<<

Purdue University Purdue e-Pubs

Open Access Dissertations Theses and Dissertations

January 2016 DEVELOPMENT OF NOVEL DESIGN AND CONTROL APPROACHES FOR INTEGRATED OPERATION AND SYSTEMS Yang Yang Purdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_dissertations

Recommended Citation Yang, Yang, "DEVELOPMENT OF NOVEL DESIGN AND CONTROL APPROACHES FOR INTEGRATED CRYSTALLIZATION OPERATION AND SYSTEMS" (2016). Open Access Dissertations. 1234. https://docs.lib.purdue.edu/open_access_dissertations/1234

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Graduate School Form 30 Updated 

PURDUE UNIVERSITY GRADUATE SCHOOL Thesis/Dissertation Acceptance

This is to certify that the thesis/dissertation prepared

By Yang Yang

Entitled DEVELOPMENT OF NOVEL DESIGN AND CONTROL APPROACHES FOR INTEGRATED CRYSTALLIZATION OPERATION AND SYSTEMS

For the degree of Doctor of Philosophy

Is approved by the final examining committee:

Zoltan K. Nagy James D. Litster Chair Gintaras V. Reklaitis

Michael T. Harris

Christopher L. Burcham

To the best of my knowledge and as understood by the student in the Thesis/Dissertation Agreement, Publication Delay, and Certification Disclaimer (Graduate School Form 32), this thesis/dissertation adheres to the provisions of Purdue University’s “Policy of Integrity in Research” and the use of copyright material.

Approved by Major Professor(s): Zoltan K. Nagy

Approved by: Sangtae Kim 9/21/2016 Head of the Departmental Graduate Program Date

i

DEVELOPMENT OF NOVEL DESIGN AND CONTROL APPROACHES FOR

INTEGRATED CRYSTALLIZATION OPERATION AND SYSTEMS

A Dissertation

Submitted to the Faculty

of

Purdue University

by

Yang Yang

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

December 2016

Purdue University

West Lafayette, Indiana

ii

ACKNOWLEDGEMENTS

I am grateful to my advisor, Prof. Zoltan K. Nagy, for his guidance and encouragement on my research throughout the last four years. It was a really wonderful and unforgettable experience working with him. He encouraged me to follow my interest on research topics and try any ideas I have. He always stood behind me and supported my ideas. I greatly appreciate all his advices and criticism on my research, and the time and effort he spent on me during the endless discussions.

In addition, I would like to thank Prof. Gintaras Reklaitis. Prof. James Litster, Prof.

Michael Harris and Dr. Christopher Burcham for serving on my dissertation committee. I acknowledge their comments and suggestions on my work.

I also acknowledge Eli Lilly and Company for providing financial support and transfer units for the MSMPRC projects. I am grateful to Christopher Burcham, Daniel Jarmer and

Christopher Polster from Eli Lilly and Company for input on MSMPRC rig configuration and discussions.

I would like to thank Hillary Hewitson, Ernie Hillier, Jennifer Halsey and Ryan Arnett from Waters Corporation for providing the PATROL UPLC® system, instrument support and comments on the UPLC project. I am grateful to GlaxoSmithKline and Kalsec for the financial support. I acknowledge Justin Quon, Charles Goss and Rahn McKeown from

GlaxoSmithKline for discussions and suggestions on the UPLC project.

iii

I would also like to thank my colleagues David Acevedo, Andy Kowswara, Ramon Peña,

Kanjakha Pal, Jennifer Lu, Qinglin Su, Liangcheng Song, Qingqing Sun, Chuntao Zhang, and undergraduate students Yuqi Zhang, Tianyue Gao, Xuan Xie, Sixing Zhao, Amiirul

Sahrir for their support on my research projects.

Finally, I wish to say thank you to my parents and fiancée who always support and trust me!

iv

TABLE OF CONTENTS

Page LIST OF TABLES ...... ix LIST OF FIGURES ...... xi LIST OF ABBREVIATIONS ...... xxi LIST OF SYMBOLS ...... xxiv ABSTRACT ...... xxviii PUBLICATIONS ...... xxxi CHAPTER 1. INTRODUCTION ...... 1 1.1 Background ...... 1 1.2 Research Aims ...... 4 1.3 Research Contributions ...... 5 1.4 Thesis Structure ...... 7 CHAPTER 2. LITERATURE REVIEW ...... 9 2.1 Fundamentals of Pharmaceutical Crystallization ...... 9 2.2 Process Analytical Technology (PAT) ...... 12 2.3 Model-Based and Model-Free Control ...... 14 2.4 Batch and Continuous Crystallization ...... 17 2.5 Combined Cooling/Antisolvent Crystallization (CCAC) System ...... 20 2.6 Integrated Crystallization and Wet Milling Operation ...... 21 2.7 Characterization of Crystallization System in the Presence of Impurities ...... 23 2.8 Summary on Literature Review ...... 26 CHAPTER 3. MODEL-BASED SYSTEMATIC DESIGN APPROACH FOR UNSEEDED COMBINED COOLING/ANTISOLVENT CRYSTALLIZATION ...... 27 3.1 Introduction ...... 27

v

3.2 Mathematical Modeling ...... 28 3.3 Experimental Section ...... 33 3.4 Results and Discussion ...... 34 3.4.1 CCAC Design Through the Optimization of Operating Procedures With Constant Cooling and Antisolvent Addition Rates ...... 34 3.4.2 CCAC Design Using a Simulation-Based Operating Map ...... 40 3.4.3 Experimental Validation of the CCAC Design Approach Using the Simulated Operating Map ...... 44 3.4.4 Full optimization of nonlinear trajectory based design of CCAC systems ... 49 3.5 Conclusions ...... 56 CHAPTER 4. COMBINED COOLING/ANTISOLVENT CRYSTALLIZATION IN MIXED SUSPENSION MIXED PRODUCT REMOVAL CASCADE CRYSTALLIZERS: STEADY-STATE AND STARTUP OPTIMIZATION ...... 58 4.1 Introduction ...... 58 4.2 Mathematical Model ...... 59 4.3 Results and Discussion ...... 62 4.3.1 Steady-State Optimization ...... 62 4.3.2 Startup Optimization ...... 67 4.3.2.1 Startup Optimization via Using Different Initial Solutions ...... 67 4.3.2.2 Startup Optimization via Applying Dynamic Operating Profiles ...... 73

4.3.2.2.1 Scenario 1: Dynamic Antisolvent Addition Profiles With Minimum t99 as Objective ...... 74

4.3.2.2.2 Scenario 2: Dynamic Temperature Profiles With Minimum t99 as Objective...... 76

4.3.2.2.3 Scenario 3: Dynamic Combined Profiles With Minimum t99 as Objective...... 77

4.3.2.2.4 Scenario 4: Dynamic Combined Profiles With Minimum t95 as Objective...... 79 4.3.2.3 Startup Optimization via Model-Free Approach ...... 80 4.4 Conclusions ...... 81

vi

CHAPTER 5. ADVANCED CONTROL APPROACHES FOR COMBINED COOLING/ANTISOLVENT CRYSTALLIZATION IN CONTINUOUS MIXED SUSPENSION MIXED PRODUCT REMOVAL CASCADE CRYSTALLIZERS ...... 83 5.1 Introduction ...... 83 5.2 Mathematical Model ...... 85 5.3 Results and Discussion ...... 86 5.3.1 Attainable region of crystal size and yield ...... 86 5.3.2 Analysis of System Dynamics ...... 89 5.3.3 Feasibility of Decentralized Proportional-Integral-Derivative (PID) Control91 5.3.4 Feasibility of Nonlinear Model Predictive Control (NMPC) ...... 94 5.3.4.1 NMPC Problem Formulation ...... 94 5.3.4.2 Servo Control ...... 96 5.3.4.3 Regulatory Control ...... 101 5.4 Conclusions ...... 105 CHAPTER 6. INTEGRATED UPSTREAM AND DOWNSTREAM APPLICATION OF WET MILLING WITH CONTINUOUS MIXED SUSPENSION MIXED PRODUCT REMOVAL CRYSTALLIZATION ...... 107 6.1 Introduction ...... 107 6.2 Experimental Section ...... 108 6.2.1 Method 1: MSMPRC without Wet Mill ...... 108 6.2.2 Method 2: MSMPRC without Wet Mill Operated Downstream in a Recycle Loop...... 109 6.2.3 Method 3: MSMPRC without Wet Mill Operated Upstream as Seed Generator ...... 112 6.3 Results and Discussion ...... 112 6.4 Conclusions ...... 123 CHAPTER 7. AUTOMATED DIRECT NUCLEATION CONTROL IN CONTINUOUS MIXED SUSPENSION MIXED PRODUCT REMOVAL COOLING CRYSTALLIZATION ...... 124 7.1 Introduction ...... 124

vii

7.2 Experimental Section ...... 125 7.3 Continuous ADNC Methodology ...... 127 7.4 Results and Discussion ...... 129 7.4.1 Startup and Control of CSD ...... 129 7.4.1.1 Continuous ADNC in Single-Stage MSMPRC ...... 129 7.4.1.2 Continuous ADNC in Two-Stage MSMPRC ...... 133 7.4.2 Disturbance Suppression ...... 139 7.5 Conclusions ...... 142 CHAPTER 8. APPLICATION OF WET MILLING-BASED AUTOMATED DIRECT NUCLEATION CONTROL IN CONTINUOUS COOLING CRYSTALLIZATION PROCESSES...... 143 8.1 Introduction ...... 143 8.2 Experimental Section ...... 144 8.2.1 Upstream Application of WMADNC Approach (Process 1) ...... 144 8.2.2 Downstream Application of WMADNC Approach (Process 2) ...... 146 8.2.3 WMADNC Methodology in Continuous Cooling Crystallization ...... 148 8.3 Results and Discussion ...... 149 8.3.1 Upstream Application of the WMADNC Approach to Achieve Closed-Loop Controlled Primary Nucleation (Process 1) ...... 149 8.3.1.1 Startup and Steady-State Dynamics ...... 150 8.3.1.2 Set-Point Change and Control of CLD ...... 151 8.3.1.3 Disturbance Rejection ...... 154 8.3.2 Downstream Application of the WMADNC Approach to Achieve Closed- Loop Controlled Secondary Nucleation (Process 2) ...... 157 8.3.2.1 Startup and Steady-State Dynamics ...... 157 8.3.2.2 Set-Point Change and Control of CLD ...... 158 8.3.2.3 Disturbance Rejection ...... 160 8.4 Conclusions ...... 163

viii

CHAPTER 9. APPLICATION OF ULTRA-PERFORMANCE LIQUID AS AN ONLINE PROCESS ANALYTICAL TECHNOLOGY TOOL IN PHARMACEUTICAL CRYSTALLIZATION ...... 164 9.1 Introduction ...... 164 9.2 Experimental Section ...... 165 9.2.1 Materials ...... 165 9.2.2 Experimental Setup ...... 165 9.2.3 UPLC Methods ...... 167 9.3 Results and Discussion ...... 169 9.3.1 Online UPLC-Based Crystallization and Degradation Monitoring ...... 169 9.3.2 Online UPLC-Based Quick Calibration for UV/vis and Raman Spectroscopy...... 171 9.3.3 Online UPLC for Real-Time Product Purity Monitoring ...... 175 9.4 Conclusions ...... 177 CHAPTER 10. CONCLUSIONS AND FUTURE WORK ...... 179 10.1 Conclusions ...... 179 10.2 Future Work ...... 181 REFERENCES ...... 183 VITA ...... 201

ix

LIST OF TABLES

Table ...... Page Table 2.1 References about application of PAT in crystallization ...... 14

Table 3.1. Growth and nucleation kinetic parameters of aspirin in ethanol-water mixture measured by Lindenberg et al., 2009...... 31

Table 3.2. Solubility of Aspirin c* (g solute/kg solvent) from Lindenberg et al., 2009. .. 32

Table 3.3. Optimal cooling time, antisolvent feeding time and final crystal properties for

S1 and S2 obtained through optimization...... 37

Table 3.4. Comparison of the simultaneous case, optimal case of cooling first, and optimal case of antisolvent first through simulation...... 43

Table 3.5. Comparison of the suboptimal trajectories obtained by optimization and simulation...... 43

Table 3.6. Comparison of simulated mean size and experimental SWMCL for the corresponding cases for conditions S2...... 46

Table 3.7. Comparison of final product properties between suboptimal and optimal trajectories...... 56

Table 4.1. Optimal operating temperatures and antisolvent addition rates for cascade of

N 1, 2, 3, and the batch system...... 65

x

Table 4.2. Comparisons between the compositions, t99 and Waste ( t99 ) of the base case

and the optimal case, which uses min{t99 } as objective...... 71

Table 4.3. Comparisons between the compositions, t95 and waste amount of the base case

and the optimal case, which uses min{t95 as objective...... 72

Table 4.4. Comparisons between the performance of all scenarios and base cases...... 80

Table 5.1. Five possible different control approaches for two-stage MSMPR cascade crystallization systems...... 87

Table 5.2. Selected operating points for local linearization and RGA analysis...... 91

Table 5.3. RGAs and input-output pairings for selected operating points of the four proposed control methods...... 93

Table 5.4. Selected set-points to evaluate NMPC feasibility...... 97

Table 5.5. NMPC performance using the five proposed control methods...... 102

Table 5.6. Two scenarios for disturbance rejection evaluation. F0, c0 are feed solution flow rate and solute concentration...... 103

Table 6.1. Summary of experimental conditions, process and product qualities using methods 1, 2 and 3...... 114

Table 7.1. Operating conditions and ADNC parameters used in every experiment...... 129

Table 9.1. Manufacture and purity of materials used...... 165

Table 9.2. Isocratic UPLC methods used in this work for different compounds...... 168

Table 9.3. Fitted coefficients and errors of UV/vis and Raman quick calibration model for aspirin...... 174

xi

LIST OF FIGURES

Figure ...... Page Figure 2.1. Crystallization phase diagram...... 10

Figure 3.1. Solubility surface of aspirin in ethanol-water mixture. The circles are solubility data obtained from literature (Lindenberg et al., 2009)...... 32

Figure 3.2. Schematic of experimental setup...... 33

Figure 3.3. Schematic of CCAC process with constant cooling rate and antisolvent addition rate. t,,, t  t  t are cooling start time and end time, antisolvent start time and end time, 1 2 3 4 respectively...... 36

Figure 3.4. Suboptimal trajectories for S1, S2, and scenarios with other initial conditions.

...... 37

Figure 3.5. Suboptimal operating curves for (a) S1 and (b) S2...... 38

Figure 3.6. (a) Suboptimal supersaturation ratio and  ()t for S1 (solid line) and S2 0 (dotted line); (b) suboptimal concentration trajectories for S1 (solid line) and S2 (dotted line)...... 38

Figure 3.7. Schematic of (a) cooling first and (b) antisolvent first strategy. t1 , t2 , t3 and t4 are cooling start time and end time, antisolvent start time and end time, respectively. ... 40

xii

Figure 3.8. The mean size surfaces of cooling first and antisolvent first strategies for scenarios S1 and S2, respectively. Points (1)(2)(3)(4) are the corresponding optimum points. Point (0) is simultaneously cooling and adding antisolvent through 0 to 30 minutes, which is used as base case for comparison...... 42

Figure 3.9. Comparison between experimental mean size (SWMCL) and simulated mean

size ( Ln ), when applying (a) cooling first strategy or (b) antisolvent first strategy corresponding to the initial condition of S2. Point A: simultaneous cooling and antisolvent; point B: optimal of cooling first; point C: non-optimal of cooling first; point D: optimal of antisolvent first; point E: non-optimum of antisolvent first. The color bar and 'sim' stand for simulated results. The number on the map and 'exp' stand for experimental results. The variation for experimental result is 2σ (95% confidence interval)...... 45

Figure 3.10. Comparison between (a) experimental mean size (SWMCL) of cooling first

strategy and (b) antisolvent first strategy, (c) simulated mean size ( Ln ) of cooling first strategy and (d) antisolvent first strategy, respectively. Case A: simultaneous cooling and antisolvent; case B: optimal of cooling first; case C: non-optimal of cooling first; case D: optimal of antisolvent first; case E: non-optimal of antisolvent first...... 47

Figure 3.11. Normalized cumulative chord length distributions at the end of batch for each case of (a) cooling first strategy and (b) antisolvent first strategy. Case A: simultaneous cooling and antisolvent; case B: optimal of cooling first; case C: non-optimal of cooling first; case D: optimal of antisolvent first; case E: non-optimal of antisolvent first...... 48

xiii

Figure 3.12. Final crystals of case A (simultaneous cooling/ antisolvent), case B (optimal of cooling first), case C (non-optimal of cooling first), case D (optimal of antisolvent first), and case E (non-optimal of antisolvent first). All images have the same magnification. . 49

Figure 3.13. Optimal trajectories for S1opt, S2opt, and scenarios with other initial conditions.

...... 51

Figure 3.14. Optimal operating curves for (a) S1opt and (b) S2opt...... 52

Figure 3.15. (a) Optimal supersaturation ratio and  ()t for S1 (solid line) and S2 0 opt opt

(dotted line); (b) optimal concentration trajectories for S1opt (solid line) and S2opt (dotted line)...... 52

Figure 3.16. Optimal trajectories (bold solid line) vs. suboptimal trajectories (thin dash line) for S1 and S2...... 55

Figure 4.1. Schematic of performing CCAC in a multistage MSMPR cascade...... 62

Figure 4.2. Optimal trajectories of 1 stage (dotted line), 2 stages (dash-dotted line), 3 stages

(dash line) cascade system and batch (solid line) system, shown in (a) 3D phase diagram and (b) its 2D projection...... 66

Figure 4.3. Final crystal mean size under various solvent volume fractions and solute concentrations of the initial solution, for cascade that has different number of stages: (a)

N 1, (b) N  2 , (c) N  3...... 69

Figure 4.4. Mean size Ltn ( ) and Waste ( t99 ) (a) before and (b) after optimizing the

compositions of the initial solutions when using min t99 as objective...... 71

Figure 4.5. Mean size Ltn ( ) and Waste ( t95 ) (a) before and (b) after optimizing the

compositions of the initial solutions when using min{t95 } as objective...... 73

xiv

Figure 4.6. (a) Optimal dynamic antisolvent addition profiles, (b) corresponding mean size

Ltn ( ) and Waste ( t99 ) when using min t99 as objective...... 75

Figure 4.7. (a) Optimal dynamic temperature profiles, (b) corresponding mean size Ltn ( )

and Waste ( t99 ) when using min{t99 } as objective...... 77

Figure 4.8. (a) Optimal dynamic antisolvent addition profiles, (b) optimal dynamic

temperature profiles, (c) corresponding mean size Ltn ( ) and Waste ( t99 ) when using

min{t99 } as objective...... 78

Figure 4.9. (a) Optimal dynamic antisolvent addition profiles, (b) corresponding mean size

Ltn ( ) and Waste ( t95 ) when using min{t95 } as objective...... 79

Figure 5.1. Schematic representation of a hierarchical NMPC implementation structure.

