.

164
Modifications in the punch face configuration have clearly produced differences in

the cohesion-adhesion equilibrium in the tablet system. As described earlier under

Section 4.3.1, particle movement and compressive force resolution are greatly

affected by the punch face configuration. The flattened surface of the FFP punch

face restricted movement of particles at the powder bed boundaries and within the

compact matrix. Due to the absence of undulations on the FFP surface, shear stress

variations on the tablet undergoing compaction was largely avoided and resulted in

a relatively more homogenous stress distribution across the top and lower flat punch

surfaces. As a consequence, powder particles at the compact-tool boundary

experienced lower amount of disruptive side sliding shear during the consolidation

phase and facilitated formation of a reduced stress gradient along the punch face-

compact interphase. This facilitated greater opportunity for contact fusion of the

compact’s surface particles to the flat punch face. Introduction of concavity to the

punch tip (FFBE and FFRE) facilitated axial and radial movement of particles,

particularly from the peripheral to the central regions of the compact. As a result,

particles in contact with punch surface during the consolidation phase were afforded

much lesser constant contact time due to greater material sliding shear forces at the

punch boundaries. At maximum compaction, during which particulate movements

are minimized, the radius/bevel edges imparted variations in the shear force

distribution in the compact. The edged punch face configurations exerted pressure

gradients across the compacting material surface and created high density areas in

the compact particularly in regions adjacent to the edges. This led to reduced liability

of the particles adhering to tool surfaces, in particular, at the punch edges. These

combined effects weakened the adhesive forces of powder to the FFRE/FFBE punch

faces and led to reduced powder adhesion.

165
The findings illustrated the importance of punch face configuration optimization for

mitigation of different types of tableting problems. FFP configuration allowed a

homogenous stress and density distribution in the tablet, favorable for densification

and elasticity reduction in the compact. However, the enlarged flat area of the FFP

geometry accentuated particle adherence (Fig. 40). Angled/curved edges of FFBE

and FFRE face configurations, which were shown to facilitate unfavorable

localization of stresses and elasticity buildup within the tablet, were beneficial in

reducing material adhesion tendency (Fig. 40). The results were similar to those

reported by Roberts et al., whose study demonstrated the advantage of concave

tooling over flat tooling in mitigating powder sticking issues [100].

The advantage of the radius edge punch in terms of compact densification was again

evident in this part of the study. FFRE tablets exhibited significantly reduced post

compaction thickness than the FFBE tablets at equivalent compaction forces (results

not shown); tablets produced by the radius edge punches were denser than those by

the bevel edge punches. This suggested greater utilization of the compaction energy

for densification purpose which probably increased cohesion within the tablet

matrix. Nonetheless, these positive effects on compact densification led to minimal

differences in terms of powder adhesion. Comparison of powder adherence data

between the edged configurations (FFBE and FFRE) showed that overall

differences, if any, were rather small in magnitude and inconclusive (Fig. 40).

Distribution of powder on the FFBE and FFRE punch faces was however found to

be different during visual inspection. Powder adhesion onto the radius edge punches

was found to extend beyond the central flat region as compared to the bevel edge

punches in which powder sticking was limited to the central flat zone and did not

extend to the angled edges (Fig. 41). These observations may be attributed to the

166
differences in stress concentration factors on the edge geometries of the two flat-

face punches. Straight angled edges of the FFBE configuration experienced much

higher stress when compared to stress concentration along the central flat zone. The

high stress regions limited powder adhesion to the flat surfaces by imparting sliding

shear force along the bevel slope, causing particles to pack better, which facilitated

the formation of stronger bonds along the edges. On the other hand, curved edges on

the FFRE face configuration dissipated stress more homogenously during

compaction and minimized differences in stress concentration factors between the

flat and edged surfaces. Hence, less densely packed surface material was presented

at the vicinity of the radius edge as compared to that of the bevel edge. This resulted

in powder adherence beyond the central flat face, along the radial curvature at lower

compaction forces.

A B

Figure 41. Distribution of residual powder on flat-face bevel (A) and radius (B) edge
punch faces.

4.4.4 Influence of compaction force on powder adhesion

Powder adherence was observed to decrease with increasing compaction force,

regardless of punch face configuration (Fig. 40). This was attributed to the enhanced

state of particulate deformation at the higher force levels. Ibuprofen particles

167
reportedly undergo predominantly elastic deformation and lactose granules

experience a mixed mechanism of brittle fracture and plastic deformation.

Nonetheless, the greater degree of physical deformation under higher compaction

loads increased the available interfacial bonding surface and led to formation of

stronger interparticulate bonds. Stronger cohesion within the tablet matrix increased

the minimum force of adhesion required to cause particle detachment off the tablet

surface. However, the effect of compaction force is debatable as contrasting results

has been reported in studies [61, 87]. This could be attributed to different pressure-

sensitive characteristics of the compaction materials used in experiments.

Quantification of powder adherence on the punch faces indicated a greater sticking

tendency on the lower punches, for all punch face configurations and force levels

investigated. This phenomenon was attributed to the force application method

(uniaxial compaction) during tableting by the manual press. Compaction load was

exerted on the powder bed confined in the die by upper punch displacement while

the lower punch remained stationary. However, force transmission across the

compact between the two punches was probably incomplete largely due to partial

utilization of compaction energy to overcome frictional resistance. As a

consequence, the force experienced at the top half of the tablet was greater than at

the bottom regions of the compact. Interparticulate bonds at the peripheral regions

adjacent to the upper punch surface were enforced, leading to less powder adhesion

on the upper punch face. Conversely, the relatively longer contact time between the

tableting material and the lower punch face could also have strengthened adhesive

forces at the punch face boundary interface.

168
4.4.5 Influence of magnesium stearate concentration

Magnesium stearate is a widely used boundary lubricant with anti-adherent

properties and could possibly be used to negate material adhesion issue during

tableting. It acts by readily adsorbing onto surfaces of metal or particles to form

protective lubricant film layers. These layers reduce contact between tablet feed

particles to compaction tools, and could possibly weaken the bond strength of

particles to tool surfaces when formed thus providing anti-adherent function.

However, the use of a higher concentration of magnesium stearate in the

experimented formulation unexpectedly increased powder adherence levels on both

the upper and lower punches for all punch face configurations (Fig. 42) (p < 0.05).

This effect has been reported in literature, attributed to the formation of a eutectic

system between ibuprofen and magnesium stearate which possibly reduced the

thermal stability of ibuprofen [71].

During the compaction process, friction conditions within the die created localized

high temperature areas that exceeded the melting point of ibuprofen. The increased

temperature caused some degree of pressurized melting and concomitant changes on

the tablet surface, while under compressive stress. Rapid recrystallization occurred

on the surface of the compact with the removal of the compaction load and possibly

accelerated adhesion of ibuprofen to the punch surface. In addition, the formation of

thicker lubricant film surrounding the particles at higher magnesium stearate

concentrations probably weakened the cohesive forces within the tablet matrix by

hindering interparticulate bond formation between the powder components, shifting

the cohesion-adhesion equilibrium towards metal surface adhesion.