FC, TC, LC are flow rate controller, temperature controller and level controller, respectively...... 84

Figure 5.2. Schematic representation of the continuous CCAC process in a two stage

MSMPR cascade crystallizer...... 86

Figure 5.3. Schematic representation of investigated multi-input multi-output system. .. 87

Figure 5.4. Attainable regions of crystal size and yield in the case of the proposed five control methods. Operating points 1–7 were used in decentralized PID control analysis.

Operating points 1, 8, 9 were used in NMPC analysis...... 89

Figure 5.5. Response of crystal mean size and yield with respect to step change in (a)

antisolvent addition rate at stage 1 ( F1 ), (b) antisolvent addition rate at stage 2 ( F2 ), (c)

temperature at stage 1 (T1 ), and (d) temperature at stage 2 (T2 )...... 90

xv

Figure 5.6. NMPC performance of the antisolvent control method: (a) crystal size, yield and coefficient of variation (C.V.) of the CSD (red dotted lines are set-points); (b) antisolvent addition rates at stage 1 and 2; (c) concentration and supersaturation ratio; and

(d) CPU time...... 98

Figure 5.7. NMPC implementations: (a) controlled variables of nucleation control; (b) manipulated variables of nucleation control; (c) controlled variables of growth control; (d) manipulated variables of growth control (red dotted lines are set-points). C.V. is coefficient of variation of the CSD...... 100

Figure 5.8. NMPC implementations: (a) controlled variables of temperature control; (b) manipulated variables of temperature control; (c) controlled variables of global control; (d) manipulated variables of global control (red dotted lines are set-points). C.V. is coefficient of variation of the CSD...... 101

Figure 5.9. Attainable region of crystal size and yield when feed flow rate increases from

5ml/min to 6ml/min. Operating point 1 is the basic operating set-point...... 103

Figure 5.10. Performance of the global control NMPC for disturbance rejection: (a) scenario 1, disturbance profile in feed flow rate (F0), crystal size, yield, and coefficient of variation (C.V.) of CSD (red dotted lines are set-points); (b) scenario 1, manipulated variables: F1, F2, T1, T2 are antisolvent addition rates and temperatures at stages 1 and 2;

(c) scenario 2, disturbance profile in feed concentration (c0), crystal size, yield, and coefficient of variation (C.V.) of the CSD (red dotted lines are set-points); (d) scenario 2 manipulated variables...... 104

xvi

Figure 6.1. Experimental setup for (a) method 1 (without wet mill), (b) method 2 (with wet mill operated downstream) and (c) method 3 (with wet mill operated upstream). The parts shown by the solid dotted rectangles in each experimental setup are different among the three methods...... 110

Figure 6.2. Paracetamol powder seeds used in method 1...... 112

Figure 6.3. Process temperature, FBRM total counts and SWMCL in the MSMPRC in (a) experiment 1, (b) experiment 2, (c) experiment 3. Fluctuations in the temperature profile in (a) are caused by disturbances from the wet mill cooling water, but their influences on total counts and mean chord length are negligible...... 115

Figure 6.4. Process temperature, FBRM total counts and SWMCL in the MSMPRC in (a) experiment 4, (b) experiment 5, (c) experiment 6...... 116

Figure 6.5. (a) SWCLDs, (b) unweighted chord length distributions and (c) unweighted relative frequency distributions obtained from FBRM at steady-state (11st RT)...... 119

Figure 6.6. Microscope images for crystals obtained at steady-state (11st RT) in (a) experiment 1, (b) experiment 2, (c) experiment 3, (d) experiment 4, (e) experiment 5 and

(f) experiment 6...... 121

Figure 7.1. Experimental setup of continuous two-stage MSMPRC...... 126

Figure 7.2. ADNC set-point and heating/cooling algorithm...... 128

Figure 7.3. CryMOCO user interface for ADNC implementation...... 128

Figure 7.4. Process temperature, ADNC jacket set-point, FBRM total chord counts, ADNC total counts set-point and SWMCL in single-stage MSMPRC: (a) without ADNC

(experiment 1), (b) with ADNC (experiment 2)...... 131

xvii

Figure 7.5. Crystal chord length distribution (CLD) at steady-state 1 and steady-state 2 in experiment 2 when ADNC is used in single-stage MSMPRC...... 132

Figure 7.6. Crystal microscope images of (a) experiment 1, (b) experiment 2 steady-state

1, and (c) experiment 2 steady-state 2...... 133

Figure 7.7. Process temperature, ADNC jacket set-point, FBRM total chord counts, ADNC total counts set-point and SWMCL in the second stage: (a) without ADNC (experiment 3),

(b) with ADNC implemented in the second stage (experiment 4). The variation in total counts between 3.5 and 4 RT is due to probe fouling and cleaning...... 134

Figure 7.8. CLD at steady-state 1 and steady-state 2 in experiment 4 when ADNC is used in the second stage...... 135

Figure 7.9. Crystal microscope images of (a) experiment 3, (b) experiment 4 steady-state

1, and (c) experiment 4 steady-state 2...... 136

Figure 7.10. Process temperature, ADNC jacket set-point, FBRM total chord counts,

ADNC total counts set-point and SWMCL in (a) first stage with ADNC (experiment 5), (b) second stage with ADNC (experiment 5)...... 137

Figure 7.11. CLD at steady-state 1 and steady-state 2 in experiment 5 when ADNC is used in both stages...... 138

Figure 7.12. Crystal microscope images of (a) experiment 5 steady-state 1, (b) experiment

5 steady-state 2...... 138

Figure 7.13. Process temperature, ADNC jacket set-point, FBRM total chord counts,

ADNC total counts set-point and SWMCL in experiment 6 with disturbances introduced artificially...... 140

xviii

Figure 7.14. CLD at different times in experiment 6 with disturbances introduced artificially...... 140

Figure 8.1. Experimental setups for (a) process 1 and (b) process 2...... 147

Figure 8.2. WMADNC methodology...... 149

Figure 8.3.(a) Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL during startup and steady-state operation in process 1 with WMADNC approach. (b) Comparison of total particle chord counts dynamics in process 1 without control and with WMADNC approach.

...... 150

Figure 8.4. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 1 with

WMADNC approach under set-point change...... 152

Figure 8.5. SWCLDs of steady-state 1 and steady-state 2 in process 1 with WMADNC approach...... 153

Figure 8.6. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 1 under disturbance with WMADNC approach...... 154

Figure 8.7. SWCLDs at steady-state 2 and steady-state 3 in process 1 with WMADNC implemented upstream...... 155

Figure 8.8. Microscope images of crystals obtained in (a) steady-states 1, (b) steady-state

2 and (c) steady-state 3 in process 1 with WMADNC implemented upstream...... 156

Figure 8.9. Raman spectra of crystals obtained in steady-state 1, steady-state 2 and steady- state 3 in process 1 with WMADNC implemented upstream...... 156

xix

Figure 8.10. (a) Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL during startup and steady-state operation in process 2 with WMADNC approach. (b) Comparison of total particle chord counts dynamics in process 2 without control and with WMADNC approach.

...... 158

Figure 8.11. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 2 with

WMADNC approach under set-point change...... 159

Figure 8.12. SWCLDs of steady-state 4 and steady-state 5 in process 2 with WMADNC approach...... 159

Figure 8.13. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 2 with

WMADNC approach under disturbance...... 160

Figure 8.14. SWCLDs of steady-state 5 and steady-state 6 in process 2 with WMADNC implemented downstream...... 161

Figure 8.15. Microscope images of crystals obtained in (a) steady-states 4, (b) steady-state

5 and (c) steady-state 6 in process 2 with WMADNC implemented downstream...... 162

Figure 8.16. Raman spectra of crystals obtained in steady-state 4, steady-state 5 and steady- state 6 in process 2 with WMADNC implemented downstream...... 162

Figure 9.1. Schematic of experimental setup...... 167

Figure 9.2. Example chromatogram of (a) caffeine and (b) aspirin...... 168

Figure 9.3. Temperature, concentration, total counts and unweighted mean chord length during caffeine crystallization process...... 169

xx

Figure 9.4. Temperature and concentrations of aspirin (ASA) and degradation byproduct salicylic acid (SA), during crystallization of aspirin in water...... 170

Figure 9.5. (a) UV/vis first derivative spectrum, and (b) Raman spectrum of a 25 oC saturated aspirin-ethanol solution...... 172

Figure 9.6. UPLC-based UV/vis and Raman quick calibration for aspirin...... 174

Figure 9.7. Real-time aspirin (ASA) product purity monitoring with salicylic acid (SA) present as impurity...... 176

Figure 9.8. Real-time aspirin (ASA) product purity monitoring with paracetamol (PCM) present as impurity...... 177

xxi

LIST OF ABBREVIATIONS

AA Acetic Acid

ACN Acetonitrile

ADNC Automated Direct Nucleation Control

API Active Pharmaceutical Ingredient

ASA Acetylsalicylic Acid (Aspirin)

ATR Attenuated Total Reflectance

CCAC Combined Cooling/Antisolvent Crystallization

CLD Chord Length Distribution

CPP Critical Process Parameter

CQA Critical Quality Attribute

CryMOCO Crystallization Monitoring and Control System

CryPRINS Crystallization Process Informatics System

CSD Crystal Size Distribution

CSO Controlled State of Operation

CV Coefficient of Variation

CWMC Continuous Wet Milling-Crystallization

DNC Direct Nucleation Control

FBRM Focused Beam Reflectance Measurement

xxii

FTIR Fourier Transform Infrared spectroscopy

HPLC High-Performance Liquid Chromatography

LMPC Linear Model Predictive Control

LTI Linear Time Invariant

MeOH Methanol

MIMO Multi-Input Multi-Output

MSMPR Mixed Suspension Mixed Product Removal

MSMPRC Mixed Suspension Mixed Product Removal Crystallizer

MPC Model Predictive Control

NMPC Nonlinear Model Predictive Control

ODE Ordinary Differential Equation

PAT Process Analytical Technology

PBE Population Balance Equation

PBM Population Balance Model

PCM Paracetamol

PFC Plug Flow Crystallizer

PID Proportional-Integral-Derivative

PVM Particle Vision and Measurement

QbC Quality-by-Control

QbD Quality-by-Design

RGA Relative Gain Array

RT Residence Time

xxiii

SA Salicylic Acid

SD Standard Deviation

SQP Sequential Quadratic Programming

SSC Supersaturation Control

SWCLD Square Weighted Chord Length Distribution

SWMCL Square Weighted Mean Chord Length

TO Turnover

TOPRT Turnover Per Residence Time

UPLC Ultra-Performance Liquid Chromatography

WM Wet Milling

WMADNC Wet Milling-Based Automated Direct Nucleation Control

WMSG Wet Mill Seed Generation

YI Yield Index

xxiv

LIST OF SYMBOLS

abc,, fitted coefficients [varies] B nucleation rate in batch crystallizer [#/m3 s-1]

3 -1 Bi nucleation rate at stage i [#/cm min ]

3 -1 Bprim primary nucleation rate [#/cm min ]

3 -1 Bsec secondary nucleation rate [#/cm min ] c* solubility [g solute/g solvent] c solute concentration [g solute/g solvent] c0 solute concentration of feed solution [g solute/g solvent] solute concentration at stage i in MSMPR cascade ci crystallizers [g solute/g solvent mixture] c0 () API API concentration at time 0 [g solute/g solvent] c0 () imp impurity concentration at time 0 [g solute/g solvent] ct () API API concentration at time t [g solute/g solvent] ct () imp impurity concentration at time t [g solute/g solvent] c absolute supersaturation [g solute/g solvent]

DDBG, dissolution rate [varies]

Dr wet mill rotor outer diameter [m]

-1 Fw addition rate of water in batch crystallizer [g min ]

-1 F0 volumetric flow rate of feed solution [ml min ] volumetric antisolvent addition rate at stage i in MSMPR Fi cascade crystallizers [ml min-1]

xxv

F ()i volumetric outlet flow rate at stage i [ml min-1] out fi frequency of chord length included in i th interval [-] G growth rate in batch crystallizer [m s-1]

-1 Gi growth rate at stage i in MSMPR crystallizers [cm min ] J objective [varies]

Kc Cooling gain [-]

Kh Heating gain [-] kGGGBBB1, k 2,  k 3,  k 1,  k 2,  k 3  model parameters [varies]

3 -1 kprim primary nucleation rate constant [#/cm min ]

3 -1 ksec secondary nucleation rate constant [#/cm min ] kv volume shape factor [-] L particle characteristic size [μm] L average chord length [μm]

Li chord length of i th interval [μm]

Ln number-average mean size [μm] M control horizon [-]

2 m2 second moment of CSD [cm ] mliq mass of mother liquor/solvent mixture [g] n number density [#/m3] n redefined number density [#] number density at stage i in MSMPR cascade crystallizers ni [#/cm3] nprim primary nucleation order [-] nsec secondary nucleation order [-] P prediction horizon [-] peak height of a selected wavelength in UV/vis first derivative p or Raman spectrum of aspirin [varies]

xxvi

Q weighting matrix in penalty term [-]

3 -1 qmill recycle volumetric flow rate through wet mill [cm min ]

3 -1 qMSMPRC volumetric flow rate through MSMPRC [cm min ] R ideal gas constant [J mol-1 K-1] S supersaturation ratio [-] T temperature [K or oC]

o Ti temperature at stage i in MSMPR cascade crystallizers [ C]

TP length of prediction horizon [min]  residence time [min] residence time at stage i in MSMPR cascade crystallizers  i [min] t time [min]

t1 cooling start time in batch crystallizer [min]

t2 cooling end time in batch crystallizer [min]

t3 antisolvent start time in batch crystallizer [min]

t4 antisolvent end time in batch crystallizer [min]

tbatch batch time [min]

ttot total operation time [min] u Input variable [varies] V batch crystallizer volume [m3]

Vi volume of mother liquor/solvent mixture [ml] v 0 ethanol volume fraction of feed solution [-] v i ethanol volume fraction at stage i [-] -1 vts wet mill rotor tip speed [cm min ] w weight for penalty term [-] x state vector in state-space model [varies] Y yield [-]

xxvii y output variable [varies] yset set-point of output variable [varies]

 i dilution factor at stage i in MSMPR cascade crystallizers [-]  sampling time [min]

ij i , j th element in relative gain array [-]

-3 c crystal density [g cm ]

-3 ethanol ethanol density [g cm ]

-3 liq density of mother liquor/solvent mixture [g cm ]

-3 w water density [g cm ]  wet mill rotor angular speed [rpm]

i i th moment in batch crystallizer

i redefined i th moment in batch crystallizer

()i  j j th moment at stage i in MSMPR cascade crystallizers

xxviii

ABSTRACT

Yang, Yang. Ph.D., Purdue University, December 2016. Development of Novel Design and Control Approaches for Integrated Crystallization Operation and Systems. Major Professor: Zoltan K. Nagy.

Crystallization is an important separation and purification technique for a wide variety of solid products in pharmaceutical, food and fine chemical industries. The production of more than 90% of the active pharmaceutical ingredients (APIs) involves crystallization.

Therefore, it is crucial to design and control pharmaceutical crystallization processes so that the desired process requirement and product critical quality attributes (CQAs), including process yield, crystal purity, crystal size distribution (CSD), crystal shape and polymorphic form can be obtained. Studies about cooling or antisolvent only batch crystallization of pure APIs have been extensively reported in the literature in the last century. But design and control of more complex and integrated crystallization operation and systems require further development to meet the needs of improved CQAs, simplified and automated process control in pharmaceutical manufacturing.

This dissertation focuses on the development of novel design and control approaches for three different types of integrated crystallization operation and systems: (1) combined cooling/antisolvent crystallization (CCAC) systems; (2) integrated crystallization and wet milling operation; (3) crystallization systems in the presence of impurities. The design and

xxix control of both batch and, more recently, continuous crystallization systems were studied using population balance model (PBM), process analytical technology (PAT) and feedback control strategies.

First, CCAC systems in both batch and continuous mixed suspension mixed product removal (MSMPR) cascade crystallizer configurations were designed and controlled via model-based approaches. PBM-based optimization techniques were used. A linear operating profile based fast design strategy was developed to achieve desired crystal size and yield for batch CCAC systems, which was experimentally validated. The steady-state operation of continuous MSMPR cascade crystallizers was optimized and compared with batch operation. An efficient startup procedure was modeled and demonstrated able to provide large improvement on startup duration and waste minimization. In addition, two control frameworks, decentralized proportional-integral-derivative (PID) control and nonlinear model predictive control (NMPC), were investigated in two stage MSMPR cascade crystallizers, with both yield and crystal size targeted as controlled variables, providing a systematic strategy to compute attainable regions of particle size under different control configurations.

The second part of this dissertation is dedicated to design and control integrated crystallization and wet milling operation. A toothed rotor-stator lab scale wet mill unit was integrated with continuous mixed suspension mixed product removal crystallizer

(MSMPRC). It was used upstream as a continuous nucleator for in situ seed generation, and used downstream as a continuous size reduction unit through recycling. The experimental results indicate that the integrated crystallization and wet milling operation can lead to improved yield, CSD and process startup compared to crystallization operation

xxx without wet milling. A PAT-based feedback control approach called automated direct nucleation control (ADNC), was developed to achieve consistent and automated closed- loop control on CSD in continuous MSMPR cascade crystallizers. This ADNC approach was also extended to integrated continuous crystallization and wet milling operation, and a wet milling-based automated direct nucleation control (WMADNC) approach was developed to achieve improved control of CSD.

In the last part of this dissertation, an ultra-performance liquid chromatography (UPLC) system with automated process sampling and dilution was developed as an online PAT tool for the first time to monitor crystallization systems in the presence of impurities. Three online UPLC-based applications were proposed and implemented, including concentration monitoring in crystallization and degradation, quick calibration for UV/vis and Raman spectroscopy, and real-time crystallization product purity monitoring.