169
A 70
Powder adherence (µg/mm2)

60

50

40

30

20
FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE
10
FFP
FFP

FFP
FFP

FFP
FFP
0
5 7.5 10
Compaction force (kN)
B 70

60
Powder adherence (µg/mm2)

50

40

30

20
FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE

FFBE
FFBE
FFRE
FFRE

10
FFP
FFP

FFP
FFP

FFP
FFP

0
5 7.5 10
Compaction force (kN)

Figure 42. Effect of punch face configuration on powder adherence on the upper (■)
and lower (□) punch faces after compacting 30 (A) and 50 (B) tablets using an
ibuprofen-lactose formulation containing 1.5 %, w/w magnesium stearate.

170
4.4.6 Influence of compaction method (compaction simulator)

The influence of punch face configuration on powder adhesion was similarly

investigated under accelerated compaction conditions using the compaction

simulator. The compaction simulator enables tablets to be produced under conditions

representative of rotary production. Tablets were compacted using single

compaction cycle (total compaction time 100 ms, dwell time 20 ms). Fig. 43 shows

the amount of powder adhered on the upper punches after a continuous tablet

production cycle (50 tablets) using a compaction simulator, quantified by UV

spectroscopic assay measurements. Powder residue on the lower punches was not

evaluated due to possible inaccuracies caused by the horizontal shear caused by the

tablet removal scrapping action. In comparison to upper punch sticking data from

the single station press (Fig. 40), some quantitative differences in powder adherence

were noted although the trend remained and the overall amounts generally decreased

when tablets were compacted using the compaction simulator (Fig. 43).

20
Powder adherence (µg/mm2)

16

12

4
FFBE
FFRE

FFBE
FFRE

FFBE
FFRE
FFP

FFP

FFP

0
5 7.5 10
Compaction force (kN)

Figure 43. Effect of punch face configuration on powder adherence on the upper
punch face after tablet production in the compaction simulator, using an ibuprofen-
lactose formulation containing 1 %, w/w magnesium stearate.

171
This was attributed to differences in the compaction kinematics between the

simulator and the manual press. A much shortened compaction cycle in the simulator

caused more rapid punch movements and accelerated the movement of particles and

air pockets from the central regions of the compact towards the annular rings of the

die, creating a highly fluidized state throughout the powder bed. Material shear along

the punch face-compact boundary increased considerably as a consequence and

weakened the adhesive bonding between the particles at the compact surface and

punch during compaction cycle. In contrast, the reduced punch vertical velocity

during compaction in the manual press afforded much more time for particle

rearrangement and migration of air pockets within the powder bed, resulting in a

relatively reduced material shear at the interface. Increased contact fusion time

between particles and punch surface during the use of the manual press had probably

contributed to the disparity in the degree of powder adherence observed between the

two compaction equipment.

4.4.7 Influence of punch face configuration (compaction simulator)

The effect of punch face configuration on powder adhesion tendency was analyzed

using powder quantification data of the upper punches (Fig. 43) and take-off force

values measured from the lower punches (Fig. 44). FFP punches consistently

exhibited higher powder adherence propensity than the edged face configurations

used. Under ideal tableting conditions with formulations not prone to sticking issues,

a much lower take-off force would be required to dislodge the tablet from the lower

punch of the FFP face type than the edged punch faces. However, the amount of

residual powder and take-off forces for the FFP tablets were much higher than the

other two corresponding tablet types, at equivalent compaction forces. Therefore,

results supported the conclusion that the apparent adhesion behavior was clearly a

172
A 60

50

Take-off force (N)


40

30

20

10

0
0 10 20 30 40 50
Tablet sequence
B 60

50
Take-off force (N)

40

30

20

10

0
0 10 20 30 40 50
Tablet sequence
C
60

50
Take-off force (N)

40

30

20

10

0
0 10 20 30 40 50
Tablet sequence

Figure 44. Take-off force profile during the time course of tablet production using
the compaction simulator for the FFP (○), FFRE (●) and FFBE (▲) punch face
geometries at compaction force 5.0 kN (A), 7.5 kN (B) and 10 kN (C) using an
ibuprofen-lactose formulation containing 1 %, w/w magnesium stearate.

173
consequence of the higher sticking potential of the FFP face configuration. Based on

the powder quantification data (Fig. 43) and take-off force profiles (Fig. 44), the

differences in the powder adhesion tendency between FFBE and FFRE face

configurations could be considered marginal and inconclusive in attempts to rank

the sticking behaviors between the two punch edge geometries. Nonetheless, the

visually observed adherence pattern of powder on the FFBE and FFRE punches

resembled the results obtained using the manual press (Fig. 41). Powder adhesion

on the beveled punch surface was largely restricted to the central flat region of the

face configuration with minimal sticking on the angled surface, while residual

powder often extended beyond the flat area and occupied portions of the curved

surface of the radius edge geometry. This may again be due to the phenomenon of

homogenous stress distribution of compaction load along the radius edge of the

FFRE configuration, as explained earlier.

4.4.8 Influence of compaction force (compaction simulator)

Based on Fig. 43, it could be seen that the powder adherence on the upper punch

surface generally decreased with increasing compaction force, matching the trend

observed for the manual press. Analysis of the take-off force profiles (Fig. 44) during

the time course of tablet production indicated an influential effect of compaction

force on powder adherence process. At the lowest compaction force (5 kN), huge

variations in the take-off values were recorded in the profiles obtained for the various

punch face configurations. The variability seen on the profiles narrowed at higher

compaction force levels, suggesting a more consistent material transfer process at

the tablet-punch surface interphase. At the lower compaction force level (5 kN),

cohesion within the compact matrix was relatively weak due to reduced particle

deformation, thus the energy required for detachment of material from the tablet

174
surface was relatively reduced. This allowed varying amounts of material to

accumulate on the punch face. As the material mass on the punch face attained a

critical size, it broke off and become integrated into the subsequent tablet,

manifesting in a higher measured take-off force. At higher compaction forces, the

interparticulate bonds and cohesive forces within the compact were strengthened.

This increased the energy required to detach particles off the tablet surface and

facilitated in maintaining a relatively consistent bidirectional transfer of material

between the tablet and punch face.

4.4.9 Ejection force as sticking parameter

Tablet ejection force is a function of the pressure required to push the tablet out of

the die (influenced by residual die wall pressure) and the adhesion between the

circumferential side of the tablet and the die wall (influenced by material property

and contact surface area). Ejection force could therefore reflect the extent of powder

sticking to tool surface. The general trend of the ejection force values for the FFP

tablets was significantly higher than that of the FFRE and FFBE tablets over the

entire force range (5-10 kN) (p < 0.05), attributed to increased contact surface area

of FFP compact with the die wall and the resultant higher residual die wall pressure

(Fig. 45). Differences in ejection force between FFRE and FFBE tablets could be

considered marginal.

During the time course of tablet production, it was observed that the force required

to eject tablets varied greatly for the tablets produced using the FFP punches at the

lowest compaction force 5 kN, as demonstrated by the larger scatter of the ejection

forces in the profile. Ejection forces for the FFP punch started at lower magnitudes

and increased to levels higher than those observed for the tablets compacted using

the edged punches and varied drastically during the course of tableting.

175
A
400

350
Ejection force (N)

300

250

200

150

100
0 10 20 30 40 50
Tablet sequence
B 400

350
Ejection force (N)

300

250

200

150

100
0 10 20 30 40 50
Tablet sequence
C 400

350
Ejection force (N)

300

250

200

150

100
0 10 20 30 40 50
Tablet sequence

Figure 45. Ejection force profile during the time course of tablet production using
the compaction simulator for the FFP (○), FFRE (●) and FFBE (▲) punch face
geometries at compaction force 5.0 kN (A), 7.5 kN (B) and 10 kN (C) using an
ibuprofen-lactose formulation containing 1 %, w/w magnesium stearate.

176
The ejection force profiles showed less variability and reduced in overall force

magnitude at higher compaction force levels. The reduction in ejection force values

at increasing compaction forces validates the occurrence of more powder adhesion

to the die wall at lower compaction forces. In comparison to the FFP compacts, the

ejection forces for the radius and straight beveled punches remained relatively

constant over time at the compaction force levels studied

4.4.10 Summary

In this study, the influence of punch face configuration on powder adhesion tendency

onto the punch face under different compaction conditions was investigated. The use

of punch face configuration with a concavity (radius and bevel edge geometry) was

significantly advantageous as the presence of a cup reduced residual powder

adherence significantly. The outcome was similar to those reported by researchers

who have studied the relationship between punch face configuration and powder

adhesion [99-100]. The flat-face plain configuration reduced movement of particles

along the punch surface-compact boundaries. Absence of undulations on the flat-

face plain configuration ensured a relatively more homogenous stress distribution

across the punch face surfaces and this reduced the level of shear stress variations

across the compact surface. As a consequence, powder particles at the compact-tool

boundary experienced lower amount of disruptive sliding material shear during the

compaction process and this clearly facilitated particle fusion to the punch surface.

Powder adhesion data between the edged punch face configurations showed that the

overall differences were generally marginal and inconclusive, with similar degrees

of material sticking were observed. Process factors, apart from tooling configuration,

were observed to be more influential determinants of powder adhesion. Powder

adherence was observed to decrease with increasing compaction force, attributed to

177
the enhanced state of particulate consolidation at the higher force levels.

Acceleration of the compaction process was found to reduce punch sticking

tendency. Punch-material sticking increased significantly with an increase in

magnesium stearate concentration in the formulation, attributed to interaction

between magnesium stearate and ibuprofen particles which possibly reduced thermal

stability of ibuprofen.

The results from this experiment clearly showed the limitations of edge

modifications on flat-face punches. While the introduction of a curvature to the

punch face edge could strengthen the punch and improve tablet mechanical

properties, it may not provide a solution to tablet adhesion during tableting.

Manipulation of process factors such as tableting speed and compaction force could

be more effective in mitigating tablet adhesion incidences. The findings could be

useful to tablet manufacturers in making tooling considerations to overcome sticking

issues during production.

178
CHAPTER 5

CONCLUSION

179
5. CONCLUSION

In this thesis, the influence of tablet compaction tooling configuration on compaction

process and the resultant effect on tablet physical quality was investigated. Tableting

punch head and punch face configurations are often reworked by tooling

manufacturers for optimal tableting performance. Changes to punch head

configuration impacts different phases of the compaction process, particularly the

dwell phase.

In the first part of the study, the effect of formulation variables on dwell phase

compact deformation was better understood. Tableting formulations possessing

different densification characteristics were employed and reaction of compaction

systems to changes in dwell phase was studied. Plastic flow during the period of

constant compact strain (dwell phase) was observed in the experimented compact

systems, albeit at different extents. Compact systems containing plastically

deforming components reacted positively to dwell time increases; more significant

plastic flow was observed with dwell extension. Rate of stress decay was negatively

influenced by reductions in compact internal porosity caused by extensive

deformation and strain hardening of the plastic components. While dwell time

extension generally improved physical quality of tablets produced from more plastic

compact systems, the changes in tablet strength were however not consistent with

dwell changes and this could be indicative of time dependent bond reinforcement

process in compacts, showing also the viscoelastic character during the dwell phase.

In the subsequent experimental studies, effect of punch head configurational (head

flat and radius) modifications on tableting performance during rotary production was

investigated. Tablet formulations with viscoelastic deformation characteristics were

utilized for high speed tableting. As the dwell period was adjusted via head

180
configurational modifications rather than changes to the compaction kinematics, the

findings provided fresh insights into the effect of dwell during high speed tableting.

The head flat was found to be a much more influential feature than the head radius

in impacting the compaction cycle and tableting outcomes. The presence of a

physical head flat on the punch head was found to be advantageous, as tablets

produced from punches with a flat head configuration consistently displayed better

mechanical properties than tablets made from punches that lacked the head flat. The

flat top on the punch head allowed a state of constant strain for the compacts to

undergo stress relief (greater plastic flow). Tableting performance was however

greatly influenced by inherent deformation nature of the tableting materials and

process factors such as loading rate and compaction force, conditions that facilitated

the build-up of elasticity in the tablet matrix. Benefits of dwell time extension was

limited in more elastic compact systems, particularly at high compaction forces. Use

of a small compaction roll for tablet production decreased the compaction cycle time

that could potentially increase the contribution of dwell duration by proportion and

consequently enhanced stress relaxation in the compact but was insufficient to

counteract the rise in elasticity in the compact due to the higher loading rate during

the consolidation phase.

In the subsequent parts of this study, effect of edge modifications of flat-face

punches on the tableting process were studied. The quality of tablets produced from

different flat-face punch configurations have not been reported in literature, the data

from the next two parts of this study provided new knowledge on the tableting

performance of flat-face punch configurations under various compaction conditions.

Flat-face punch configurations with different edge geometries were used for tablet

production under different compaction conditions. Modification of the angled edge

181
of the bevel face (FFBE configuration) from straight to a curved edge with a radius

(FFRE configuration), enabled deeper punch penetration in the die cavity during the

compaction cycle, bringing about greater compact densification. As a result, tablets

produced from the FFRE face configuration consistently displayed better physical

quality; higher tensile strength and lower capping tendency. Improved die fill

packing by the radius edge geometry increased interparticulate bond formation and

helped to dissipate destructive elasticity within the compact, consequently reduced

tablet expansion during the decompression phase. The positive impact of face edge

modifications of flat-face punches was also more noticeable at a higher turret speed.

The application of the precompaction force along with dwell time extension

amplified the tableting performance of FFRE face configuration to a greater extent

when compared to the FFBE face configuration, attributed to the enhanced packing

efficiency at both precompaction and main compaction stages.