In this dissertation, a series of novel applications were proposed and validated via simulation and/or experimental studies that demonstrated the significance of quality-by- control (QbC) approaches for complex and integrated crystallization operation and systems.

xxxi

PUBLICATIONS

1. Yang, Y., Nagy, Z.K., 2014. Model-based systematic design and analysis approach for

unseeded combined cooling and antisolvent crystallization (CCAC) systems. Cryst.

Growth Des. 14, 687–698. Chapter 3.

2. Yang, Y., Nagy, Z.K., 2015. Combined cooling and antisolvent crystallization in

continuous mixed suspension, mixed product removal cascade crystallizers: steady-

state and startup optimization. Ind. Eng. Chem. Res. 54, 5673–5682. Chapter 4.

3. Yang, Y., Nagy, Z.K., 2015. Advanced control approaches for combined

cooling/antisolvent crystallization in continuous mixed suspension mixed product

removal cascade crystallizers. Chem. Eng. Sci. 127, 362–373. Chapter 5.

4. Yang, Y., Song, L., Gao, T., Nagy, Z.K., 2015. Integrated upstream and downstream

application of wet milling with continuous mixed suspension mixed product removal

crystallization. Cryst. Growth Des. 15, 5879–5885. Chapter 6.

5. Yang, Y., Song, L., Nagy, Z.K., 2015. Automated direct nucleation control in

continuous mixed suspension mixed product removal cooling crystallization. Cryst.

Growth Des. 15, 5839–5848. Chapter 7.

6. Yang, Y., Song, L., Zhang, Y., Nagy, Z.K., 2016. Application of wet milling-based

automated direct nucleation control in continuous cooling crystallization processes. Ind.

Eng. Chem. Res. 55, 4987–4996. Chapter 8.

xxxii

7. Yang, Y., Zhang, C., Pal, K., Koswara, A., Quon, J., McKeown, R., Goss, C., Nagy,

Z.K. Application of ultra-performance liquid chromatography as an online process

analytical technology tool in pharmaceutical crystallization. Cryst. Growth Des.

Submitted. Chapter 9.

8. Yang, Y., Pal, K., Koswara, A., Sun, Q., Zhang, Y., Quon, J., McKeown, R., Goss, C.,

Nagy, Z.K. Application of feedback control and in situ milling to improve particle size

and shape in the crystallization of a slow growing needle-like active pharmaceutical

ingredient. Int. J. Pharm. Submitted.

9. Yang, Y., Nagy, Z.K., 2015. Application of nonlinear model predictive control in

continuous crystallization systems. Proceedings of the American Control Conference,

4282–4287.

10. Acevedo, D., Peña, R., Yang, Y., Barton, A., Firth, P., Nagy, Z.K., 2016. Evaluation

of mixed suspension mixed product removal crystallization processes coupled with a

continuous system. Chem. Eng. Process. Process Intensif. 108, 212–219.

1

CHAPTER 1. INTRODUCTION

1.1 Background

Crystallization is a very efficient and economical separation and purification technique, in which a solute can transfer from liquid solution to a pure crystalline phase. It is one of the oldest unit operations used in chemical, food and pharmaceutical industries (Mullin,

2001). More than 90% of the active pharmaceutical ingredients (APIs) are crystals of small organic molecules and would require at least one crystallization step in their manufacture

(Alvarez et al., 2010; Wong, et al., 2012). In pharmaceutical industry, crystallization is widely used to purify the APIs produced from organic synthesis or fermentation and then followed by filtration, , milling, granulation, tableting, etc. Due to the safety concern of impurities and strict regulatory requirements, purity is always one of the most important critical quality attributes (CQAs) in pharmaceutical crystallization process. In general, very high purity (e.g. > 99.9%) is required in the crystallization of APIs. Besides, the crystal polymorphic form is another important CQA, since different polymorphs of a same API can have significantly different solubility, dissolution behavior and bioavailability in human body. In addition, crystallization is not only a separation technique, but also an important particle technology. The crystal size distribution (CSD) and crystal shape significantly influence the efficiency of downstream operations such as filtration and drying (Braatz, 2002; Nagy et al., 2013; Yu et al., 2007).

2

Typically, large and uniform crystals that have small aspect ratio are desired considering manufacturability. But in some situations small and uniform crystals are also of great importance due to their improved bioavailability compared to large crystals. Considering the requirements of various CQAs as well as the process yield, the design and control of crystallization process is important and challenging in pharmaceutical industry.

There are various ways to carry out crystallization, such as through cooling, antisolvent addition, or evaporation. Cooling and antisolvent addition are the two common ways used in pharmaceutical crystallization because of the purification purpose. The design and control of cooling only or antisolvent only crystallization processes to achieve desired

CQAs have been extensively reported in literature since last century (Aamir et al., 2010;

Hermanto et al., 2010; Ma and Wang, 2012; Nagy et al., 2013; Nowee et al., 2008).

However, in pharmaceutical industry, in many situations a combined cooling/antisolvent crystallization (CCAC) method is actually used in order to maximize the process yield.

Due to the extra freedom in process parameters, the integrated CCAC system, which has not yet been widely studied in literature, is more challenging to design and control but also more promising to further improve the desired CQAs in both batch and continuous processes (Nagy et al., 2008b).

In addition, crystallization is typically investigated in literature on its own as a single unit operation. As a result, the ability to control CQAs, such as size and shape, is very limited. The integration of other unit operations into crystallization process is of great interests in recent years. For example, wet milling is a widely used unit operation to tailor the particle size distribution (PSD) of slurry. It is typically used in downstream to improve the CSD obtained from a batch crystallization process. However, in recent years it was

3 found that by integrating wet milling with batch crystallization, one could not only reduce the number of unit operations and intensify the process, but also achieve better control on crystal quality (Engstrom et al., 2013; Kougoulos et al., 2011). Therefore, integrated crystallization and wet milling operation, especially for integrated continuous crystallization, is a very novel and promising direction in industrial crystallization.

What is more, impurity is a crucial aspect in crystallization process development.

Generally speaking, crystallization of multi-component system (e.g. API with impurities) is much more challenging to study than single component system (e.g. pure API). There are several reasons. First of all, pure API of a certain polymorph in a particular solvent always has a fixed solubility curve which is not difficult to measure or search for in literature, whereas the API solubility could be different when impurity is present (Simone et al., 2015c). In addition, the nucleation and growth kinetics of the components in solution could be influenced by each other (Abbou Oucherif et al., 2013). Lastly, the compositions of solid and liquid phases of a multicomponent system are difficult to characterize in real- time. For instance, spectroscopy techniques are useful to measure concentration of single component system however often not useful when impurity is present due to the overlapped spectra. Because of the above reasons, although large amount of crystallization papers are available in literature, most of them use pure compounds in order to simplify the study.

However, in pharmaceutical industry, most crystallization processes are with impurities.

There is significant demand of developing novel approaches to characterize crystallization systems in the presence of impurities.

4

1.2 Research Aims

The aims of this dissertation are to develop novel design and control approaches for integrated crystallization operation and systems, including combined cooling/antisolvent crystallization (CCAC) systems, integrated crystallization and wet milling operation, and crystallization systems in the presence of impurities:

1) Modeling and simulating integrated CCAC systems in both batch and continuous

MSMPR processes in MATLAB using population balance model (PBM) to study the evolution of CSD under different operation conditions, including cooling rate, antisolvent addition rate, batch time or residence time. Then using optimization techniques (e.g. sequential quadratic programming (SQP)) to optimize the operation parameters to achieve certain objectives, such as maximization of average crystal size, or minimization of startup duration in continuous MSMPR process. Finally, based on the optimization results, summarizing the general rules behind and developing systematic design and control approaches which can be used to simplify the design of CCAC experiments in batch and continuous MSMPR processes.

2) Studying different ways to integrate wet milling operation with continuous

MSMPR crystallization, including using it upstream as a continuous nucleator or downstream as a size reduction unit. Investigating the influences on process yield, startup duration, average crystal size and width of size distribution, and finding out the advantages and disadvantages of both the upstream and downstream applications of wet milling. Then, to develop a PAT-based feedback control approach for a simple continuous MSMPR crystallization process to achieve closed-loop controlled CSD. Finally, applying and adapting this PAT-based feedback control approach to the more complex integrated

5 continuous crystallization and wet milling processes, therefore combining the advantages of both wet milling and closed-loop control for continuous crystallization.

3) Developing a novel online PAT tool to characterize the crystallization systems in the presence of impurities, which are difficult to monitor and control using the existing online or inline PAT. More specifically, studying how ultra-performance liquid chromatography (UPLC) can be applied as an online PAT tool in pharmaceutical crystallization process, what are the benefits compared to the existing PAT, and how to monitor crystal purity, which is the most important CQA, in real-time based on online

UPLC.

1.3 Research Contributions

The main contributions of this dissertation include:

1) Developed a model-based systematic design approach for unseeded CCAC batch process using PBM. The approach can be easily implemented in practice using automated lab reactors, and gives the optimal times for the start of a linear cooling profile and/or constant antisolvent addition rate. It can be used to produce operating surfaces based on which the performance of CCAC systems can be understood and operating strategies rapidly designed depending on the initial concentration, solvent composition and desired yield (Yang and Nagy, 2014).

2) Found out the relationship and general rule between optimal batch CCAC system and optimal continuous MSMPR CCAC system. This is important in transferring existing

CCAC knowledge from batch to continuous MSMPR process. In addition, it was found that CCAC in continuous MSMPR crystallization could achieve significant reduction of

6 the startup duration time and waste by using dynamic startup operating profiles (i.e. temperature, antisolvent addition rate) instead of the typically used fixed operating parameters.

3) Demonstrated the feasibility and benefits of applying advanced control approach, more specifically, nonlinear model predictive control (NMPC) in continuous MSMPR

CCAC system for fast target process and product property (i.e. yield and crystal size) change-over and disturbance rejection. On the other hand, the simple decentralized proportional-integral-derivative (PID) control was proved infeasible.

4) Integrated rotor-stator wet mill with continuous MSMPR crystallization through upstream and downstream applications of wet milling. The benefits of both applications on process yield, crystal size, width of size distribution and startup duration were identified compared to crystallization without wet milling.

5) Developed a PAT-based feedback control approach, automated direct nucleation control (ADNC), for continuous multistage MSMPR cooling crystallization. The proposed approach was observed to provide quick startup, high quality control of CSD, as well as automated and effective disturbance suppression.

6) Developed a novel wet milling based automated direct nucleation control

(WMADNC) approach for continuous MSMPR cooling crystallization. It could be used to achieve closed-loop controlled primary nucleation or secondary nucleation kinetics, depending on the target particle size and whether WMADNC was used upstream or downstream. It not only combined the advantages of wet milling and ADNC, but also showed good control performance on particle size due to decoupled nucleation and growth kinetics.

7

7) Applied UPLC as an online PAT tool in pharmaceutical crystallization for the first time. Three online UPLC-based applications were developed, including crystallization and degradation monitoring, quick calibration for UV/vis and Raman spectroscopy, and real- time crystallization product purity monitoring.

1.4 Thesis Structure

Chapter 2 presents a literature review about pharmaceutical crystallization process development, including fundamentals of crystallization process, PAT, model-based and model-free control, batch and continuous crystallization, CCAC, integrated crystallization and wet milling operation, and characterization of crystallization systems in the presence of impurities.

Chapter 3 introduces a PBM-based systematic design approach for unseeded CCAC batch process to maximize average crystal size under yield and operating constraints.

Experimental validation of the proposed fast design approach is also presented.

Chapter 4 describes the steady-state and startup optimization of CCAC in multistage continuous MSMPR crystallization process via PBM to maximize the average crystal size and minimize the startup duration and waste amount, by optimizing cooling rates, antisolvent addition rates and residence times.

In Chapter 5, the feasibility of applying different control approaches, decentralized PID control or NMPC, in CCAC in two stage continuous MSMPR process is analyzed, with yield and crystal size aimed as controlled variables.

8

Chapter 6 shows the upstream and downstream applications of wet milling in continuous MSMPR cooling crystallization. The improvements on CSD, yield and startup duration is summarized.

Chapter 7 describes the development of a PAT-based feedback control, ADNC, in one and two stage continuous MSMPR cooling crystallization process. Its benefits on set-point change and disturbance rejection in continuous crystallization are introduced.

In Chapter 8, the proposed ADNC approach is integrated with wet milling technique, and a WMADNC approach is developed for continuous cooling crystallization. The control performance of WMADNC is evaluated.

Chapter 9 presents the application of UPLC as a novel online PAT tool in pharmaceutical crystallization. It can be used to monitor concentration for multiple components, to provide quick calibration for UV/vis and Raman spectroscopy, and to measure crystallization product purity in real-time.

Chapter 10 summarizes the main conclusions of this dissertation, and recommendations for future works are introduced as well.

9

CHAPTER 2. LITERATURE REVIEW

2.1 Fundamentals of Pharmaceutical Crystallization

Crystallization is a process in which randomly arranged molecules, atoms or ions assemble and form an ordered solid structure called crystal (Mullin, 2001). Since molecules that have same or similar structures are more likely to come together, crystallization is a good separation and purification technique. It is also an important particle technology to control various particulate properties of crystals, such as particle size distribution (PSD)

(or crystal size distribution (CSD)) and crystal shape (Mullin, 2001).

Pharmaceutical crystallization is typically carried out as suspension crystallization, which starts from a liquid solution that is a homogeneous mixture of the continuous phase called solvent and the dispersed phase called solute. Typically, the solutes in pharmaceutical crystallization include certain desired organic API as well as some impurities. Commonly used solvent include water, lower alcohols, ketones and esters, and so on (Mullin, 2001). One of the most important thermodynamic properties of a solution is solubility, which represents the equilibrated concentration of the solute in that specific solvent at a certain temperature. Since solubility is affected by temperature and solvent composition, solubility curves or solubility surfaces are built to report solubility as a function of temperature or/and solvent composition, as shown in Figure 2.1.

10

Figure 2.1. Crystallization phase diagram.

Solution is undersaturated and dissolution of existing crystals happens if the solute concentration is below solubility curve. The region above solubility curve is supersaturated.

At high supersaturation above nucleation metastable limit, both formation of new crystals, which is called nucleation, and growth of existing crystals can occur. However, at low supersaturation nucleation is difficult but crystal growth still happens. The region between nucleation metastable limit and solubility curve is called metastable zone, in which crystallization is typically operated. Both thermodynamic and kinetics factors, such as chemical potential of nucleation and cooling rate being used can influence the position of nucleation metastable limit and the width of metastable zone (Davey and Garside, 2000).

Supersaturation is the thermodynamic driving force for a crystallization process. There are three common ways to express supersaturation: concentration driving force, c , supersaturation ratio, S , and relative supersaturation  (Mullin, 2001):

11

c  c  c* (2.1)

c S  (2.2) c*

c   S 1 (2.3) c*

* where c is concentration of solute, c is equilibrated concentration or solubility. There are various ways to generate supersaturation in suspension crystallization, including cooling, antisolvent addition and solvent evaporation. Cooling or/and antisolvent addition are the most common approaches used in pharmaceutical crystallization.

Nucleation and growth are the two most important kinetics in crystallization. Both of them are strong functions of supersaturation. There are two major types of nucleation: primary nucleation and secondary nucleation (Mullin, 2001). Nucleation that occurs in a solution which is free of the solute crystals is defined as primary nucleation. On contrary, secondary nucleation is defined as nucleation that happens with existing solute crystals. In general, primary nucleation requires higher supersaturation than secondary nucleation. In engineering applications, empirical equations are often used to express primary and

secondary nucleation kinetics, Bprim and Bsec (Yang et al., 2016):

nprim Bprim k prim c (2.4)

nsec Bsec k sec c m2 (2.5)

where kprim and ksec are primary and secondary nucleation rate constants, nprim and nsec

are primary and secondary nucleation orders, c is concentration driving force, and m2

12 is the second moment of the CSD (Frawley et al., 2012). For size independent crystal growth, the following empirical equation is often used to express growth kinetics:

g G kg c (2.6)

where G is growth rate, kg is growth rate constant and g is the growth order. kprim , ksec ,

sometimes are used as constant numbers, but sometimes are expressed as Arrhenius equations which depend on temperature and nucleation or growth activation energy

(Frawley et al., 2012).

2.2 Process Analytical Technology (PAT)1

In 2004, the U.S. Food and Drug Administration (FDA) published the guidance for pharmaceutical industry about process analytical technology (PAT), in which PAT was considered as “a system for designing, analyzing, and controlling manufacturing through timely measurements (i.e., during processing) of critical quality and performance attributes of raw and in-process materials and processes, with the goal of ensuring final product quality” (FDA, 2004). In terms of pharmaceutical crystallization processes, the CQAs of the final product typically include chemical purity, polymorphic form, CSD and crystal shape (King et al., 2015; McGlone et al., 2015; Park et al., 2016; Zhang et al., 2015). The

CQAs are affected by important process parameters such as the concentration and supersaturation profiles of the API and impurities. Many PAT tools have been developed in recent decades to monitor and control CQAs and important process parameters inline,

1 Reproduced with permission from Cryst. Growth Des., submitted for publication. Unpublished work copyright 2016 American Chemical Society.

13 online, atline or offline in pharmaceutical crystallization (Simon et al., 2015). In inline analyses, the sample interface is located inside the process therefore separate sampling system is not needed (Hassell and Bowman, 1998). On the other hand, online analyses require automated sampling and transport through a sample line to an automated analyzer, whereas atline or offline analyses need manual sampling with transport to analyzer (Hassell and Bowman, 1998). Focused beam reflectance measurement (FBRM) was a laser technique extensively used in pharmaceutical crystallization as an inline PAT tool to provide in situ measurement of crystal chord length distribution (CLD) that reflects CSD

(Li et al., 2013, 2014; Peña and Nagy, 2015; Yang and Nagy, 2014; Yang et al., 2015d,

2015e, 2016). Inline video microscopy system such as particle vision and measurement

(PVM) are techniques that can record crystal appearance (i.e. shape and size) in real-time

(Acevedo and Nagy, 2014; Acevedo et al., 2015; Borsos et al., 2016). Raman spectroscopy with immersion optics has been developed to measure crystal polymorphic form inline

(Simone et al., 2014a). There are also some important atline or offline spectroscopy techniques such as mid-infrared (IR), near infrared (NIR), Raman and x-ray diffraction

(XRD), which can quantify polymorphs and crystallinity (Steen et al., 2015; Barba et al.,

2013). Inline or online PAT are favorable because they can avoid manual sampling errors and allow real-time closed-loop process control (Barrett et al., 2010; Chew et al., 2007;

Khan et al., 2011; Ma and Wang, 2012). Simone et al. developed a composite sensor array

(CSA) based on multiple inline PAT including Raman, NIR, and attenuated total reflectance ultraviolet/visible (ATR-UV/vis) spectroscopy, FBRM and PVM for an automated intelligent decision support and control system (Simone et al., 2015b). The proposed CSA can exploit signals from all in situ PAT measurements to implement

14 automated feedback control. Table 2.1 presents some references about application of inline

PAT for various purposes, such as monitoring of crystallization product, characterization of polymorphic transformation, determination of nucleation and growth kinetics.