In the final part of the study, the impact of face configurational modification on

tablet material sticking tendency was studied using flat-face plain, bevel and radius

edged face configurations. The use of punch face configuration with a concavity

(radius and bevel edge geometry) was shown to be advantageous as the presence of

a cup reduced residual powder adherence significantly. Flattened surface of the flat-

face plain (FFP) configuration restricted movement of particles at the powder bed

boundaries and within the compact matrix to a greater extent than the other

experimented punch face configurations. Modification of the angled edge of the

FFBE geometry to a curved radius edge did not improve powder sticking to punch

surfaces, indicating equivalent tableting performance in terms of tablet material

adhesion tendency.

182
Thus, through the studies conducted in this study, design modifications to tableting

punch head and face proved to be influential factors in determining tableting

outcomes. The findings provided a better understanding on the advantages and

limitations of punch head and face configurational changes with respect to tablet

properties and may aid in making processing considerations for a successful

tableting program. The findings could also facilitate the design of better tablet

compaction tools.

Finally, two research areas with potential scientific impact are identified for future

investigation. First, a comprehensive evaluation of punch face configuration on the

tableting process should also include other tablet shape categories e.g. capsule and

oval shaped configurations. The most challenging aspect will probably be the need

to standardize the tablet characterization procedure, which may vary with different

tablet shapes. Second, the effect of tooling specialty metal and coating on tablet

quality is a relatively unexplored area of research. It is generally understood that

certain types of specialty metal and coating, usually of the chromium variant, could

be beneficial in reducing sticking tendency. However, benefits of such tooling

modifications have yet to be proven in a scientific approach.

183
CHAPTER 6

REFERENCES

184
6. REFERENCES

[1] M. Turkoglu, A. Sakr, Tablet dosage forms, in: G.S. Banker, C.T. Rhodes,
ed(s), Modern pharmaceutics, Fourth edition, New York: Marcel Dekker,
2009, pp. 481-497.
[2] J.F. Skaftason, T. Johannesson, Tablets and tablet production - with special
reference to Icelandic conditions, Laeknabladid, 99 (2013) 197-202.
[3] S. Mohan, Compression physics of pharmaceutical powders: a review,
International Journal of Pharmaceutical Sciences and Research, 3 (2012)
1580-1592.
[4] M. Çelik, Compaction of multiparticulate oral dosage forms, in: I. Ghebre-
Sellassie, ed(s), Multiparticulate oral drug delivery, New York: Marcel
Dekker, 1994, pp. 181-215.
[5] E.N. Hiestand, J.E. Wells, C.B. Peot, J.F. Ochs, Physical processes of
tableting, Journal of Pharmaceutical Sciences, 66 (1977) 510-519.
[6] B.V. Derjaguin, I.I. Abrikosova, E.M. Lifshitz, Direct measurement of
molecular attraction between solids separated by a narrow gap, Progress in
Surface Science, 40 (1992) 77-82.
[7] M. Duberg, C. Nyström, Studies on direct compression of tablets XVII.
Porosity—pressure curves for the characterization of volume reduction
mechanisms in powder compression, Powder Technology, 46 (1986) 67-75.
[8] D. Sixsmith, The compression characteristics of microcrystalline cellulose
powders, Journal of Pharmacy Pharmacology, 34 (1982) 345-346.
[9] M.J. Adams, M.A. Mullier, J.P.K. Seville, Agglomerate strength
measurement using a uniaxial confined compression test, Powder
Technology, 78 (1994) 5-13.
[10] S.A. Altaf, S.W. Hoag, Deformation of the Stokes B2 rotary tablet press:
quantitation and influence on tablet compaction, Journal of Pharmaceutical
Sciences, 84 (1995) 337-343.
[11] P.C. Schmidt, R. Herzog, Calcium phosphates in pharmaceutical tableting,
Pharmacy World and Science, 15 (1993) 105-115.
[12] E.T. Cole, J.E. Rees, J.A. Hersey, Relations between compaction data for
some crystalline pharmaceutical materials, Pharmaceutica acta Helvetiae, 50
(1975) 28-32.
[13] K. Van der Voort Maarschalk, K. Zuurman, H. Vromans, G.K. Bolhuis, C.F.
Lerk, Stress relaxation of compacts produced from viscoelastic materials,
International Journal of Pharmaceutics, 151 (1997) 27-34.
[14] K. van der Voort Maarschalk, K. Zuurman, H. Vromans, G.K. Bolhuis, C.F.
Lerk, Porosity expansion of tablets as a result of bonding and deformation of
particulate solids, International Journal of Pharmaceutics, 140 (1996) 185-
193.
[15] J.M. Sonnergaard, Quantification of the compactibility of pharmaceutical
powders, European Journal of Pharmaceutics and Biopharmaceutics, 63
(2006) 270-277.
[16] B. Johansson, M. Wikberg, R. Ek, G. Alderborn, Compression behaviour and
compactability of microcrystalline cellulose pellets in relationship to their
pore structure and mechanical properties, International Journal of
Pharmaceutics, 117 (1995) 57-73.
[17] P.J. Denny, Compaction equations: a comparison of the Heckel and Kawakita
equations, Powder Technology, 127 (2002) 162-172.

185
[18] R.J. Roberts, R.C. Rowe, Brittle/ductile behaviour in pharmaceutical
materials used in tabletting, International Journal of Pharmaceutics, 36 (1987)
205-209.
[19] J.S.M. Garr, M.H. Rubinstein, The effect of rate of force application on the
properties of microcrystalline cellulose and dibasic calcium phosphate
mixtures, International Journal of Pharmaceutics, 73 (1991) 75-80.
[20] F. Kiekens, A. Debunne, C. Vervaet, L. Baert, F. Vanhoutte, I. Van Assche,
F. Menard, J.P. Remon, Influence of the punch diameter and curvature on the
yield pressure of MCC-compacts during Heckel analysis, European Journal
of Pharmaceutical Sciences, 22 (2004) 117-126.
[21] J.M. Sonnergaard, A critical evaluation of the Heckel equation, International
Journal of Pharmaceutics, 193 (1999) 63-71.
[22] D.W. Danielson, W.T. Morehead, E.G. Rippie, Unloading and
postcompression viscoelastic stress versus strain behavior of pharmaceutical
solids, Journal of Pharmaceutical Sciences, 72 (1983) 342-345.
[23] S. Shlanta, G. Milosovich, Compression of pharmaceutical powders I.Theory
and instrumentation, Journal of Pharmaceutical Sciences, 53 (1964) 562-564.
[24] S.T. David, L.L. Augsburger, Plastic flow during compression of directly
compressible fillers and its effect on tablet strength, Journal of Pharmaceutical
Sciences, 66 (1977) 155-159.
[25] M. Rehula, R. Adamek, V. Spacek, Stress relaxation study of fillers for
directly compressed tablets, Powder Technology, 217 (2012) 510-515.
[26] L. Casahoursat, G. Lemagnen, D. Larrouture, The use of stress relaxation
trials to characterize tablet capping, Drug Development and Industrial
Pharmacy, 14 (1988) 2179-2199.
[27] J.E. Rees, P.J. Rue, Time-dependent deformation of some direct compression
excipients, Journal of Pharmacy Pharmacology, 30 (1978) 601-607.
[28] F. Ebba, P. Piccerelle, P. Prinderre, D. Opota, J. Joachim, Stress relaxation
studies of granules as a function of different lubricants, European Journal of
Pharmaceutics and Biopharmaceutics, 52 (2001) 211-220.
[29] W.R. Vezin, H.M. Pang, K.A. Khan, S. Malkowska, The effect of
precompression in a rotary machine on tablet strength, Drug Development and
Industrial Pharmacy, 9 (1983) 1465-1474.
[30] M. Celik, Pharmaceutical powder compaction technology, Informa
Healthcare, New York, 2011.
[31] S.K. Dwivedi, R.J. Oates, A.G. Mitchell, Peak offset times as an indication of
stress relaxation during tableting on a rotary tablet press, Journal of Pharmacy
Pharmacology, 43 (1991) 673-678.
[32] P.C. Schmidt, P.J. Vogel, Force-time-curves of a modern rotary tablet
machine I. Evaluation techniques and characterization of deformation
behaviour of pharmaceutical substances, Drug Development and Industrial
Pharmacy, 20 (1994) 921-934.
[33] P.J. Vogel, P.C. Schmidt, Force-time curves of a modern rotary tablet machine
II. Influence of compression force and tableting speed on the deformation
mechanisms of pharmaceutical substances, Drug Development and Industrial
Pharmacy, 19 (1993) 1917-1930.
[34] P.C. Schmidt, M. Leitritz, Compression force/time-profiles of
microcrystalline cellulose, dicalcium phosphate dihydrate and their binary
mixtures—a critical consideration of experimental parameters, European
Journal of Pharmaceutics and Biopharmaceutics, 44 (1997) 303-313.

186
[35] P. Konkel, J.B. Mielck, Associations of parameters characterizing the time
course of the tabletting process on a reciprocating and on a rotary tabletting
machine for high-speed production, European Journal of Pharmaceutics and
Biopharmaceutics, 45 (1998) 137.
[36] M. Leitritz, M. Krumme, P.C. Schmidt, Force‐time curves of a rotary tablet
press. Interpretation of the compressibility of a modified starch containing
various amounts of moisture, Journal of Pharmacy and Pharmacology, 48
(1996) 456-462.
[37] O. Antikainen, J. Yliruusi, Determining the compression behaviour of
pharmaceutical powders from the force–distance compression profile,
International Journal of Pharmaceutics, 252 (2003) 253-261.
[38] G. Ragnarsson, J. Sjogren, Work of friction and net work during compaction,
Journal of Pharmacy and Pharmacology, 35 (1983) 201-204.
[39] H. Nakamura, Y. Sugino, S. Watano, In-die evaluation of capping tendency
of pharmaceutical tablets using force-displacement curve and stress relaxation
parameter, Chemical & Pharmaceutical Bulletin (Tokyo), 60 (2012) 772-777.
[40] A. Nokhodchi, M.H. Rubinstein, H. Larhrib, J.C. Guyot, The effect of
moisture content on the energies involved in the compaction of ibuprofen,
International Journal of Pharmaceutics, 120 (1995) 13-20.
[41] H. Takeuchi, S. Nagira, H. Yamamoto, Y. Kawashima, Die wall pressure
measurement for evaluation of compaction property of pharmaceutical
materials, International Journal of Pharmaceutics, 274 (2004) 131-138.
[42] G.F. Palmieri, E. Joiris, G. Bonacucina, M. Cespi, A. Mercuri, Differences
between eccentric and rotary tablet machines in the evaluation of powder
densification behaviour, International Journal of Pharmaceutics, 298 (2005)
164-175.
[43] P. Anbalagan, C.V. Liew, P.W.S. Heng, Role of dwell on compact
deformation during tableting: an overview, Journal of Pharmaceutical
Investigation, 47 (2017) 173-181.
[44] K.M. Picker, The 3-D model: Comparison of parameters obtained from and
by simulating different tableting machines, AAPS PharmSciTech, 4 (2003)
55-61.
[45] C.E. Ruegger, M. Çelik, The influence of varying precompaction and main
compaction profile parameters on the mechanical strength of compacts,
Pharmaceutical Development and Technology, 5 (2000) 495-505.
[46] Courtoy, GEA - Courtoy tabletting equipment, 2014 [Accessed 11 December
2016]. Available from: http://www.s3process.co.uk/solid-dosage/gea-
courtoy/tabletting
[47] Korsch, High speed, fully automated production, 2016 [Accessed 11
December 2016]. Available from:
http://www.korsch.de/en/applications/pharmaceutical-industry/production
[48] S.D. Bateman, M.H. Rubinstein, R.C. Rowe, R.J. Roberts, P. Drew, A.Y.K.
Ho, A comparative investigation of compression simulators, International
Journal of Pharmaceutics, 49 (1989) 209-212.
[49] F. Michaut, V. Busignies, C. Fouquereau, B.H. de Barochez, B. Leclerc, P.
Tchoreloff, Evaluation of a rotary tablet press simulator as a tool for the
characterization of compaction properties of pharmaceutical products, Journal
of Pharmaceutical Sciences, 99 (2010) 2874-2885.
[50] M. Çelik, K. Marshall, Use of a compaction simulator system in tabletting
research, Drug Development and Industrial Pharmacy, 15 (1989) 759-800.