Table 2.1 References about applications of PAT in crystallization

PAT Property measured Reference FBRM Particle chord counts, CLD Barrett and Glennon, 2002; Kempkes et al., 2008; Li et al., 2013, 2014; Peña and Nagy, 2015; Saleemi et al., 2012a, 2012b; Simone et al., 2014b, 2015b, 2015c; Su et al., 2010; Zhao et al., 2013. PVM Size and shape Acevedo et al., 2015; Barrett and Glennon, 2002; Barrett et al., 2010; Borsos et al, 2016; Kempkes et al., 2008; Simon et al., 2009b; Simone et al., 2015b; Su et al., 2010; Zhao et al., 2013. UV/vis Concentration Aamir et al., 2010; Bakar et al., 2009; Saleemi et al., 2012b; Thompson et al., 2005; Simone et al., 2014b, 2014c, 2015a, 2015b, 2015c. Raman Polymorph, concentration Simone et al., 2014a, 2014b, 2014c, 2015a, 2015b; Su et al., 2010; Zhao et al., 2013.

2.3 Model-Based and Model-Free Control2

In general, crystallization process can be controlled via either model-based or model- free approaches (Nagy et al., 2008b, 2013). The applications of both approaches in batch crystallization have been studied in depth in literature. In model-based approaches, typically a PBM is used to describe the evolution of the CSD in the crystallization process.

By applying optimization techniques, open-loop optimal temperature or/and antisolvent

2 Reproduced with permission from Yang, Y., Nagy, Z.K., 2015. Chem. Eng. Sci. 127, 362–373. Copyright 2015 Elsevier. http://www.sciencedirect.com/science/article/pii/S0009250915000895

15 addition profiles could be obtained to produce desired CSD without violating equipment limitations (e.g. maximum/minimum temperature, cooling rate or antisolvent addition rate) or productivity/yield constraint. Plenty of papers have shown the benefits of using model- based optimization and offline calculated optimal operating profiles in unseeded/seeded batch/continuous cooling and/or antisolvent crystallization (Mullin and Nyvlt, 1971;

Rawlings et al., 1993; Vetter et al., 2014; Worlitschek and Mazzotti, 2004; Xie et al., 2001;

Yang and Nagy, 2015c; Zhang and Rohani, 2003).

In addition, more advanced model-based approaches that solve the open-loop optimization repeatedly, such as model predictive control (MPC) has been applied in batch crystallization process (Hermanto et al., 2009; Nagy and Braatz, 2003; Kalbasenka et al.,

2007). MPC uses the mathematical model and real-time measurements to optimize the current operating curve, based on the predicted future behavior of the system. For systems of high complexity and nonlinearity, nonlinear model predictive control (NMPC) is used instead of linear model predictive control (LMPC). Hermanto et al. utilized NMPC strategy to control polymorphic transformation in a batch crystallization process (Hermanto et al.,

2009). It was proved to be more robust than other existing control strategies like temperature control or concentration control. A robust extended Kalman filter-based

NMPC framework was developed by Nagy and Braatz for batch crystallization processes

(Nagy and Braatz, 2003). The proposed NMPC approach considerably enhanced robust performance as compared with open-loop optimal control. Although these model-based approaches require significant time and engineering effort to develop the model, and have difficulties to model practical objectives (e.g. purity, filterability), they can enhance

16 process understanding, provide theoretically optimal recipe, and require much smaller number of experiments than statistical experimental design (Nagy et al., 2008b).

Model-free (or direct design) approaches such as supersaturation control (SSC) (Nagy et al., 2008a; Nagy and Aamir, 2012; Gron et al., 2003), direct nucleation control (DNC) or automated direct nucleation control (ADNC) (Abu Bakar et al., 2009; Saleemi et al.,

2012a, 2012b) are methodologies that maintain the operating curve within the metastable zone to avoid nucleation or to generate controlled dissolution. They are based on the direct use of PAT tools (e.g. FBRM, UV/Vis, IR spectroscopy) in closed-loop control of crystallization processes. Gron et al. developed a closed-loop supersaturation control approach of batch crystallizer using in-process ATR-FTIR spectroscopy (Gron et al., 2003).

This strategy could produce crystals of desired product properties, notably uniformity in crystal size and shape. Nagy et al. investigated a model-free concentration control approach to maintain the supersaturation within the metastable zone in batch cooling crystallization

(Nagy et al., 2008a). It showed lower sensitivity to disturbances than model-based temperature control approach. Abu Bakar et al. implemented a DNC approach to directly control the amount of nuclei present using measurements taken by FBRM in closed-loop control framework through heating/cooling and addition of solvent/antisolvent (Abu Bakar et al., 2009). It was found DNC approach resulting in larger crystal size than uncontrolled crystallization experiments. Saleemi et al. developed an ADNC approach to automatically detect secondary nucleation and accidental seeding in the system and remove fines in situ

(Saleemi et al., 2012a, 2012b). These model-free approaches only require basic crystallization concepts and simple mathematical tools without any modeling effort. They can provide fast approaches for highly automated recipe design. The disadvantage is that

17 the set-point operating profiles are suboptimal and typically chosen arbitrarily or by trial- and-error (Fujiwara et al., 2005). The model-free feedback control approaches are aligned with the novel quality-by-control (QbC) concept which allows larger applicable design space and more robust process operation than the commonly used quality-by-design (QbD) approaches (Yang et al., 2015e).

2.4 Batch and Continuous Crystallization3

Crystallization is implemented as a batch operation for most of the time in pharmaceutical industry. Batch crystallization has large operation flexibility and only requires a short development time (Ward et al., 2006). However, in batch crystallization the operating conditions change with time, resulting in control difficulty and batch to batch inconsistency (Alvarez et al., 2011). The process and equipment efficiency is also limited in batch operation (Alvarez et al., 2011). These intrinsic disadvantages of batch crystallization have strongly motivated the development of continuous crystallization systems during the recent decade. Continuous crystallization is operated at steady-state, in which crystals of consistent purity, yield, CSD and polymorphic form can be produced.

Therefore, in a long run continuous crystallization can have better quality consistency, process and equipment efficiency, and productivity as compared with batch crystallization

(Alvarez et al., 2011; Quon et al., 2012; Zhang et al., 2012).

Several types of continuous crystallizers have been developed, including mixed suspension mixed product removal (MSMPR) crystallizer, plug flow crystallizer (PFC) and

3 Reproduced with permission from Yang, Y., Nagy, Z.K., 2015. Chem. Eng. Sci. 127, 362–373. Copyright 2015 Elsevier. http://www.sciencedirect.com/science/article/pii/S0009250915000895

18 oscillatory baffled crystallizer (OBC) to name only a few. Compared to the other two,

MSMPR crystallization processes are more frequently applied in industrial crystallization since they provide a smaller technology change from batch and are simpler to operate

(Vetter et al., 2014). MSMPR crystallizer is an ideally well-mixed vessel that has feed solution continuously entering in and product slurry continuously withdrawn. The output slurry is assumed to have exactly the same composition as the vessel content. It can be used in various configurations, including single stage, multistage, or with recycle loops

(Acevedo et al., 2016; Alvarez et al., 2011; Quon et al., 2012; Vetter et al., 2014; Wong et al., 2012; Zhang et al., 2012). It is operated at steady-state, which could be close to batch equilibrium condition by increasing the number of stages. As a result, it is relatively easy to convert batch knowledge into continuous MSMPR crystallization systems (Wong et al.,

2012). In addition, it has large tolerance for high suspension density, and can be used for crystallization of slow growing crystals which require long residence times (Quon et al.,

2012). Cooling, or antisolvent addition, or both could be applied to generate supersaturation at a certain MSMPR crystallization stage. Vetter et al. modeled MSMPR cascade crystallization process using PBM (Vetter et al., 2014). Cooling only crystallization, antisolvent addition only crystallization, and CCAC were used as three case studies, in which the particle size attainable regions of MSMPR cascade crystallizers of different number of stages were obtained and compared with batch crystallizer and PFC.

Quon et al. implemented and optimized a two stage MSMPR cascade crystallization process for continuous reactive cooling crystallization of aliskiren hemifumarate (Quon et al., 2012). Crystals of both high yield and purity were obtained. Alvarez et al. implemented a three stage continuous MSMPR cascade crystallizer to produce cyclosporine crystals

19 using cooling to generate supersaturation (Alvarez et al., 2011). A PBM was developed in the same work, and used to optimize crystal size, yield and purity, as well as to obtain crystallization kinetic parameters. Wong et al. constructed two continuous single stage

MSMPR crystallizers with recycle loop for cooling crystallization and CCAC (Wong et al., 2012). They experimentally demonstrated the feasibility and potential advantages (e.g. high yield) of having recycle loop in single stage MSMPR crystallizer. A continuous

CCAC process using two stage MSMPR cascade crystallizer was implemented by Zhang et al., and was experimentally validated to be able to well control CSD, yield and purity

(Zhang et al., 2012). PFC is another common approach to perform continuous crystallization. In PFC, supersaturation is generated by cooling or antisolvent addition along the tube to promote nucleation and growth (Ferguson et al., 2012). One of the major challenges in PFC is to keep crystals suspended (Vetter et al., 2014). Typically, a high flow rate is necessary, and a long tube length is required to achieve the desired residence time.

As a result, PFC is usually used for fast growing compounds with low residence times. It usually produces small crystals with narrow size distribution (Quon et al., 2012). Alvarez and Myerson applied static mixing elements into the antisolvent crystallization of ketoconazole in a PFC (Alvarez and Myerson, 2010). Eder et al. developed a PFC with continuous seeding for cooling crystallization of aspirin in ethanol (Eder et al., 2010). In addition, Lawton et al. presented a continuous OBC, which behaves similar as ideal PFC but without the limitation of high flow rates (Lawton et al., 2009).

20

2.5 Combined Cooling/Antisolvent Crystallization (CCAC) System4

The fundamental driving force for crystallization is supersaturation, which typically is generated by cooling or antisolvent addition in pharmaceutical industry. Significant amount of work exist that studies the design and control problem of cooling only or antisolvent addition only crystallization (Aamir et al., 2010; Hermanto et al., 2010; Ma and

Wang, 2012; Nagy and Braatz, 2012; Nagy et al., 2013; Nowee et al., 2008). Besides using these two approaches independently, CCAC is another technique commonly used in the pharmaceutical industries. Compared with cooling or antisolvent addition only crystallization, the major advantage of CCAC is the higher yield, due to the extra driving force of this combination (Sheikhzadeh et al., 2008). The final product properties (e.g.

CSD) can also be improved if the process can be controlled well (Nagy et al., 2008b). In addition, since solvent/antisolvent ratio can also have effect on crystal shape and polymorphic form, CCAC can provide extra control on multiple crystal properties simultaneously (Yang and Nagy, 2014).

Nagy et al. have firstly presented a comparative overview about the control problem of

CCAC through both model-free and model-based framework (Nagy et al., 2008b). In this work the optimal simultaneous cooling and antisolvent addition operating profiles in terms of various single objectives were obtained through model-based optimization. Both the simulation and experimental validation results indicated that the CCAC approach can yield better crystal quality than the cooling or antisolvent only processes. Trifkovic et al. showed an experimental study of offline and online optimal control for a CCAC process (Trifkovic

4 Reproduced with permission from Yang, Y., Nagy, Z.K., 2014. Cryst. Growth Des. 14, 687–698. Copyright 2013 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/cg401562t

21 et al., 2009). A neuro-fuzzy logic controller was also developed by the same group to increase the theoretical yield and control the supersaturation through combination of cooling and antisolvent addition (Sheikhzadeh et al., 2008). Recently an operating strategy with combined cooling and antisolvent addition was utilized to produce consistent CSD

(Hermanto et al., 2012). Lindenberg et al. determined nucleation and growth kinetics of aspirin (acetylsalicylic acid) using in situ monitoring techniques (Lindenberg et al., 2009).

In the same work, a strategy with constant cooling and antisolvent addition rate under the same feeding time for seeded CCAC process was discussed, indicating that it should be favored as compared to processes where the antisolvent was added all at once at the beginning or end of the cooling ramp.

2.6 Integrated Crystallization and Wet Milling Operation5

Traditionally in crystallization design large and uniform crystals are desired in order to facilitate downstream operations, such as filtration and drying. But small and uniform crystals are also of great interest due to their high dissolution rate and bioavailability, therefore may be used directly in formulation so that downstream processes like milling and granulation can be avoided. One approach to generate small and uniform particles is to use impinging jet mixer (Hacherl et al., 2003; Johson and Prud’homme, 2003; Kamahara et al., 2007). The residence time inside the impinging jet mixer can be manipulated and either crystalline or amorphous particles can be obtained. However typically the jet mixer

5 Reproduced with permission from Yang, Y., Song, L., Gao, T., Nagy, Z.K., 2016. Cryst. Growth Des. 15, 5879–5885. Copyright 2016 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/acs.cgd.5b01290

22 requires two streams that have similar volumes therefore it is mainly applied in antisolvent or reactive crystallization (Kamahara et al., 2007). Wet mill is another commonly used approach that can reduce particle size and size distribution, as well as normalize API physical properties in pharmaceutical industry (Burcham et al., 2008; Harter et al., 2013).

Harter et al. investigated the scale up approach of terminal particle size and milling time required to achieve target particle size in wet mill batch operation (Harter et al., 2013).

Kougoulos et al. compared the PSD obtained in high shear wet mill with other particle engineering techniques such as sonocrystallization and dry impact milling after a conventional batch cooling crystallization (Kougoulos et al., 2011). Engdtrom et al. integrated toothed rotor-stator wet mills with batch cooling crystallizers in a recycle loop across lab and pilot plant scale (Engstrom et al., 2013). It was found that the normalized energy imparted by the wet mills could be correlated to the particle size reduction by using an energy factor and a probability factor. Kamahara et al. coupled a high-speed rotor-stator wet mill with a semibatch antisolvent crystallization process using a recycle loop to achieve controlled secondary nucleation under high shear environment during crystallization and successfully scaled up this approach (Kamahara et al., 2007). As compared to the traditionally used dry milling technique which follows crystallization, these integrated wet milling and crystallization techniques have the advantages of achieving process intensification (i.e. eliminated separate unit operation of milling), better API chemical and physical stability, as well as safety (e.g. reduced exposure, prevented API dust explosions)

(Burcham et al., 2008). In addition, according to the literature, both primary and secondary nucleation rates increase dramatically under high shear environment (Allen et al., 2008;

Gray et al., 1995; Kamahara et al., 2007; Reguera et al., 2003). Therefore, instead of using

23 rotor-stator wet mill downstream, potentially it may also be used upstream as a continuous in situ seed generation tool for crystallization processes. This continuous wet seeding technique can have advantages on seed quality consistency as well as safety compared with commonly used dry seeding technique. But the upstream application of continuous wet milling in pharmaceutical crystallization has not been studied in literature yet.

2.7 Characterization of Crystallization System in the Presence of Impurities6

The development of PAT that can provide fast and accurate inline or online concentration measurement is of great importance in the characterization, design and control of pharmaceutical crystallization processes. Inline spectroscopy techniques have been extensively reported to monitor concentration in crystallization processes. Cornel et al. demonstrated the quantitative application of in situ attenuated total reflectance Fourier transform infrared (ATR-FTIR) and Raman spectroscopy with multivariate analysis and principal component analysis (PCA) to monitor liquid composition during cooling crystallization, precipitation and polymorphic transformation of L-glutamic acid (Cornel et al., 2008). Zhou et al. used ATR-FTIR spectroscopy and a partial least square (PLS) model for in situ concentration measurement and closed-loop control of supersaturation in antisolvent crystallization of paracetamol (Zhou et al., 2006). Thompson et al. measured solute concentration in the crystallization of an important organic compound using ATR-

UV spectroscopy, a calibration curve, a PLS temperature correction model and orthogonal signal correction (Thompson et al., 2005). Some key advantages of spectroscopic

6 Reproduced with permission from Cryst. Growth Des., submitted for publication. Unpublished work copyright 2016 American Chemical Society.

24 concentration analysis (e.g. ATR-FTIR, ATR-UV, Raman) are that it is typically rapid

(<30 seconds per measurement) and there are many commercially available probe type tools for in situ measurement. There are several disadvantages of spectroscopy approaches, however, that strongly limit their applications in pharmaceutical manufacturing. First, spectroscopic techniques are challenging to use in processes that contain multiple components, especially for structurally related chemicals due to their overlapped spectra.

Many industrial pharmaceutical crystallization processes contain more than one component. For this and other reasons, knowledge of chemometrics is important to process the spectroscopy data in order to obtain accurate concentration values. Commonly used statistical and mathematical tools include data preprocessing techniques (e.g. spectral normalization, linear baseline correction, standard normal variate, multiplicative scatter correction) and multivariate data analysis (e.g. PCA, PLS) (Cornel et al., 2008). This is also because spectroscopic signals, for example Raman, can be influenced by not only liquid composition but also solid density, particle size, temperature, etc. Sometimes standard univariate approaches based on peak height or area, which are much simpler than chemometrics techniques, are also feasible. Saleemi et al. were able to obtain good concentration measurement using ATR-UV/vis spectroscopy and a simple non-linear calibration model based on single wavelength data (Saleemi et al., 2012a, 2012b). Due to the large number of factors that can influence spectroscopic signals and the limitation of concentration measurement, a long time (e.g. a week) is typically required in the standard calibration approach in order to collect data at different concentrations, temperatures, etc. to train the calibration model (Saleemi et al., 2012b). In addition, the spectroscopy calibration model can be system and instrument dependent, which could make the model

25 none-transferable or in need of recalibration due to instrument drifting. One would even need to re-calibrate every few weeks or months due to instrument drifting. Finally, probe based PAT tools such as ATR-UV/vis, ATR-FTIR, Raman, FBRM, and PVM, are not accurate in the presence of fouling, which is a very common phenomenon in crystallization processes.