187
[51] J.T. Fell, J.M. Newton, Determination of tablet strength by the diametral-
compression test, Journal of Pharmaceutical Sciences, 59 (1970) 688-691.
[52] S.T. David, L.L. Augsburger, Flexure test for determination of tablet tensile
strength, Journal of Pharmaceutical Sciences, 63 (1974) 933-936.
[53] C. Nystrom, K. Malmqvist, J. Mazur, W. Alex, A.W. Holzer, Measurement
of axial and radial tensile strength of tablets and their relation to capping, Acta
pharmaceutica Suecica, 15 (1978) 226-232.
[54] A. Adam, L. Schrimpl, P.C. Schmidt, Factors influencing capping and
cracking of mefenamic acid tablets, Drug Development and Industrial
Pharmacy, 26 (2000) 489-497.
[55] I. Akseli, A. Stecula, X. He, N. Ladyzhynsky, Quantitative correlation of the
effect of process conditions on the capping tendencies of tablet formulations,
Journal of Pharmaceutical Sciences, (2014).
[56] A. Belič, I. Škrjanc, D.Z. Božič, R. Karba, F. Vrečer, Minimisation of the
capping tendency by tableting process optimisation with the application of
artificial neural networks and fuzzy models, European Journal of
Pharmaceutics and Biopharmaceutics, 73 (2009) 172-178.
[57] E.N. Hiestand, Principles, tenets and notions of tablet bonding and
measurements of strength, European Journal of Pharmaceutics and
Biopharmaceutics, 44 (1997) 229-242.
[58] E.N. Hiestand, D.P. Smith, Indices of tableting performance, Powder
Technology, 38 (1984) 145-159.
[59] I. Akseli, N. Ladyzhynsky, J. Katz, X. He, Development of predictive tools to
assess capping tendency of tablet formulations, Powder Technology, 236
(2013) 139.
[60] K.K. Lam, J.M. Newton, Influence of particle size on the adhesion behaviour
of powders, after application of an initial press-on force, Powder Technology,
73 (1992) 117-125.
[61] S. Paul, L.J. Taylor, B. Murphy, J. Krzyzaniak, N. Dawson, M.P. Mullarney,
P. Meenan, C.C. Sun, Mechanism and kinetics of punch sticking of
pharmaceuticals, Journal of Pharmaceutical Sciences, 106 (2017) 151-158.
[62] J.J. Wang, T. Li, S.D. Bateman, R. Erck, K.R. Morris, Modeling of adhesion
in tablet compression I. Atomic force microscopy and molecular simulation,
Journal of Pharmaceutical Sciences, 92 (2003) 798-814.
[63] J.J. Wang, M.A. Guillot, S.D. Bateman, K.R. Morris, Modeling of adhesion
in tablet compression II. Compaction studies using a compaction simulator
and an instrumented tablet press, Journal of Pharmaceutical Sciences, 93
(2004) 407-417.
[64] S.W. Booth, J.M. Newton, Experimental investigation of adhesion between
powders and surfaces, Journal of Pharmacy and Pharmacology, 39 (1987)
679-684.
[65] M. Eriksson, G. Alderborn, The effect of particle fragmentation and
deformation on the interparticulate bond formation process during powder
compaction, Pharmaceutical Research, 12 (1995) 1031-1039.
[66] R.J. Roberts, R.C. Rowe, The compaction of pharmaceutical and other model
materials - a pragmatic approach, Chemical Engineering Science, 42 (1987)
903-911.
[67] C. Nyström, G. Alderborn, M. Duberg, P.G. Karehill, Bonding surface area
and bonding mechanism-two Important factors for the understanding of
powder comparability, Drug Development and Industrial Pharmacy, 19
(1993) 2143-2196.

188
[68] M. Rahmouni, V. Lenaerts, D. Massuelle, E. Doelker, J.C. Leroux, Influence
of physical parameters and lubricants on the compaction properties of
granulated and non-granulated cross-linked high amylose starch, Chemical &
Pharmaceutical Bulletin (Tokyo), 50 (2002) 1155-1162.
[69] B. van Veen, G.K. Bolhuis, Y.S. Wu, K. Zuurman, H.W. Frijlink, Compaction
mechanism and tablet strength of unlubricated and lubricated (silicified)
microcrystalline cellulose, European Journal of Pharmaceutics and
Biopharmaceutics, 59 (2005) 133-138.
[70] M.J. Mollan, M. Çelik, The effects of lubrication on the compaction and post-
compaction properties of directly compressible maltodextrins, International
Journal of Pharmaceutics, 144 (1996) 1-9.
[71] M. Roberts, J.L. Ford, P.H. Rowe, A.M. Dyas, G.S. MacLeod, J.T. Fell, G.W.
Smith, Effect of lubricant type and concentration on the punch tip adherence
of model ibuprofen formulations, Journal of Pharmacy and Pharmacology, 56
(2004) 299-305.
[72] F. Waimer, M. Krumme, P. Danz, U. Tenter, P.C. Schmidt, A novel method
for the detection of sticking of tablets, Pharmaceutical Development and
Technology, 4 (1999) 359-367.
[73] L. Shi, C.C. Sun, Overcoming poor tabletability of pharmaceutical crystals by
surface modification, Pharmaceutical Research, 28 (2011) 3248-3255.
[74] K.K. Lam, J.M. Newton, Investigation of applied compression on the
adhesion of powders to a substrate surface, Powder Technology, 65 (1991)
167-175.
[75] K. Kakimi, T. Niwa, K. Danjo, Influence of compression pressure and
velocity on tablet sticking, Chemical & Pharmaceutical Bulletin (Tokyo), 58
(2010) 1565-1568.
[76] R.J. Roberts, R.C. Rowe, The effect of punch velocity on the compaction of a
variety of materials, Journal of Pharmacy and Pharmacology, 37 (1985) 377-
384.
[77] R.J. Roberts, R.C. Rowe, The effect of the relationship between punch
velocity and particle size on the compaction behaviour of materials with
varying deformation mechanisms, Journal of Pharmacy and Pharmacology,
38 (1986) 567-571.
[78] N. Armstrong, L.P. Palfrey, The effect of machine speed on the consolidation
of four directly compressible tablet diluents, Journal of Pharmacy and
Pharmacology, 41 (1989) 149-151.
[79] R. Ishino, H. Yoshino, Y. Hirakawa, K. Noda, Influence of tabletting speed
on compactibility and compressibility of two direct compressible powders
under high speed compression, Chemical & pharmaceutical bulletin, 38
(1990) 1987.
[80] J.S.M. Garr, M.H. Rubinstein, An investigation into the capping of
paracetamol at increasing speeds of compression, International Journal of
Pharmaceutics, 72 (1991) 117-122.
[81] P.V. Marshall, P. York, J.Q. Maclaine, An investigation of the effect of the
punch velocity on the compaction properties of ibuprofen, Powder
Technology, 74 (1993) 171-177.
[82] O.F. Akande, M.H. Rubinstein, P.H. Rowe, J.L. Ford, Effect of compression
speeds on the compaction properties of a 1:1 paracetamol–microcrystalline
cellulose mixture prepared by single compression and by combinations of pre-
compression and main-compression, International Journal of Pharmaceutics,
157 (1997) 127-136.

189
[83] C.E. Ruegger, M. Çelick, The effect of compression and decompression speed
on the mechanical strength of compacts, Pharmaceutical Development and
Technology, 5 (2000) 485-494.
[84] H.A. Garekani, J.L. Ford, M.H. Rubinstein, A.R. Rajabi-Siahboomi, Effect of
compression force, compression speed, and particle size on the compression
properties of paracetamol, Drug Development and Industrial Pharmacy, 27
(2001) 935-942.
[85] H. Larhrib, J.I. Wells, M.H. Rubinstein, Compressing polyethylene glycols:
The effect of compression pressure and speed, International Journal of
Pharmaceutics, 147 (1997) 199-205.
[86] A. Nokhodchi, J.L. Ford, P.H. Rowe, M.H. Rubinstein, The effects of
compression rate and force on the compaction properties of different viscosity
grades of hydroxypropylmethylcellulose, International Journal of
Pharmaceutics, 129 (1996) 21-31.
[87] M. Roberts, J.L. Ford, G.S. MacLeod, J.T. Fell, G.W. Smith, P.H. Rowe,
Effects of surface roughness and chrome plating of punch tips on the sticking
tendencies of model ibuprofen formulations, Journal of Pharmacy and
Pharmacology, 55 (2003) 1223-1228.
[88] D. Natoli, Tooling for pharmaceutical processing, in: L.L. Augsburger, S.W.
Hoag, ed(s), Pharmaceutical dosage forms - Tablets, Third edition, New York:
Informa Healthcare USA, 2008, pp. 1165-1212.
[89] American Pharmacists Association, Tableting Specification Manual,
American Pharmacists Association, 2006.
[90] L.L. Augsburger, S.W. Hoag, Pharmaceutical dosage forms: tablets, Informa
Healthcare USA, New York, 2008.
[91] B. Eiliazadeh, K. Pitt, B. Briscoe, Effects of punch geometry on powder
movement during pharmaceutical tabletting processes, International Journal
of Solids and Structures, 41 (2004) 5967-5977.
[92] D. Sixsmith, D. McCluskey, The effect of punch tip geometry on powder
movement during the tableting process, Journal of Pharmacy and
Pharmacology, 33 (1981) 79-81.
[93] A. Djemai, I.C. Sinka, NMR imaging of density distributions in tablets,
International Journal of Pharmaceutics, 319 (2006) 55-62.
[94] V. Busignies, B. Leclerc, P. Porion, P. Evesque, G. Couarraze, P. Tchoreloff,
Quantitative measurements of localized density variations in cylindrical
tablets using X-ray microtomography, European Journal of Pharmaceutics
and Biopharmaceutics 64 (2006) 38-50.
[95] H. Diarra, V. Mazel, V. Busignies, P. Tchoreloff, Investigating the effect of
tablet thickness and punch curvature on density distribution using finite
elements method, International Journal of Pharmaceutics, 493 (2015) 121-
128.
[96] J.M. Newton, I. Haririan, F. Podczeck, The influence of punch curvature on
the mechanical properties of compacted powders, Powder Technology, 107
(2000) 79-83.
[97] D.C. Andersson, P. Larsson, A. Cadario, P. Lindskog, On the influence from
punch geometry on the stress distribution at powder compaction, Powder
Technology, 202 (2010) 78-88.
[98] M.S. Kadiri, A. Michrafy, The effect of punch's shape on die compaction of
pharmaceutical powders, Powder Technology, 239 (2013) 467-477.
[99] F. Waimer, M. Krumme, P. Danz, U. Tenter, P.C. Schmidt, The influence of
engravings on the sticking of tablets. Investigations with an instrumented

190
upper punch, Pharmaceutical Development and Technology, 4 (1999) 369-
375.
[100] M. Roberts, J.L. Ford, G.S. MacLeod, J.T. Fell, G.W. Smith, P.H. Rowe, A.M.
Dyas, Effect of punch tip geometry and embossment on the punch tip
adherence of a model ibuprofen formulation, Journal of Pharmacy and
Pharmacology, 56 (2004) 947-950.
[101] R.S. Ransing, D.T. Gethin, A.R. Khoei, P. Mosbah, R.W. Lewis, Powder
compaction modelling via the discrete and finite element method, Materials
& Design, 21 (2000) 263-269.
[102] D.T. Gethin, R.S. Ransing, R.W. Lewis, M. Dutko, A.J.L. Crook, Numerical
comparison of a deformable discrete element model and an equivalent
continuum analysis for the compaction of ductile porous material, Computers
and Structures, 79 (2001) 1287-1294.
[103] C.Y. Wu, O.M. Ruddy, A.C. Bentham, B.C. Hancock, S.M. Best, J.A. Elliott,
Modelling the mechanical behaviour of pharmaceutical powders during
compaction, Powder Technology, 152 (2005) 107-117.
[104] Y. Hayashi, T. Miura, T. Shimada, Y. Onuki, Y. Obata, K. Takayama,
Prediction of tablet characteristics from residual stress distribution estimated
by the finite element method, Journal of Pharmaceutical Sciences, 102 (2013)
3678-3686.
[105] J.C. Cunningham, I.C. Sinka, A. Zavaliangos, Analysis of tablet compaction.
I. Characterization of mechanical behavior of powder and powder/tooling
friction, Journal of Pharmaceutical Sciences, 93 (2004) 2022-2039.
[106] I.C. Sinka, J.C. Cunningham, A. Zavaliangos, The effect of wall friction in
the compaction of pharmaceutical tablets with curved faces: a validation study
of the Drucker–Prager Cap model, Powder Technology, 133 (2003) 33-43.
[107] I.C. Sinka, J.C. Cunningham, A. Zavaliangos, Analysis of tablet compaction.
II. Finite element analysis of density distributions in convex tablets, Journal
of Pharmaceutical Sciences, 93 (2004) 2040-2053.
[108] T. Sinha, R. Bharadwaj, J.S. Curtis, B.C. Hancock, C. Wassgren, Finite
element analysis of pharmaceutical tablet compaction using a density
dependent material plasticity model, Powder Technology, 202 (2010) 46-54.
[109] C.Y. Wu, B.C. Hancock, A. Mills, A.C. Bentham, S.M. Best, J.A. Elliott,
Numerical and experimental investigation of capping mechanisms during
pharmaceutical tablet compaction, Powder Technology, 181 (2008) 121-129.
[110] L.H. Han, J.A. Elliott, A.C. Bentham, A. Mills, G.E. Amidon, B.C. Hancock,
A modified Drucker-Prager Cap model for die compaction simulation of
pharmaceutical powders, International Journal of Solids and Structures, 45
(2008) 3088-3106.
[111] H. Sato, T. Miura, H. Furuichi, FEM Analysis of tablet hardness and punch
strength using ANSYS software, Journal of Computational Science and
Technology, 5 (2011) 120-133.
[112] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of
particulate systems: A review of major applications and findings, Chemical
Engineering Science, 63 (2008) 5728-5770.
[113] A. Mehrotra, B. Chaudhuri, A. Faqih, M.S. Tomassone, F.J. Muzzio, A
modeling approach for understanding effects of powder flow properties on
tablet weight variability, Powder Technology, 188 (2009) 295-300.
[114] M. Khanal, W. Schubert, J. Tomas, DEM simulation of diametrical
compression test on particle compounds, Granular Matter, 7 (2005) 83-90.