High-performance liquid chromatograph (HPLC) is widely used to measure analyte concentration in pharmaceutical analysis because it can provide very accurate results for multiple components (i.e. API and impurities) using simple linear calibration models based on the peak area of each component. HPLC often requires up to half an hour per measurement which significantly limits its use for process monitoring. Revolutionary advances in chromatography columns and instrumentation, such as reduced particle size, higher pressure capabilities and minimal system dispersion led to the introduction of ultra- performance liquid chromatography (UPLC) in 2004. Compared to HPLC, UPLC offers improved resolution and sensitivity, lower solvent consumption, and analysis times of only a few minutes (Swartz, 2005). Together, these improvements make UPLC very attractive for PAT applications. In addition, there are a number of different ways to calibrate HPLC or UPLC, such as single external standard, calibration curve or internal standard, that are all simpler than the spectroscopy calibration described earlier because chromatographic peak area is typically a linear function of analyte concentration. Therefore, online UPLC would be a good replacement for spectroscopy when the characterization of crystallization systems in the presence of impurities is required.

HPLC and UPLC have been widely used in the analysis of pharmaceutical products.

Purity is one of the most important CQAs in pharmaceutical crystallization due to strict

26 regulatory requirements. Purity, however, is mainly measured atline in literature by manually taking samples, filtering, washing, diluting and running HPLC atline (Alvarez et al., 2011; Li et al., 2016; Simone et al., 2015c). This commonly used atline purity measurement method is inconvenient, incompatible with real-time process control, and sometimes inaccurate due to the manual sampling errors. As a result, the development of an online crystallization product purity measurement approach is extremely important for crystallization systems in the presence of impurities.

2.8 Summary on Literature Review

This chapter presents a literature review for recent development in pharmaceutical crystallization, including fundamentals of crystallization process, PAT, model-based and model-free control, batch and continuous crystallization, CCAC, integrated crystallization and wet milling operation, and characterization of crystallization systems in the presence of impurities. In general, simple crystallization operation and systems have been extensively studied in literature, whereas there is a lack of investigation about more complex and integrated crystallization operation and systems, including CCAC systems, integrated crystallization and wet milling operation, and crystallization systems in the presence of impurities.

27

CHAPTER 3. MODEL-BASED SYSTEMATIC DESIGN APPROACH FOR

UNSEEDED COMBINED COOLING/ANTISOLVENT CRYSTALLIZATION7

3.1 Introduction

The fundamental driving force for crystallization is supersaturation, which can be generated by cooling or antisolvent addition. Besides using these two supersaturation generation approaches independently, CCAC is another technique, often applied in the pharmaceutical industries. Despite the relatively widespread industrial use of CCAC, the modeling, design and control of these systems have not been studied in depth in literature yet, motivating the growing interest in this area in recent years.

The goal of this work is to develop a systematic approach for unseeded batch CCAC processes using a model-based simulation and optimization framework. The approach is based on the PBM of the CCAC solved using the method of moments (Nagy et al., 2008b).

Under fixed batch time, this model is then used to optimize the cooling time and antisolvent feeding time, which both have constant rates, to obtain suboptimal but easily implementable operating trajectories.

7 Reproduced with permission from Yang, Y., Nagy, Z.K., 2014. Cryst. Growth Des. 14, 687–698. Copyright 2013 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/cg401562t

28

Multiple scenarios with different initial solution conditions (initial temperature, solvent/antisolvent ratio) were analyzed and a CCAC design surface was generated that can be used to provide a map that indicating whether the process was controlled by antisolvent addition, cooling or both, and could also be used to estimate operating procedures quickly for different initial conditions that can provide a desired product property (e.g. mean size) under fixed batch time and yield requirements. The simulation results were also validated through experiments, performed in an automated laboratory crystallizer. Finally, the suboptimal trajectories obtained by our methodology were compared to the optimal trajectories obtained by full optimization of nonlinear trajectory.

The methodology that created the CCAC design surface for simplified operating trajectories could significantly reduce the number of experiments required for the design of unseeded CCAC process. The model system used in the evaluation studies was aspirin

(acetylsalicylic acid) in ethanol-water mixture.

3.2 Mathematical Modeling

The population balance equation (PBE) (Ramakrishna, 2000; Randolph, 1988) for a batch or semi-batch crystallizer in which crystal growth is size independent, crystal shape is uniform, crystal breakage or aggregation is negligible can be written as:

(nLtVt , ()) ( nLtGtVt , ()()) B()() t V t (3.1) tL where n is the volumetric number density(#/m3), L is the characteristic crystal size (m),

G is the growth rate (m s-1), and B is the nucleation rate (#/m3s-1).

29

For a semi-batch crystallizer where antisolvent addition is applied and total volume is time-varying, a redefined number density n(,) L t can be used to simplify equation (3.1)

(Nagy et al., 2008b):

n L,,() t  n L t V t (3.2) where V is the total volume of mother liquor (m3).

The corresponding redefined PBE becomes:

n L, t ( n L , t G ( t )) B()() t V t (3.3) tL

The above PBE was solved using the method of moments. The j th moment of n(,) L t is defined as (Nagy et al., 2008b):

  t Lj n L, t dL ,  j  0,1,  ,  (3.4) j      0

where 0 is the total number of crystals in the crystallizer, 10/ is the number-average mean size. Assuming the nuclei formed through primary nucleation in unseeded batch crystallizer has zero size, the ordinary differential equations (ODEs) of each moment can be written as (Nagy et al., 2008b):

dt   0  B t V t (3.5) dt

dt   j j  t G t,  j  1,2,  ,  (3.6) dt j1

The mass balance equations used in the model for the semi-batch crystallizer is derived as (Nagy et al., 2008b):

dm liq  F (3.7) dt w

30

dc c 2  Fw 3 c k v G (3.8) dt mliq m liq

dcethanol c ethanol  Fw (3.9) dt mliq

1 liq  (3.10) cethanol/ ethanol (1c ethanol ) / w

m V  liq (3.11) liq

where mliq (g) is the mass of mother liquor mixture, Fw (g water/min) is the flow rate of

antisolvent water, cs is the solute concentration (g solute/g solvent), c is the crystal

3 density (g/cm ), kv is the crystal volume shape factor, cethanol is the ethanol concentration

3 (g ethanol/g solvent), liq is the solvent mixture density (g/cm ). In this study, the values

3 of c and kv are fixed at 1.4 g/cm ,  /6, respectively.

Different empirical or semi-empirical equations are used in the literature to express nucleation and growth kinetics, which usually depend on the supersaturation and the temperature. In this chapter, the growth rate and primary nucleation rate of aspirin in ethanol-water mixture are expressed by empirical equations (Lindenberg et al., 2009):

kG2 * kG3 G kG1 exp  ( c S  1 ) (3.12) RT

k k Bkexp( B2 )exp(  B3 ) (3.13) B1 RTln 2 S

S c/ c* (3.14)

31

* where S is the supersaturation ratio, c is the solute concentration (g solute/g solvent), c

is the solubility (g solute/g solvent) at a certain temperature and ethanol concentration, kG1

, kG 2 , kG3 , kB1 , kB2 , kB3 are empirical parameters from literature (Lindenberg et al.,

2009), as given in Table 3.1.

Table 3.1. Growth and nucleation kinetic parameters of aspirin in ethanol-water mixture measured by Lindenberg et al., 2009.

Parameters Values

41 kG1 3.21 0.18  10 ms 

41 Growth kG2 2.58 0.14  10 J  mol

kG3 1.00 0.01

21 3 1 kB1 1.15 0.51 10 ms

41 Nucleation kB2 7.67 0.11  10 J  mol

kB3 0.16 0.01

There have been a variety of different approaches to calculate the solubility data at a certain temperature and ethanol concentration using scattered experimental data. The mathematical approach in this work is biharmonic interpolation (Sandwell, 1987). Table

3.2 presents some experimental solubility data determined by ATR-FTIR from literature

(Lindenberg et al., 2009). A surface described by biharmonic interpolation via the

MATLAB function griddata is shown as well (Figure 3.1).

32

Table 3.2. Solubility of Aspirin c* (g solute/kg solvent) from Lindenberg et al., 2009.

* cwater [g water/g solvent] cethanol [g ethanol/g solvent] T [℃] c [g solute/kg solvent] 0 1.00 25 237.9 0 1.00 35 378.4 0 1.00 50 648.4 0.10 0.90 25 283.2 0.10 0.90 35 419.1 0.10 0.90 50 689.2 0.25 0.75 25 266.6 0.25 0.75 35 395.3 0.25 0.75 50 679.1 0.40 0.60 25 186.2 0.40 0.60 35 304.8 0.40 0.60 50 570.8 0.55 0.45 25 87.9 0.55 0.45 35 149.5 0.55 0.45 50 377.0 0.75 0.25 25 16.2 0.75 0.25 35 32.9 0.75 0.25 50 101.6

0.8

0.6

0.4

[g/g solvent]

n

i r

p 0.2

s A C 0 100 C 50 e tha 50 40 no [%g/g solvent] l 30 C] 0 20 Temperature [

Figure 3.1. Solubility surface of aspirin in ethanol-water mixture. The circles are solubility data obtained from literature (Lindenberg et al., 2009).

33

3.3 Experimental Section

Acetylsalicylic acid (>99%, Sigma-Aldrich, USA), ethanol (pure, 200 pf, KOPTEC,

USP Grd) and deionized water were used in all tests. The experiments were carried out in

Easymax synthesis workstation (Mettler Toledo, USA), including an Easymax 102 reactor block, a 100 mL teflon-covered round-bottom glass reactor, an overhead glass stirrer, an external solid state thermostat and a SP-50 dosing unit. Water was used as antisolvent and was fed into the reactor by the SP-50 dosing unit. The iControl Easymax 4.2 software

(Mettler Toledo, USA) was used to design and control all experimental procedures, including the control of the reactor temperature, cooling rate, cooling time, dosing rate, dosing time, and stirring rate. Additionally, the reactor was equipped with a FBRM G400 probe (Mettler Toledo, USA) for in situ measurements of the CLD, the total counts and the square weight mean chord length (SWMCL) of the crystals in the slurry, and was connected with iC FBRM 4.3 software (Mettler Toledo, USA) for data analysis. The FBRM data was measured every 10 seconds and averaged at 20 seconds interval. All experiments were carried out at a stirring rate of 400 rpm. The setup is shown in Figure 3.2.

Figure 3.2. Schematic of experimental setup.

34

All crystallization experiments were unseeded. The initial solubility data is calculated via 2D interpolation from the solubility surface constructed from experimental data as described before, according to the initial temperature and ethanol concentration, which are

(48.5℃, 55% ethanol). The initial total mass of solvent and supersaturation ratio are fixed at 25g and 1.03, respectively, from which the initial mass of aspirin, ethanol and water are calculated. The solution was first heated to 5 degrees above the desired initial temperature and stabilized for 15 minutes. Then it was slowly cooled to the desired initial temperature for the controlled temperature profile and stabilized for 5 more minutes before the crystallization experiment began. The FBRM data was collected during the experiment.

Then after the 30 minutes batch time ended, the experiment was stopped immediately and slurry samples were taken out through a pipet and quickly filtered. The crystal samples were dried in hood and kept in sealed tube. A Nikon SMZ1500 microscope was used to take images of the crystal samples.

3.4 Results and Discussion

3.4.1 CCAC Design Through the Optimization of Operating Procedures With Constant

Cooling and Antisolvent Addition Rates

In order to make the cooling and antisolvent addition profiles easy to implement in industrial process, the cooling rate and antisolvent addition rate are kept constant in this study. The trajectories computed this way will be suboptimal compared to allowing time- varying cooling and antisolvent addition rates, but are easily implementable in industrial crystallization systems. The control problem of computing these operating trajectories can be formulated as an optimization problem, which was solved by SQP. The initial and final

35 temperatures, ethanol concentrations, as well as the batch time are fixed at certain desired values. The optimization variables are the times when the cooling profile starts and ends (

t1 and t2 , respectively) as well as the times when the antisolvent feeding starts and stops (

t3 , and t4 , respectively), as shown in Figure 3.3. The optimization problem can be written as follows:

max{Jn. m . s .  L n ( t batch )   1 ( t batch )/( 0 t batch )} (3.15) t1,,, t 2 t 3 t 4

subject to: model equations (3.5) to (3.14)

0t1 ,  t 2 ,  t 3 ,  t 4  tbatch  (3.16)

tt12 (3.17)

tt34 (3.18)

T(0) Tinitial ,  T ( t batch )  T final ,  mliq 0 mm liq,, initial ,  m liq tbatch  liq final (3.19)

where t1,,, t 2  t 3  t 4 are cooling start and end times, antisolvent start and end times,

respectively, tbatch is the batch time, which is fixed at 30 minutes. Tinitial , Tfinal , mliq, initial , m are the predetermined temperatures and masses of mother liquor for the initial and liq, final the final solutions. The inequality constraints from (3.16) to (3.18) limit the operation time of cooling and antisolvent addition in the proper range. The last equality constraints are used to fix both the initial and the final conditions of the solution at desired points. The objective function J is some desired crystal properties at the end of batch. In this study the number-average mean size was considered: J L()()/() t t t . n. m . s . n batch 1 batch 0 batch

36

55 40 Temperature 50 t1 Mass of solvent t 4 35

45

C]  40 30 35

Temperature [ 30 25 Mass of [g]solvent t 25 3 t2 20 20 0 5 10 15 20 25 30 Time [min]

Figure 3.3. Schematic of CCAC process with constant cooling rate and antisolvent addition rate. t,,, t  t  t are cooling start time and end time, antisolvent start time and end time, 1 2 3 4 respectively.

Multiple scenarios with different initial conditions but same initial solubility and concentration were evaluated to obtain the suboptimal trajectories (Figure 3.4). Especially, two scenarios were analyzed in detail, (i) Scenario 1 (S1), when the system followed the trajectory from point A (40.0 ℃, 75% ethanol) to point C (25 ℃, 40% ethanol) in the phase diagram, and (ii) Scenario 2 (S2), in which the process followed on operating curve from a different initial point B (48.5 ℃, 55% ethanol) to the same final point C (25 ℃, 40% ethanol) in the phase diagram (as shown in Figure 3.4). The initial points A and B were selected so that point A locates in an area of the solubility surface that has high sensitivity for temperature change but low sensitivity for change in antisolvent concentration, whereas point B locates in an area of the phase diagram where the solubility surface shows high sensitivity to change in antisolvent concentration but low sensitivity for temperature change. In addition, points A and B have the same solubility, indicating that both

37 trajectories, S1 and S2, provide the same concentration change. Thus, S1 and S2 are two typical scenarios in a CCAC process and are comparable with each other.

Figure 3.4. Suboptimal trajectories for S1, S2, and scenarios with other initial conditions.

Table 3.3 and Figure 3.5 present the suboptimal cooling times and antisolvent feeding times and corresponding operating profiles for S1 and S2. Figure 3.6 shows how the supersaturation ratio,  ()t and concentration change with time. 0

Table 3.3. Optimal cooling time, antisolvent feeding time and final crystal properties for

S1 and S2 obtained through optimization.

S1 S2

t1[min] 0.0 17.2

t2[min] 30.0 30.0

t3[min] 12.9 0.0

t4[min] 30.0 14.7

Lmn [] 529 527 CV..[] 0.284 0.238

38

C]

C]  40  50 35 40 30 (a) (b) 30

25 Temperature [ 0 5 10 15 20 25 30 Temperature [ 0 5 10 15 20 25 30 Time [min] Time [min] 1.5 50 1 35 1 40 (a) 0.5 30 0.5 Flow rate 30 (b) Flow rate Mass of solvent Mass of solvent

Flow rate [g/min]Flow 0 20

Flow rate [g/min]Flow 0 25 Mass of [g]solvent 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Mass of [g]solvent Time [min] Time [min] Figure 3.5. Suboptimal operating curves for (a) S1 and (b) S2.

4 x 10 1.8 10 (a) S1 (b) S1 S2 0.6 S2 8 1.6 0.5

6 0.4

1.4 [-] 0

 0.3 4 0.2 1.2 Supersaturation ratio Supersaturation [-] 2

0.1 Concentration [g Concentration solute/g solvent]

1 0 0 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Time [min] Time [min]

Figure 3.6. (a) Suboptimal supersaturation ratio,  ()t for S1 (solid line) and S2 (dotted 0 line); (b) suboptimal concentration trajectories for S1 (solid line) and S2 (dotted line).

For both S1 and S2, the supersaturation ratios remain high in the first 15 minutes to promote nucleation in the system, and then decrease to a low but relatively constant level to maintain growth (see Figure 3.6(a)). In addition, it can be found that the primary nucleation was induced by cooling only in S1, whereas was induced by antisolvent only in

S2 (see Figure 3.5 and Figure 3.6(a)). Trajectory S1 gives slightly larger particles than S2

(see Table 3.3) since in this trajectory nucleation occurs sooner (longer time available for

39 growth) and the number of particles is lower (see Figure 3.6(a)) although the initial and final concentrations in both cases are the same and the concentration trajectories are significantly different (see Figure 3.6(b)). Since nucleation occurs later for S2 a larger number of particles are required to achieve similar final concentration in shorter time as for S1.

Several other trajectories, which all have the same initial solubility and concentrations as S1 and S2, were also computed using a similar optimization framework but starting from different initial conditions in the phase diagram as shown in Figure 3.4. From these trajectories, it is easy to recognize that, in cooling sensitive regions, such as point A, cooling first strategy, or cooling induced nucleation is preferred. However, in antisolvent sensitive region, like point B, antisolvent first strategy or antisolvent induced nucleation is preferred. This phenomenon can be explained as an early primary nucleation, which needs relatively high supersaturation, is quite desirable in an unseeded batch crystallization to ensure that there is longer time for growth. The concepts of temperature-controlled operating region and antisolvent-controlled operating region were first introduced by Nagy et al., 2008b, but the present work provides a systematic analysis and application of these into CCAC design and control practice.

In the case of the trajectories obtained for scenarios S1 and S2 it can be seen that when cooling first is preferred, as expected cooling must start at 0 minute to ensure an early nucleation, and antisolvent should stop at the end of batch to maintain a maximum growth.