191
[115] Y. Sheng, C.J. Lawrence, B.J. Briscoe, C. Thornton, Numerical studies of
uniaxial powder compaction process by 3D DEM, Engineering Computations,
21 (2004) 304-317.
[116] G. Frenning, An efficient finite/discrete element procedure for simulating
compression of 3D particle assemblies, Computer Methods in Applied
Mechanics and Engineering, 197 (2008) 4266-4272.
[117] S.M. Siiriä, O. Antikainen, J. Heinämäki, J. Yliruusi, 3D Simulation of
internal tablet strength during tableting, AAPS PharmSciTech, 12 (2011) 593-
603.
[118] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of
particulate systems: Theoretical developments, Chemical Engineering
Science, 62 (2007) 3378-3396.
[119] O.F. Akande, J.L. Ford, P.H. Rowe, M.H. Rubinstein, The effects of lag-time
and dwell-time on the compaction properties of 1:1
paracetamol/microcrystalline cellulose tablets prepared by pre-compression
and main compression, Journal of Pharmacy and Pharmacology, 50 (1998)
19.
[120] C.K. Tye, C. Sun, G.E. Amidon, Evaluation of the effects of tableting speed
on the relationships between compaction pressure, tablet tensile strength, and
tablet solid fraction, Journal of Pharmaceutical Sciences, 94 (2005) 465-472.
[121] K. Hyun Jo, V. Gopi, Original Articles : Compaction simulator study on
pectin introducing dwell Time, International Journal of Pharmaceutical
Investigation, 35 (2005) 243-247.
[122] S. Sarkar, S.M. Ooi, C.V. Liew, P.W. Heng, Influence of rate of force
application during compression on tablet capping, Journal of Pharmaceutical
Sciences, 104 (2015) 1319-1327.
[123] S. Aoki, J. Uchiyama, M. Ito, Development of new punch shape to replicate
scale‐up issues in laboratory tablet press II: A new design of punch head to
emulate consolidation and dwell times in commercial tablet press, Journal of
Pharmaceutical Sciences, 103 (2014) 1921-1927.
[124] G. Thoorens, F. Krier, B. Leclercq, B. Carlin, B. Evrard, Microcrystalline
cellulose, a direct compression binder in a quality by design environment - A
review, International Journal of Pharmaceutics, 473 (2014) 64-72.
[125] H.Y. Saw, C.E. Davies, A.H.J. Paterson, J.R. Jones, Correlation between
powder flow properties measured by shear testing and Hausner ratio, Procedia
Engineering, 102 (2015) 218-225.
[126] L.E. Dahlinder, M. Johansson, J. Sjögren, Comparison of methods for
evaluation of flow properties of powders and granulates, Drug Development
and Industrial Pharmacy, 8 (1982) 455-461.
[127] C. Cahyadi, L.W. Chan, P.W.S. Heng, The reality of in-line tablet coating,
Pharmaceutical Development and Technology, 18 (2013) 2-16.
[128] R.V. Haware, I. Tho, A. Bauer-Brandl, Evaluation of a rapid approximation
method for the elastic recovery of tablets, Powder Technology, 202 (2010)
71-77.
[129] L. Maganti, M. Çelik, Compaction studies on pellets I. Uncoated pellets,
International Journal of Pharmaceutics, 95 (1993) 29-42.
[130] P. Anbalagan, S. Sarkar, C.V. Liew, P.W.S. Heng, Influence of the punch
head design on the physical quality of tablets produced in a rotary press,
Journal of Pharmaceutical Sciences, 106 (2016) 356-365.

192
[131] P. Anbalagan, P.W.S. Heng, C.V. Liew, Tablet compression tooling – Impact
of punch face edge modification, International Journal of Pharmaceutics, 524
(2017) 373-381.
[132] M. Ucgul, J.M. Fielke, C. Saunders, 3D DEM tillage simulation: Validation
of a hysteretic spring (plastic) contact model for a sweep tool operating in a
cohesionless soil, Soil and Tillage Research, 144 (2014) 220-227.
[133] S. Paul, L.J. Taylor, B. Murphy, J.F. Krzyzaniak, N. Dawson, M.P.
Mullarney, P. Meenan, C.C. Sun, Powder properties and compaction
parameters that influence punch sticking propensity of pharmaceuticals,
International Journal of Pharmaceutics, 521 (2017) 374-383.
[134] D. Khossravi, W.T. Morehead, Consolidation mechanisms of pharmaceutical
solids: A multi-compression cycle approach, Pharmaceutical Research, 14
(1997) 1039-1045.
[135] S.R. Perumalla, L. Shi, C.C. Sun, Ionized form of acetaminophen with
improved compaction properties, CrystEngComm, 14 (2012) 2389-2390.
[136] S. Karki, T. Friscic, L. Fabian, P.R. Laity, G.M. Day, W. Jones, Improving
mechanical properties of crystalline solids by cocrystal formation: New
compressible forms of paracetamol, Advanced Materials, 21 (2009) 3905-
3909.
[137] C.C. Sun, Y.H. Kiang, On the identification of slip planes in organic crystals
based on attachment energy calculation, Journal of Pharmaceutical Sciences,
97 (2008) 3456-3461.
[138] E. Shotton, B.A. Obiorah, Effect of physical properties on compression
characteristics, Journal of Pharmaceutical Sciences, 64 (1975) 1213-1216.
[139] S. Abdel-Hamid, M. Koziolek, G. Betz, Study of radial die-wall pressure
during high speed tableting: effect of formulation variables, Drug
Development and Industrial Pharmacy, 38 (2012) 623-634.
[140] H. Takeuchi, S. Nagira, H. Yamamoto, Y. Kawashima, Die wall pressure
measurement for evaluation of compaction property of pharmaceutical
materials, International Journal of Pharmaceutics, 274 (2004) 131-138.
[141] A.P. Grima, P.W. Wypych, Development and validation of calibration
methods for discrete element modelling, Granular Matter, 13 (2011) 127-132.

193
CHAPTER 7

LIST OF PUBLICATIONS
AND PRESENTATIONS

194
7. LIST OF PUBLICATIONS AND PRESENTATIONS

International Journals
1. Anbalagan P, Sarkar S, Liew CV, Heng PWS 2016. Influence of the punch
head design on the physical quality of tablets produced in a rotary press.
Journal of Pharmaceutical Sciences 106:356-365.

2. Anbalagan P, Liew CV, Heng PWS 2017. Role of dwell on tablet


compaction. Journal of Pharmaceutical Investigation 47(3):173-181.

3. Anbalagan P, Heng PWS, Liew CV 2017. Tablet compression tooling -


Impact of punch face edge modification. International Journal of
Pharmaceutics 524:373-381.

4. Desai PM, Anbalagan P, Heng PWS, Liew CV 2017. Evaluation of tablet


punch configuration on mitigating capping by a quality by design approach.
Drug Delivery and Translational Research (in press).

Oral presentations
1. Anbalagan P, Liew CV, Heng PWS 2016. Influence of tableting punch head
design on tablet compression in a rotary press: An experimental and
computational approach. Asian Graduate Congress on Pharmaceutical
Technology. Singapore.

2. Anbalagan P, Liew CV, Heng PWS 2016. Tablet compression tooling design
modifications to optimize tablet properties. Globalization Pharmaceutics
Education Network (GPEN) conference. Kansas, USA.
Poster presentations
1. Anbalagan P, Liew CV, Heng PWS 2014. Tablet compression tooling design
modifications for optimal tablet manufacture. International Society for
Pharmaceutical Engineering (ISPE) conference. Singapore.

2. Anbalagan P, Liew CV, Heng PWS 2015. Influence of punch geometry on


tablet capping during manufacturing: An experimental and computational
approach. 7th Asian Association of Schools of Pharmacy (AASP) conference.
Taipei, Taiwan.

3. Anbalagan P, Liew CV, Heng PWS 2016. Influence of tableting punch head
design on tablet compression in a rotary press: An experimental and
computational approach. American Association of Pharmaceutical Scientists
Annual Meeting and Exposition. Denver, USA.

195

Potrebbero piacerti anche