In other words, for cooling first strategy t1 and t4 can be fixed at 0 and 30 minute

respectively. Similarly, when antisolvent first is preferred, t2 and t3 can be fixed at 30 and

40

0 minute respectively. From the various additional operating trajectories obtained for different initial conditions presented in Figure 3.4, the same general conclusions can be achieved. Consequently, the operating trajectories can be simplified for the two strategies to have only two optimization variables each, as shown in Figure 3.7. In these simplified

cases the effect of the two design parameters ( t2 and t3 for cooling first, t1 and t4 for

antisolvent first) on the desired final product quality (expressed as the mean size Ln ) can be represented as an operating map using model-based simulation.

55 40 55 40 (a) Temperature (b) Temperature 50 Mass of solvent 50 t1 Mass of solvent 35 t4 35

45 45

C]

C]

  40 40 30 30

35 35 Temperature [ 30 Temperature [

Mass of [g]solvent 30 25 25 Mass of [g]solvent t 25 3 25 t2 20 20 20 20 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Time [min] Time [min]

Figure 3.7. Schematic of (a) cooling first and (b) antisolvent first strategy. t1 , t2 , t3 and t4 are cooling start time and end time, antisolvent start time and end time, respectively.

3.4.2 CCAC Design Using a Simulation-Based Operating Map

Because the suboptimal trajectories belong to either cooling first or antisolvent first strategies, the optimization problem discussed in the previous section may be converted into a simple simulation based evaluation problem. The key advantage of this approach is that a generic operating map can be developed that can be used for better understanding of

41 the crystallization operating conditions that will affect a particular crystal property and can be used for the rapid design of an operating strategy that will produce for example a desired mean crystal size and yield, without the requirement of performing model based optimizations.

As shown in Figure 3.7, the operating trajectories of cooling first strategy can be found

by simulating the CCAC processes when both t2 and t3 vary from 0 to 30 minute. Similarly, the operating trajectories for the antisolvent first strategy can be obtained by simulations

by varying both t1 and t4 from 0 to 30 minute. For the generation of the operation maps, t1 ,

t2 , t3 and t4 were simulated every 1 minute. The simulation results are presented as operating maps in form of 'mean size surface', in which the z value stands for simulated

number-average mean size Ln , and the x, y values stand for t2 and t3 for cooling first

strategy, or t1 and t4 for antisolvent first strategy, respectively. The Scenarios S1 and S2, corresponding to the initial conditions discussed previously were also evaluated here. The results are shown in Figure 3.8 and Table 3.4. The simultaneous approach (point 0) according to which both cooling and antisolvent additions occur throughout the entire batch duration from 0 to 30 minute is used as base case for comparison. Points 1, 2, 3 and 4 are the optimum points for each strategy in the two initial conditions investigated (1–S1, cooling first; 2– S1, antisolvent first; 3– S2, cooling first; and 4– S2, antisolvent first). It can be found that for initial condition of S1, 1is larger than 2, indicating that cooling first strategy is better than antisolvent first strategy. However, for initial condition of S2, 3 is smaller than 4, indicating that cooling first strategy is worse than antisolvent first strategy.

This difference is due to the different sensitivities for the change of temperature and

42 antisolvent concentration of the solubility surface in the initial points of S1 and S2 (Nagy et al., 2008b). In both scenarios the optimum points of the cooling first or antisolvent first strategies are better than the results obtained using the simultaneous cooling and antisolvent addition case. These results illustrate that significant improvement in product quality can be achieved by using the proposed systematic CCAC design approach.

(a) S1, cooling first (b) S1, antisolvent first (2) 600 (1) 350 (0)

m] 300

m] 500

  400 250

300 (0) 200

Mean size Mean [ size Mean size Mean [ size 200 150 30 30 30 Antisolvent20 start time 30 Antisolvent20 end time 20 20 (t 10 (t 10 3 ) [min] 10 4 ) [min] 10 0 0 Cooling end time 0 0 Cooling start time (t ) [min] ) [min] 2 (t 1

(c) S2, cooling first (d) S2, antisolvent first

500 (3) 600

(4) m]

m] (0)   400 400 (0)

300 200

Mean size Mean [ size Mean [ size 200 0 30 30 20 30 20 30 Antisolvent start time 20 Antisolvent end time 20 (t 10 (t 10 3 ) [min] 10 4 ) [min] 10 0 0 0 0 Cooling end time Cooling start time (t ) [min] ) [min] 2 (t 1

Figure 3.8. The mean size surfaces of cooling first and antisolvent first strategies for scenarios S1 and S2, respectively. Points (1)(2)(3)(4) are the corresponding optimum points. Point (0) is simultaneously cooling and adding antisolvent through 0 to 30 minutes, which is used as base case for comparison.

43

Table 3.4. Comparison of the simultaneous case, optimal case of cooling first, and optimal case of antisolvent first through simulation.

S1 S2 Simultaneous Optimal Optimal of Simultaneous Optimal Optimal of (0) of antisolvent (0) of antisolvent cooling first (2) cooling first (4) first (1) first (3)

t1[min] 0 0 22 0 0 17

t2[min] 30 30 30 30 30 30

t3[min] 0 13 0 0 11 0

t4[min] 30 30 25 30 30 14

Lmn [] 319 529 334 384 452 524 CV..[] 0.376 0.283 0.319 0.300 0.272 0.219

Both the full four parameter optimization and the simplified two parameter simulation based results are summarized in Table 3.5 for comparison.

Table 3.5. Comparison of the suboptimal trajectories obtained by optimization and simulation.

S1 S2 Optimization Simulation Optimization Simulation

t1[min] 0.0 0 17.2 17

t2[min] 30.0 30 30.0 30

t3[min] 12.9 13 0.0 0

t4[min] 30.0 30 14.7 14

Lmn [] 529 529 527 524

CV..[] 0.284 0.283 0.238 0.219

44

It can be seen that the differences between optimization and simulation results are negligible (and are within the discretization used in the simulation based approach). This indicates that the proposed simulation-based operating map can be used as a quick CCAC design approach.

3.4.3 Experimental Validation of the CCAC Design Approach Using the Simulated

Operating Map

For the CCAC process of both cooling first and antisolvent first strategies corresponding to the initial condition S2, five cases were evaluated experimentally and were compared with the simulation results. These cases were selected based on the mean size maps (Figure

3.8(c)(d)) developed using the simulations, as shown in Figure 3.9. The five cases correspond to the simultaneous cooling and antisolvent addition (case A), the optimal operating trajectories (corresponding to the largest mean sizes) for the cooling first and antisolvent first scenarios (cases B and D, respectively), as well as two additional non- optimal points in the operating map corresponding to some smaller target mean sizes for both the cooling first and antisolvent first scenarios (cases C and E, respectively), which should produce the smallest crystals in both scenarios. The corresponding operating

procedures were implemented experimentally. Both the nucleation time tnuc (induction time) and the SWMCL obtained from the FBRM (as a qualitative estimate of the mean size used as the quality index) were used as criteria for comparison. The results are shown in

Table 3.6. The experimental SWMCL was computed as the average of the measurements during the last minute of the batch (with 95% confidence interval).

45

(a) Cooling first Mean size [microns] (sim) 450

164+16- (exp) B 400 151+6- (exp) A 125+6- (exp) 30 30 C 350 Antisolvent start time 20 20 (t 3 ) [min] 300 10 10 ) [min] Cooling(t 2 end time 0 0

(b) Antisolvent first Mean size [microns] (sim) 500 137+8- (exp) E 170+16- (exp) 400 151+6- (exp) D A 30 30 300 Antisolvent end time 20 20 (t 200 4 ) [min] 10 10 ) [min] Cooling(t 1 start time 0 0

Figure 3.9. Comparison between experimental mean size (SWMCL) and simulated mean

size ( Ln ), when applying (a) cooling first strategy or (b) antisolvent first strategy corresponding to the initial condition of S2. Point A: simultaneous cooling and antisolvent; point B: optimal of cooling first; point C: non-optimal of cooling first; point D: optimal of antisolvent first; point E: non-optimum of antisolvent first. The color bar and 'sim' stand for simulated results. The number on the map and 'exp' stand for experimental results. The variation for experimental result is 2σ (95% confidence interval).

According to Figure 3.10, the experimental times for the onset of nucleation (induction time) agree very well with the simulated induction times, indicating that the primary

46 nucleation kinetics from the literature (Lindenberg et al., 2009) is suitable for the experimental setup and conditions used in this work.

Table 3.6. Comparison of simulated mean size and experimental SWMCL for the corresponding cases for conditions S2.

Simultaneous Optimal Non-optimal Optimal Non-optimal cooling and cooling cooling first antisolvent antisolvent antisolvent first first first (case A) (case B) (case C) (case D) (case E)

t1[min] 0 0 0 17 17

t2[min] 30 30 15 30 30

t3[min] 0 11 11 0 0

t4[min] 30 30 30 14 30

* 384 452 358 524 272 Ln ()[] sim  m 

SWMCL(exp)[ m ] 151±6 164±16 125±6 170±16 137±8 * 'sim' stands for simulated, 'exp' stands for experimental.

In the case of cooling first strategy the induction times are relatively close to each other, nevertheless follow the same relative order both in simulation and experimentally. In the case of the antisolvent first strategy the induction time differences are larger, which is also in good agreement with the simulation results. The SWMCL at the end of the batches for both the cooling first and the antisolvent first strategies indicate that larger crystals are obtained in the cases of the optimal strategies (B and D). The selected cases for smaller desired mean particle size (C and E) result in significantly smaller SWMCL as expected.

The simultaneous cooling and antisolvent case produces particles with intermediate size.

47

The experimental results are again in excellent agreement with the simulations in terms of the relative values. It might be noticed that in Figure 3.9, Figure 3.10 and Table 3.6, the absolute values of experimental SWMCL are smaller than the simulated mean sizes. This is a general trend in the case of FBRM data because the short chord lengths of crystals usually have higher probabilities to be measured by FBRM than longer chord lengths.

250 (a) Cooling first 250 (b) Antisolvent first 200

200

]

]

m

m 

[ 150 

[ 150

100 100 SWMCL

SWMCL SWMCL Simultaneous (case A) 50 Simultaneous (case A) 50 Optimal Optimal (case B) (case D) Non-optimal Non-optimal (case C) (case E) 0 0 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Time [min] Time [min]

600 600 (c) Cooling first (d) Antisolvent first 500 500 Simultaneous (case A) Optimal (case D) 400 400 m]

m] Non-optimal (case E)

 

300 300

200

200 Mean [ size Mean size Mean [ size Simultaneous (case A) 100 100 Optimal (case B) Non-optimal (case C) 0 0 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Time [min] Time [min]

Figure 3.10. Comparison between (a) experimental mean size (SWMCL) of cooling first

strategy and (b) antisolvent first strategy, (c) simulated mean size ( Ln ) of cooling first strategy and (d) antisolvent first strategy, respectively. Case A: simultaneous cooling and antisolvent; case B: optimal of cooling first; case C: non-optimal of cooling first; case D: optimal of antisolvent first; case E: non-optimal of antisolvent first.

48

Figure 3.11 and Figure 3.12 present the normalized cumulative chord length distributions and crystal images at the end of batch for each case, showing that the optimal cases do produce more large crystals, while the non-optimal cases selected with the goal of smaller sized crystal production create more fines. The experimental mean sizes indicated on Figure 3.9, are in very good agreement with the trend suggested by the mean size surface map. These result demonstrate that simulated mean size surface can be used as an efficient tool for the design of CCAC systems, reducing significantly the number of experiments required for the design of these processes to produce e.g. a desired target mean particle size.

Figure 3.11. Normalized cumulative chord length distributions at the end of batch for each case of (a) cooling first strategy and (b) antisolvent first strategy. Case A: simultaneous cooling and antisolvent; case B: optimal of cooling first; case C: non-optimal of cooling first; case D: optimal of antisolvent first; case E: non-optimal of antisolvent first.

49

Figure 3.12. Final crystals of case A (simultaneous cooling/ antisolvent), case B (optimal of cooling first), case C (non-optimal of cooling first), case D (optimal of antisolvent first), and case E (non-optimal of antisolvent first). All images have the same magnification.

3.4.4 Full optimization of nonlinear trajectory based design of CCAC systems

Based on the nonlinear model presented in the paper, the full optimization of the CCAC process was also evaluated. The initial condition, final condition, batch time and objective function are the same as section for the optimization problem of the simplified scenarios presented earlier (equations (3.15)–(3.19)). The computation of the optimal temperature trajectory and the antisolvent addition rate profile is formulated as a nonlinear optimization problem solved using a standard constrained nonlinear optimization approach based on

SQP implemented in MATLAB. The batch time is discretized into N ( N  8) intervals, and the temperatures and antisolvent addition rates at every discrete time kN1,2,..., are

50 considered as the optimization variables. The optimization problem can be written as follows:

max{Jn. m . s . L n ( t batch ) 1 ( t batch )/( 0 t batch )} (3.20) T( k ), Fw ( k )

subject to:

model equations (3.5)–(3.14)

Tmin T(k) T max (3.21)

dT RR (3.22) T,mindt T,max

Fw,min F w (k) F w,max (3.23)

T(0) Tinitial , T ( N )  T final , mliq 0  m liq,, initial , m liq N  m liq final (3.24)

where Tmin , Tmax , RT,min , RT,max are the minimum and maximum temperatures and

temperature ramp rates, respectively. Fw,min , Fw,max are the minimum and maximum antisolvent addition flow rates. T , T , m , m are the predetermined initial final liq, initial liq, final temperatures and masses of mother liquor for the initial and the final solutions. The first three constraints can ensure that the temperatures and antisolvent addition rates are in a proper operating range. The last equality constraints are used to fix both the initial and the final conditions of the solution at desired points.

Multiple scenarios were evaluated here using the full optimization to obtain the optimal trajectories, which are shown in Figure 3.13. Similar to the previous section, scenarios with initial conditions corresponding to points A and B (scenarios S1opt and S2opt) were evaluated and analyzed in detail.

51

Solubility [g/g solvent]

0.6

B 0.5 A 50 0.4 S2opt S1opt 45 0.3 100 C] 40  C 80 0.2 e 35 tha no [% g/g solvent]60 l 30 0.1 C Temperature [ 40 25

Figure 3.13. Optimal trajectories for S1opt, S2opt, and scenarios with other initial conditions.

The optimal operating profiles, supersaturation ratio,  ()t and concentration 0 trajectories for S1opt and S2opt are shown in Figure 3.14 and Figure 3.15. For both S1opt and

S2opt, it can be observed that the temperature decreases fast before primary nucleation occurs, and then remains almost constant during nucleating. The fast cooling at the beginning generates supersaturation and makes the solution reach fast the metastable limit for primary nucleation. The constant temperature stage afterwards ensures no more supersaturation is created during nucleation. This cooling pattern guarantees an early and well controlled nucleation, which allows the generation of an optimum amount of nuclei early in the process which then have longer time until the end of the batch to grow to larger sizes. These well controlled, early nucleation events can be observed from Figure 3.15(a).

The results indicated a quick increase in supersaturation until the onset of nucleation after which the supersaturation declines and is maintained at a lower constant level to maintain the number of particles avoiding further nucleation. Additionally, according to Figure

52

3.15(b), although the two scenarios have very different initial temperature and solvent/antisolvent ratio, they prefer almost the same concentration trajectory, which further demonstrates that the trajectories and the operating curves provided in Figure 3.13

and Figure 3.14 are indeed optimal.

C]

C]  40  50 35 40 30 (a) 30 (b)

25 Temperature [ 0 5 10 15 20 25 30 Temperature [ 0 5 10 15 20 25 30 Time [min] Time [min]

4 50 4 35 45 3 Flow rate 3 Flow rate 40 2 Mass of solvent 2 Mass of solvent 30 35

1 (a) 30 1 (b) Flow rate [g/min]Flow

0 25 rate [g/min]Flow 0 25 Mass of [g]solvent 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Mass of [g]solvent Time [min] Time [min]

Figure 3.14. Optimal operating curves for (a) S1opt and (b) S2opt.

4 x 10 1.8 4 S1 S1 (a) opt (b) opt S2 0.6 S2 opt opt 1.6 3 0.5

0.4

1.4 2 [-] 0

 0.3

0.2

1.2 1 Supersaturation ratio Supersaturation [-]

0.1 Concentration [g Concentration solute/g solvent]

1 0 0 0 5 10 15 20 25 30 0 5 10 15 20 25 30 Time [min] Time [min]

Figure 3.15. (a) Optimal supersaturation ratio and  ()t for S1 (solid line) and S2 0 opt opt

(dotted line); (b) optimal concentration trajectories for S1opt (solid line) and S2opt (dotted line).

53

These results indicate that in principle the optimal operating trajectory for any crystallization processes is given by the optimal supersaturation profile, which is the actual driving force for both nucleation and growth. The optimization of the cooling and/or antisolvent addition will in principle provide only the means to achieve the optimal supersaturation trajectory within the existing operating constraints. Therefore, multiple combined cooling and antisolvent addition trajectories may lead to identical product quality since they may create similar supersaturation trajectory.

An interesting observation from the full optimization results is that although the initial conditions given by points A and B have very different sensitivities for change of temperature and antisolvent concentration, the trajectories for both S1opt and S2opt have similar patterns, both characterized by cooling only induced nucleation, as well as combined cooling and antisolvent addition based growth. This phenomenon can be explained by the pronounced temperature effect on the nucleation and growth kinetics compared to the effect of supersaturation. The solubility surface based sensitivity analysis to determine whether a particular initial condition leads to antisolvent first or cooling first procedure only considers the effect of the supersaturation as the main driving force for achieving early primary nucleation. However, in the full optimization problem, the solubility surface based sensitivity no longer dominates the achievable results, because an early primary nucleation can be easily obtained by using a variable cooling rate or antisolvent addition rate (e.g. initial fast cooling followed by a slow cooling period), while the excessive nucleation can be avoided by decreasing the cooling (addition) rate immediately after nucleation occurred. When the cooling/addition rates are fixed this level of control on nucleation is not possible and a uniform rate of supersaturation generation is

54 found that is a result of the compromise between triggering fast the crystallization and avoiding excessive nucleation Since both of these kinetic phenomena also depend on the temperature in addition to the supersaturation in some cases this effect becomes more dominant. According to Table 3.1, the activation energy for the nucleation rate is three times higher than the activation energy for the growth rate. As a result, the nucleation rate will be much more sensitive to temperature change than the growth rate. In order to obtain a gentle well controlled primary nucleation, as well as to avoid nucleation during growth, a lower temperature is preferred, which can be achieved during the initial phase of the process, before the onset of primary nucleation. Thus, from this point of view, cooling is better than antisolvent to generate supersaturation for primary nucleation at the beginning of batch, to also bring the temperature to a level where nucleation rate can be better controlled relatively to the growth. However, to be able to achieve the early nucleation generation (for longer growth time) and the fine control of nucleation rate a varying cooling rate is necessary. As indicated by the optimal cooling profiles in Figure 3.14, the initial cooling rate is large followed by a period of practically no cooling. When the simplified scenario, with fixed cooling rate, is used for optimization this type of combined fast nucleation generation followed by gentle nucleation procedure cannot be achieved in both cases, hence the optimal operating procedures are different in those cases. This can be observed when both the full optimal and simplified trajectories corresponding to the initial points A and B are plotted in Figure 3.16 for comparison.

55

Solubility [g /g solvent]

B 0.6 S2 A opt 0.5 (optimal) S2 50 (suboptimal) 0.4 45 S1 S1 0.3 opt C] (suboptimal) 40  100 (optimal) 0.2 C 80 35 eth an [% g/g solvent] ol 60 30 0.1 40 C Temperature [ 25

Figure 3.16. Optimal trajectories (bold solid line) vs. suboptimal trajectories (thin dash line) for S1 and S2.

When starting from a temperature sensitive region, both the optimal and the suboptimal trajectories perform cooling first, to ensure an early and gentle, well-controlled primary nucleation. However, when starting from antisolvent concentration sensitive region, the simplified trajectory cannot guarantee early primary nucleation and low temperature growth at the same time. The results provide an illustrative example how the proposed framework using the solubility surface and property map based simplified design of CCAC systems can be used for the rapid determination of simplified operating procedures for

CCAC processes, but also illustrate that full optimization of nonlinear trajectory can yield better product properties (see Table 3.7) due to the finer control of nucleation that can be achieved by the time varying cooling and antisolvent addition rates. Because the differences between final crystal qualities achieved using the simplified CCAC design procedure and the full optimization are due to the requirement to initiate the crystallization

56 process as soon as possible via controlled nucleation, therefore it is expected that in the case of seeded CCAC processes the differences between the approaches would be less significant.

Table 3.7. Comparison of final product properties between suboptimal and optimal trajectories.

S1 S2 Suboptimal Optimal Suboptimal Optimal

Lmn [] 529 722 527 878 CV..[] 0.284 0.251 0.238 0.232

3.5 Conclusions

A novel model-based systematic analysis and design approach for unseeded combined cooling and antisolvent crystallization processes has been proposed and evaluated for the crystallization of aspirin in ethanol-water mixture. The cooling rate and antisolvent addition rate were kept constant to ensure that the operating curves are easy to implement in an industrial setup, whereas the cooling time and antisolvent feeding time were optimized to obtain large crystals. A strategy with simultaneous constant cooling and antisolvent addition, where cooling time and antisolvent feeding time were equal, was used as the base strategy for comparison. The simplified operating procedure allows the construction of operating maps for the combined cooling and antisolvent crystallization systems that can be used for better understanding of the conditions that govern these systems and for the rapid design of operating procedures that can lead to desired product

57 property. The advantage of the proposed methodology was presented through both simulations and experimental investigations, indicating that the base strategy could be further improved considerably. According to the proposed analysis, based on the sensitivity of the solubility surface, the simplified operating procedures should either belong to cooling first or antisolvent first strategy, depending on whether the initial solution locates at temperature sensitive region or antisolvent concentration sensitive region. The design based on the full optimization of nonlinear trajectory can provide further significant performance improvement due to the finer control of nucleation rate that can be achieved through the time-varying cooling and antisolvent addition rates. Although the proposed approach may have limitation in accuracy in the case of disturbances and model uncertainties due to the limitations of offline computed open-loop control trajectories, it provides an efficient and rapid experimental design method to significantly reduce the number of experiments required for the design of CCAC systems.

58

CHAPTER 4. COMBINED COOLING/ANTISOLVENT CRYSTALLIZATION IN

MIXED SUSPENSION MIXED PRODUCT REMOVAL CASCADE

CRYSTALLIZERS: STEADY-STATE AND STARTUP OPTIMIZATION8

4.1 Introduction

It is always important to optimize the steady-state operating conditions in the continuous crystallization process in order to obtain crystals of desired size, shape, etc. In addition, the startup problem is also crucial in the continuous crystallization, since the crystal properties are inconsistent before the continuous system can reach steady-state, and the time and the products are wasted during startup. Hence feasible methods to shorten the startup duration time and reduce the waste of material in the continuous crystallization systems are highly desirable. However, currently there are very few works reported to control and optimize the startup behavior of continuous crystallization system such as the MSMPR cascade (Su et al., 2015).

In this work, both the steady-state and the startup behaviors of CCAC in multistage

MSMPR cascade were evaluated and optimized through a model-based framework. The

PBE model was used and solved using the method of moments. Firstly, the optimal steady- state operating trajectories were obtained and compared with the optimal batch trajectory.

8 Reproduced with permission from Yang, Y., Nagy, Z.K., 2015. Ind. Eng. Chem. Res. 54, 5673–5682. Copyright 2015 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/ie5034254

59

Then different methods to shorten the startup duration and reduce the amount of waste were discussed. The methods include changing the initial solution compositions, applying dynamic antisolvent profiles, applying dynamic temperature profiles, and applying dynamic antisolvent/temperature combined profiles. Lastly, a concept of implementing model-free approach to optimize the startup behavior was proposed. Aspirin was used as the model compound in the case study, and ethanol and water were used as solvent and antisolvent, respectively.

4.2 Mathematical Model

In this work, a continuous N stage MSMPR cascade system is modeled via population balance model. All the vessels are assumed to be well-mixed. The compositions of the outlet flow are identical to the slurry inside the crystallizer. The solvent/antisolvent mixing and the crystallization process do not affect the solution volume. The crystal breakage and agglomeration are assumed negligible. Then the PBE for the ith stage could be written as

(Ramkrishna, 2000; Rawlings et al., 1993):

V n(,)(,) L t V G n L t i iF(ii 1) n( L , t )  F ( ) n ( L , t )  i i i , i  1,2,..., N (4.1) tLout i1 out i

3 where ni (,) L t is the volumetric number density (#/cm ) at stage i , L is the characteristic

-1 3 crystal size (cm), Gi is the growth rate (cm min ) at stage i , Vi (cm ) is the solution volume at stage i , F ()i is the outlet volumetric flow rate (cm3 min-1) of stage i , t is time out (min).

60

Equation (4.1) could be solved by the method of moments (Randolph, 1988). The jth moment of stage i is represented by:

  ()ijt Lnt(,)L dL, j  0,1,  ,  (4.2) j    i 0

 Take  xj ( eqn 4.1) dx , j  1. To make the equations mathematically simple, assume the 0

growth rate Gi is size-independent, and the solution volume Vi is constant. This assumption can be easily implemented in practice by using a level controller in the vessel.

Then the following equation could be obtained:

d()i Vj  F(1)(1)i i  F ()() i  i  VG j  () i , j  1 (4.3) idt out j out j i i j1

Rearrange equation (4.3):

()i (ii 1) ( ) d j FVFout i1 (i 1) out ( i ) ( i ) j   j jG i  j1, j  1 (4.4) dt Vi1 V i V i

Define the residence time and dilution factor at stage i as  i and  i , respectively:

Vi i  ()i (4.5) Fout

Vi1 i  (4.6) Vi

Then equation (4.4) could be written as the following ordinary differential equations

(ODEs):

()i d j 11(i 1) ( i ) ( i ) j   j jG i  j1 ,1  j  (4.7) dt i11  i  i

61

Similarly, the zeroth moment could be written as:

()i d0 11(ii 1) ( ) 00   Bi (4.8) dt i11  i  i

3 -1 where Bi represents the nucleation rate (#/cm min ) at stage i .

The mass balance equation for the solute at stage i was derived as:

dc d ()i Vi  F(ii 1) c  F ( ) c  k V 3 (4.9) idt out i1 outi vci dt

where ci is the solute concentration (g solute/g solvent mixture) at stage i . According to equations (4.2), (4.5) and (4.6), equation (4.9) could be simplified as:

dci c i1 c i ()i   3kGv c i 2 (4.10) dt i11  i  i

where kv is the volume shape factor, c is the crystal density. In this study, they are fixed at /6, 1.4 g/cm3, respectively (Lindengerg et al., 2009).

The mass balance equation for the solvent ethanol at stage i was derived as:

dv Vi  F(ii 1) v F ( ) v (4.11) idt out i1 out i

where vi is the volume fraction of solvent in the solvent mixture. Similarly, it could be simplified as:

dv v v i i1 i (4.12) dt i11  i  i

In this study, empirical equations (3.12), (3.13) and (3.14) were used to express nucleation rate, growth rate and relative supersaturation of aspirin in ethanol-water mixture.

The solubility is calculated based on Table 3.2 and Figure 3.1 (Lindenberg et al., 2009;

Yang and Nagy, 2014).

62

4.3 Results and Discussion

4.3.1 Steady-State Optimization

Figure 4.1 presents the schematic of using both cooling and antisolvent addition to generate supersaturation in multistage MSMPR cascade crystallizers. The feed solution is free of crystals so that the first stage is a nucleation stage while the rest stages are growth stages. Since there are two means to manipulate the supersaturation driving force, one could consider using temperature only, or antisolvent only, or both to control the supersaturation level for a certain stage.

Figure 4.1. Schematic of performing CCAC in a multistage MSMPR cascade.

The continuous MSMPR cascade system could finally reach steady-state and produce consistent products under certain fixed operating conditions. In this study, the optimal steady-state operating profiles of MSMPR cascade with N stages were obtained, using the maximum mean size as the objective. The temperature and the ethanol volume fraction of

 the feed solution are fixed as (T0 40 C, v0  79.2% ) (79.2% in volume = 75% in mass), respectively. Whereas the temperature and the ethanol volume fraction of the final stage,

63

 or Nth stage are fixed as (TN 25 C, vN  45.8% ), (45.8% in volume = 40% in mass),

respectively. The volumetric flow rate of the feed solution is a constant of F0  5 mL/min .

The optimization problem can be written as:

()()()NNN max {Jn. m . s .  L n  1 / 0 }, 1  i  N (4.13) TFi,, i i

subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10) and (4.12)

TTTT0 1  2    N (4.14)

 T0 40 C (4.15)

 TN 25 C (4.16)

N Fi 3.65 ml/min (4.17) i1

N i  30 min (4.18) i1

-1 where Ti is the temperature ( C ) at stage i , Fi is the antisolvent addition rate (ml min )

at stage i . They are the optimization variables. TN is the temperature (°C ) of the final stage or N th stage. The inequality constraint (4.14) limits the cooling rate to be nonnegative.

The equality constraints (4.15), (4.16) fix the feed solution temperature and the final temperature at stage N , while (4.17) fixes the final solvent composition at stage N . The last equality constraint (4.18) fixes the total residence time as 30 minutes. The objective function J is some desired crystal properties at stage N. In this study the number-average

mean size Ln is considered. Width of CSD can be represented by number-based coefficient

64

of variation ( CVn ). However, since there is often a tradeoff between crystal mean size and

CVn , only single objective (crystal mean size) is included here in order to simplify the study. The optimization problem is solved using SQP algorithm. MATLAB function fmincon and default specifications are used. The above solution approach is used throughout the whole study. Since it is not practical to set up more than 3 stages in the cascade in the pharmaceutical industry, the scenarios when N 1, 2, 3 were computed. In addition, the batch system that has the same initial and final conditions as the feed solution and the N th stage of the cascade system, was compared as well. The batch time is 30 minutes, which is exactly the same as the total residence time of the cascade system. The initial solutions at each stages are set to have the same ethanol volume fraction (79.2%) as the feed solution. The solute concentration of the initial solutions is at saturation of 25 °C , which equals 0.222 g /g solvent mixture.

The optimal operating profiles were presented in Table 4.1. In a certain MSMPR system,

it could be observed that the optimal residence time  i would increase as the stage number i increases. This phenomenon does make sense, because in a typical crystallization operation, the time for growth should be much longer than nucleation in order to obtain large crystals. In a MSMPR cascade, the first stage generally acts as a nucleation stage whereas the rest of the stages are designed for growth. Also one could notice that the more stages are used in the cascade, the larger crystals with smaller size variations can be obtained. The batch system can achieve better products as compared to the cascade system, since it corresponds to the limiting optimal operating trajectory that could be achieved using an infinite number of MSMPRs in the cascade. The batch crystallization is operated

65 in equilibrium state whereas the continuous crystallization is operated in steady-state (Chen et al., 2011).

Table 4.1. Optimal operating temperatures and antisolvent addition rates for cascade of

N 1, 2, 3, and the batch system.

N 1 stage 2 stages 3 stages batch

o 25 25 26.09 - T1 [ C]

o 25 25 - T2 [ C]

o 25 - T3 [ C]

-1 3.65 1.05 0.00 - F1 [mL min ]

-1 2.60 1.60 - F2 [mL min ]

-1 2.05 - F3 [mL min ]

1 [min] 30 10.74 6.61 -

 2 [min] 19.26 11.37 -

3 [min] 12.03 -

V1 [ml] 259.5 64.98 33.05 -

V2 [ml] 166.60 75.04 -

V3 [ml] 104.06 -

Ln [μm] 257.39 369.55 449.90 704.69

CVn [-] 1.000 0.790 0.683 0.250

The optimal trajectories of the four scenarios in Table 4.1 were compared in the 3D phase diagram, as shown in Figure 4.2. It can be noticed that the optimal cascade trajectory becomes more and more similar as the optimal batch trajectory when N increases, especially in the x and y directions, as plotted in Figure 4.2(b). This phenomenon

66 demonstrates that the batch trajectory can be regarded as a trajectory of MSMPR cascade of infinite number of stages for both the simple 2D (i.e. cooling or antisolvent only crystallization), as well as for the 3D crystallization (i.e. CCAC). In other words, the knowledge of the optimal operating procedure for a CCAC process in batch system could be transferred to a close to equivalent CCAC operation in MSMPR cascade system.

(a)

0.8

0.6

0.4

0.2

[g/g mixture]solvent

n

i

r

p s

A 0

C 100 Ethanol mass fraction [-] 50 75 45 40 50 35 30 C] 25 25 Temperature [

(b) Solubility [g/g solvent mixture]

0.6

0.5 50 0.4 45

C] 0.3 100 40  0.2 Ethanol mass75 fraction [-] 35 30 0.1 50 Temperature [

25 25

Figure 4.2. Optimal trajectories of 1 stage (dotted line), 2 stages (dash-dotted line), 3 stages

(dash line) cascade system and batch (solid line) system, shown in (a) 3D phase diagram and (b) its 2D projection.

67

For instance, if it is already known that cooling first is desired in CCAC in a batch system to produce large crystals, then it is very likely that cooling first should also be preferred in CCAC in MSMPR cascade system. This would be implemented by operating the first stage as a cooling crystallization and the second as antisolvent. If the batch operation requires simultaneous cooling and antisolvent addition, then most likely both stages of the MSMPR cascade would be operated using cooling as well as antisolvent addition.

4.3.2 Startup Optimization

The startup problem is very important in continuous manufacturing. The crystal properties are inconsistent before the continuous system can reach steady-state. As a result, the time and the products are wasted during startup. Therefore, the development of feasible methods to shorten the startup duration time and reduce the waste in the continuous crystallization systems can have high impact in continuous crystallization development and operation.

4.3.2.1 Startup Optimization via Using Different Initial Solutions

Typically, it is necessary to prepare some initial solutions for each stage before one turns on a continuous MSMPR cascade system. Theoretically the compositions of the initial solution at each stage should have some impact on the startup dynamic behavior, and maybe even the final steady-state product quality if multiplicity exists in the model. Hence the multiplicity of the MSMPR cascade when N equals 1, 2 and 3 was studied first. To

68 simplify the analysis, it is assumed that the initial solutions for each stage in a cascade are the same when N equals 2 or 3. Then the process is simulated via equations (3.12), (3.13),

(3.14), (4.7), (4.8), (4.10) and (4.12), with different initial solvent volume fraction and solute concentration. The results are plotted in Figure 4.3, which indicates that the cascade system would always reach the same steady-state independently of the composition of the initial solutions in the vessels. Hence there is no multiplicity in the MSMPR cascade system in our case. In other words, the cascade system is robust to the variation of initial conditions.

It also indicates that the results of the steady-state optimization are independent of how the initial solution compositions are chosen. However, the initial conditions can provide a way to shorten the startup duration time and reduce the waste by optimizing the compositions of the initial solutions.

Then a 2 stage MSMPR cascade system is modeled here, since N  2 is the most often used configuration in the pharmaceutical industry considering both performance and the operation complexity. The solvent volume fraction and solute concentration of the initial solution at each stage are the optimization variables. The objective is the minimum t99 or t95, which can be defined as the time required for the system to reach a controlled state of operation (CSO) that has less than 1% or 5% variation. The optimization problem can be written as:

min t99 , 1  i  N (4.19) cvii,

subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10) and (4.12)

0vv12 , 1 (4.20)

69

(a)

500 400

300

m]

 [

n 200 L 100 0 0.4 1 c 1 (initial) 0.2 [g/mL solvent] 0.75 0 0.5 (initial) [-] v1

(b)

500 400

300

m]

 [

n 200 L 100 0 0.4 1 c 1 ,c 0.2 [g/mL solvent]2 (initial) 0.75 0 0.5 ,v (initial) [-] v1 2

(c)

500 400

300

m]

 [

n 200 L 100 0 0.4 c 1 1 ,c 0.2 [g/mL solvent]2 ,c 0.75 3 (initial) (initial) [-] 0 0.5 ,v ,v3 v1 2

Figure 4.3. Final crystal mean size under various solvent volume fractions and solute concentrations of the initial solution, for cascade that has different number of stages: (a)

N 1, (b) N  2 , (c) N  3.

70

0c1 solubility ( T 1 , v 1 ) (4.21)

0c2 solubility ( T 2 , v 2 ) (4.22)

where v1 , v2 , c1 , c2 are the ethanol volume fraction and solute concentration of stages 1 and

2, respectively. The inequality constraints (4.20), (4.21), (4.22) can ensure that the optimization variables are in a proper range, and the solutions can be easily prepared in practice (no supersaturated initial solution is allowed).

A constant operating profile obtained from Table 4.1, when N  2 is applied in this section. A scenario with the initial solution compositions discussed in section 4.3.1 is also used here as base case for comparison ( v  79.2%, c  0.222 g/g solvent mixture or 25 °C saturated solution). This solution is used as the initial solution for both stages.

The optimization results are shown in Table 4.2 and Figure 4.4. The waste material produced in each scenario (amount of solute crystallized until the controlled state of operation is achieved) is also computed using the following equation:

t99 Waste ( t )   F()()NN ( t )  ( t ) dt (4.23) 99 v c out 3 0

The optimal case in Table 4.2 and Figure 4.4 indicates an improvement on the startup

duration time t99 as well as Waste ( t99 ) . However, the dynamic behavior of Ltn ( ) didn't change too much, in terms of the overshoots and the oscillations. Additionally, it can be noticed that in the optimal case, the initial concentration of the second stage is much smaller than the first stage. This is because the second stage, or the growth stage needs relatively smaller supersaturation, as compared with the first stage or the nucleation stage.

71

Table 4.2. Comparisons between the compositions, t99 and Waste ( t99 ) of the base case

and the optimal case, which uses min{t99 } as objective.

Base case ()t99 Optimal case ()t99

c1 [g/g solvent mixture] 0.222 0.201

v1 [-] 79.2% 74.7%

c2 [g/g solvent mixture] 0.222 0.026

v2 [-] 79.2% 81.8%

t99 [min] 143.1 106.7

Waste ( t99 ) [g] 156.6 94.0

600 600 (a) Waste (t ) = 156.6 g (b) Waste (t ) = 94.0 g 500 99 500 99

400 400

m] m] 

 300 300

[ [

n n

L L 200 200 t = 143.1 min t = 106.7 min 99 99 100 100 stage 1 stage 1 stage 2 stage 2 0 0 0 50 100 150 200 250 300 0 50 100 150 200 250 300 Time [min] Time [min]

Figure 4.4. Mean size Ltn ( ) and Waste ( t99 ) (a) before and (b) after optimizing the

compositions of the initial solutions when using min t99 as objective.

In practice, instead of 1% deviation, sometimes 5% deviation is allowable. Hence,

min{t95 } is also considered here as the optimization objective, which can be written as:

min t95 , 1  i  N (4.24) cvii,

72 where the constraints remain the same as in equations (4.20), (4.21) and (4.22). Then the

waste that uses t95 as criteria should be defined as:

t95 Waste ( t )   F()()NN ( t )  ( t ) dt (4.25) 95 v c out 3 0

The optimization results using min{t95 } as objective are listed in Table 4.3 and Figure

4.5. A similar improvement could be observed as using min{t99 } as objective. The

difference is that, when using t95 the improvement is more significant. This can be explained by the larger flexibility for optimization offered by the looser limits that define the controlled state of operation with 5% deviation as compared to the case with 1% deviation.

Table 4.3. Comparisons between the compositions, t95 and waste amount of the base case

and the optimal case, which uses min{t95 as objective.

Base case ()t95 Optimal case ()t95

c1 [g/g solvent mixture] 0.222 0.207

v1 [-] 79.2% 75.0%

c2 [g/g solvent mixture] 0.222 0

v2 [-] 79.2% 73.4%

t95 [min] 100.1 59.9

Waste ( t95 ) [g] 97.2 32.6

73

600 600 (a) Waste (t ) = 97.2 g (b) Waste (t ) = 32.6 g 500 95 500 95

400 400

m] m] 

 300 300

[ [

n n

L L 200 200 t = 100.1 min 95 t = 59.9 min 100 100 95 stage 1 stage 1 stage 2 stage 2 0 0 0 50 100 150 200 250 300 0 50 100 150 200 250 300 Time [min] Time [min]

Figure 4.5. Mean size Ltn ( ) and Waste ( t95 ) (a) before and (b) after optimizing the

compositions of the initial solutions when using min{t95 } as objective.

4.3.2.2 Startup Optimization via Applying Dynamic Operating Profiles

Typically, the continuous MSMPR cascade system should be operated under some constant operating conditions in order to maintain steady-state and consistent product properties. However, during the startup it might be helpful to apply dynamic operating profiles to shorten the startup duration and reduce the waste. In the dynamic startup study a 2 stage MSMPR cascade system is investigated. The computation of the optimal dynamic temperature or/and antisolvent addition profiles are formulated as a nonlinear optimization problem solved using a SQP approach. Since the startup duration time is typically around

100 minutes as shown before, a time period of 300 minutes is prepared for observation.

The 300 minutes’ time period is then divided into 30 intervals, and the temperatures and the antisolvent addition rates at every discrete time are considered as the optimization variables.

74

4.3.2.2.1 Scenario 1: Dynamic Antisolvent Addition Profiles With Minimum t99 as

Objective

If using t99 as criteria and dynamic antisolvent flow rates with constant temperatures as operating profiles, the optimization problem can be written as:

min t99 , 0  k  30 (4.26) F12( k ), F ( k ) subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10) and (4.12)

0F1 ( k )  10,  k  0,1,...,15 (4.27)

0F2 ( k )  10,  k  0,1,...,15 (4.28)

F1( k ) F 1,ss ,  k  16,17,...,30 (4.29)

F2( k ) F 2,ss ,  k  16,17,...,30 (4.30)

where Fk1 (), Fk2 () are the antisolvent addition rates at the kth discrete time for stage 1

and 2, respectively. F1,ss , F2,ss are the steady-state operating values obtained from Table

4.1. Inequality constraints (4.27), (4.28) limit the antisolvent addition rate in the proper operating range. Equality constraints (4.29), (4.30) force the antisolvent addition rates go back to the desired steady-state values, in order to ensure that the dynamic antisolvent profiles won't affect the product properties at the final steady-state. The initial solution compositions are the same as section 624.3.1, or the base case in section 4.3.2.1. All the other variables, such as the temperatures, the volumes, the feed solution flow rate and composition, remain the same as the steady-state values listed in Table 4.1. Thus, the base case in Table 4.2 and Figure 4.4(a) can also be used as the base case here for comparison.

75

The optimal dynamic antisolvent profiles when using t99 as criteria are plotted in Figure

4.6. The antisolvent flow rates during startup are larger than the steady-state values, which makes sense because high supersaturation is necessary in order to make the system crystallize more quickly and strongly during startup than in the steady-state. The dynamic

behavior of the mean size Ltn ( ) shows excellent trajectory, with almost no overshoot and

oscillations anymore. The t99 and Waste ( t99 ) in Figure 4.6 are also reduced significantly, as compared to the base case (Figure 4.4(a)). Thus unlike changing the initial solution composition, the dynamic profiles can have much stronger influence on the dynamic behavior during startup. In addition, another scenario that uses min Waste ( t ) instead of 99 min{t99 } as objective is also computed. Finally, a similar optimal result as in scenario 1 is

obtained. This is because overshoot and oscillations are eliminated, hence the less t99 is,

the less Waste ( t99 ) would be. Thus it is reasonable to consider min t99 as the only objective.

600 4 (a) stage 1 (b) Waste (t ) = 80.4 g 500 99 2 400

0 Flow rate [ml/min]Flow

0 50 100 150 200 250 300 m]

 300

Time [min] [ n L 4 (a) stage 2 200 t = 83.0 min 100 99 2 stage 1 stage 2 0 0 Flow rate [ml/min]Flow 0 50 100 150 200 250 300 0 50 100 150 200 250 300 Time [min] Time [min]

Figure 4.6. (a) Optimal dynamic antisolvent addition profiles, (b) corresponding mean size

Ltn ( ) and Waste ( t99 ) when using min t99 as objective.

76

4.3.2.2.2 Scenario 2: Dynamic Temperature Profiles With Minimum t99 as Objective

Alternatively, dynamic temperature profiles could be applied instead of dynamic antisolvent addition profiles. The optimization problem could be formulated in a similar way:

min t99 , 0  k  30 (4.31) T12( k ), T ( k ) subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10) and (4.12)

15T1 ( k )  40,  k  0,1,...,15 (4.32)

15T2 ( k )  40,  k  0,1,...,15 (4.33)

T1( k ) T 1,ss ,  k  0,16,17,...,30 (4.34)

T2( k ) T 2,ss ,  k  0,16,17,...,30 (4.35)

where Tk1 (), Tk2 () are the temperatures at the kth discrete time for stage 1 and 2, respectively. Equations (4.34) and (4.35) contain more equality constraints than (4.29) and

(4.30). In (4.34) and (4.35), the equality constraints include TT1(0)  1,ss and TT2(0)  2,ss ,

whereas in (4.29) and (4.30) of scenario 1, FF1(0)  1,ss and FF1(0)  1,ss are not listed as constraints. This is because in practice, the antisolvent flow rates could achieve a large step change within a negligible time period, whereas the temperatures couldn't.

The optimization results of scenario 2 are plotted in Figure 4.7, in which some similar phenomena could be observed as scenario 1. The temperatures during startup are lower than the steady-state values, resulting in relatively higher supersaturation and quicker crystallization during startup. The optimal performance in scenario 2 is also very close to

77 the performance in scenario 1, and both of these are much better than the base case (Figure

4.4(a)).

30 600 C]  (a) stage 1 (b) Waste (t ) = 81.0 g 25 99 500 20 400

Teperature Teperature [ 15

0 50 100 150 200 250 300 m]

Time [min]  300

[ n

L C] 30  (a) stage 2 200 25 t = 83.1 min 100 99 20 stage 1 stage 2 15 Temperature [ 0 0 50 100 150 200 250 300 0 50 100 150 200 250 300 Time [min] Time [min]

Figure 4.7. (a) Optimal dynamic temperature profiles, (b) corresponding mean size Ltn ( )

and Waste ( t99 ) when using min{t99 } as objective.

4.3.2.2.3 Scenario 3: Dynamic Combined Profiles With Minimum t99 as Objective

In this case startup optimization is considered using dynamic antisolvent profiles and dynamic temperature profiles simultaneously. The optimization problem using the dynamic combined profiles can be written as following:

min t99 , 0  k  30 (4.36) F1(), k F 2 (), k  T 1 (), k  T 2 () k subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10) and (4.12)

constraints (4.27)–(4.30), (4.32)–(4.35)

The optimization results of scenario 3 are shown in Figure 4.8. The performance and conclusions are generally similar to the previous two scenarios.

78

4 (a) stage 1

2

0 Flow rate [ml/min]Flow 0 50 100 150 200 250 300 Time [min]

4 (a) stage 2

2

0 Flow rate [ml/min]Flow 0 50 100 150 200 250 300 Time [min]

30 (b) stage 1 25

20

15

Teperature Teperature [ml/min] 0 50 100 150 200 250 300 Time [min] 30 (b) stage 2 25

20

15 0 50 100 150 200 250 300 Temperature [ml/min] Time [min]

600 (c) Waste (t ) = 77.4 g 99 500

400 m]

 300

[

n L 200 t = 80.1 min 99 100 stage 1 stage 2 0 0 50 100 150 200 250 300 Time [min]

Figure 4.8. (a) Optimal dynamic antisolvent addition profiles, (b) optimal dynamic

temperature profiles, (c) corresponding mean size Ltn ( ) and Waste ( t99 ) when using

min{t99 } as objective.

79

The dynamic antisolvent flow rates here have much stronger effect on decreasing startup duration than the dynamic temperatures. This is because the antisolvent flow rates can also affect the residence time whereas the temperatures cannot. Considering the dynamic operation complexity and cost, thus a practical way to implement this design is to apply dynamic antisolvent flow rates only, while keeping the temperature profiles constant, as in scenario 1.

4.3.2.2.4 Scenario 4: Dynamic Combined Profiles With Minimum t95 as Objective

As shown in the previous section the antisolvent flow rates have stronger impact on the system than the operating temperatures. Thus a scenario using dynamic antisolvent profiles

only and min{t95 } as objective was considered. The only difference between scenario 1 and scenario 4 is the objective function. All the other optimization formulations are the same as in scenario 1.

600 4 (a) stage 1 (b) Waste (t ) = 6.6 g 95 500 2 400 0

Flow rate [ml/min]Flow 0 50 100 150 200 250 300 m]

Time [min]  300

[ n

L 4 (a) stage 2 200 t = 27.5 min 2 100 95 stage 1 stage 2 0 Flow rate [ml/min]Flow 0 0 50 100 150 200 250 300 0 50 100 150 200 250 300 Time [min] Time [min]

Figure 4.9. (a) Optimal dynamic antisolvent addition profiles, (b) corresponding mean size

Ltn ( ) and Waste ( t95 ) when using min{t95 } as objective.

80

The optimal results are plotted in Figure 4.9, which indicates a great improvement on startup duration time and waste when comparing with the base case (Figure 4.5(a)). But since 5% deviation still allows room for overshoot and oscillation, they can hardly be

eliminated when setting min{t95 }as objective.

Finally, the performance of all scenarios and base cases are summarized in Table 4.4 for comparison, from which at least about 50% improvement can be observed when using dynamic operating profiles during startup.

Table 4.4. Comparisons between the performance of all scenarios and base cases.

t99 [ min ] Waste ( t99 ) [g] t95 [min] Waste ( t95 ) [g] Without dynamic profiles (base case) 143.1 156.6 100.1 97.2 Dynamic antisolvent profiles (scenario 1, 83.0 80.4 27.5 6.6 4) Dynamic temperature profiles (scenario 2) 83.1 81.0 -- -- Dynamic combined profiles (scenario 3) 80.1 77.4 -- --

4.3.2.3 Startup Optimization via Model-Free Approach

Typically, the methods to control crystallization can be distinguished as model-based approach and model-free (direct design) approach (Nagy and Braatz, 2012; Nagy et al.,

2013). In the previous discussion a model-based approach is used, which is an offline control method and it may not be robust to the outside disturbances as well as the model uncertainties. Thus a model-free, or direct design approach can also be considered. In the model-free approach of the batch system, typically two methods are used via closed-loop control. One is the DNC, which uses FBRM as the sensor to control the crystal counts in

81 the slurry in a certain range (Abu Bakar et al., 2009; Saleemi et al., 2012a, 2012b). The other method is the SSC, which uses ATR spectroscopy to measure the solute concentration and controls the supersaturation level to follow a certain trajectory (Saleemi et al., 2012b;

Woo et al., 2009). These two methods have been demonstrated to be very efficient in the control of batch operation. Two similar methodologies can be proposed in the MSMPR cascade system to improve the startup behavior. By using FBRM as the sensor in the closed-loop, the crystal counts are measured in real time. When the counts are lower than the desired steady-state value, the control loop could result in a higher antisolvent addition rate or lower temperature to increase the supersaturation level, and vice versa. By using

ATR spectroscopy, a higher antisolvent addition rate or lower temperature should be actuated when the supersaturation is measured lower than the desired steady-state value, and vice versa. These two model-free approaches are promising in the automatic control of the startup behavior of continuous MSMPR cascade, and are being evaluated in practice at the moment.

4.4 Conclusions

A continuous MSMPR cascade system of different number of stages has been modeled and utilized to optimize both the steady-state and startup behaviors when both cooling and antisolvent addition are applied as the crystallization driving forces. Aspirin-ethanol-water is used as the model system in the study. In the steady-state optimization, the optimal steady-state operating profiles have been obtained. In addition, the optimal trajectories for both MSMPR cascade and batch systems have been plotted together in 3D phase diagram, from which the fact that the batch trajectory could be regarded as a MSMPR cascade of

82 infinite number of stages has been observed in 3D crystallization (i.e. CCAC) for the first time. Besides, different methods to shorten the startup duration time and reduce the waste have been discussed and proved to be feasible. These methods include changing the initial solution compositions, applying dynamic antisolvent profiles, applying dynamic temperature profiles, and applying dynamic antisolvent/temperature combined profiles.

The most efficient and practical way to minimize startup duration and waste is to use dynamic antisolvent profiles only during startup. It was shown that this approach could reduce the startup duration and the amount of waste with approximately 50% while having no impact on the final steady-state product properties. Finally, a model-free approach to optimize the startup behavior is proposed using feedback control through FBRM or ATR spectroscopy techniques.

83

CHAPTER 5. ADVANCED CONTROL APPROACHES FOR COMBINED

COOLING/ANTISOLVENT CRYSTALLIZATION IN CONTINUOUS MIXED

SUSPENSION MIXED PRODUCT REMOVAL CASCADE CRYSTALLIZERS9

5.1 Introduction

While model-based optimization, design and experimental investigation of continuous single stage and multistage MSMPR systems is becoming more common in the literature, currently there are no published papers working on model-based advanced control approaches for CCAC in MSMPR cascade crystallizers, with both crystal size and yield controlled at the same time. To achieve this goal, this chapter systematically studied the feasibility of two different control schemes: decentralized PID control and nonlinear model predictive control (NMPC). For decentralized PID control approach, local linearization method and relative gain array (RGA) analysis were used to examine its feasibility.

Whereas for NMPC approach, its performance was evaluated under both servo control and regulatory control scenarios. In this work, not only nucleation and growth kinetics, but also controlled dissolution kinetics were considered. In the case study, aspirin (acetylsalicylic acid), ethanol and water were used as model compound, solvent and antisolvent, respectively.

9 Reproduced with permission from Yang, Y., Nagy, Z.K., 2015. Chem. Eng. Sci. 127, 362–373. Copyright 2015 Elsevier. http://www.sciencedirect.com/science/article/pii/S0009250915000895

84

The schematic representation for a closed-loop hierarchical NMPC implementation structure in continuous two stage MSMPR cascade crystallizer is presented in Figure 5.1.

In principle, PAT tools like FBRM and UV/Vis spectroscopy can be used for online measurements of CSD and solute concentration, from which mean crystal size and yield can be inferred in real time. As a result, the process model can be updated during the process using those measurements via closed-loop optimal control. Sensor noise and error are not considered in this work. In addition, in practice typically low level flow rate and temperature controllers would be used in a hierarchical control structure with the NMPC, as shown in Figure 5.1. In this work, in order to simplify the simulation study, antisolvent addition rates and temperatures were manipulated directly from NMPC scheme.

NMPC FC FC

Feed solution Antisolvent Antisolvent

FBRM UV/Vis

LC LC

TC TC

Figure 5.1. Schematic representation of a hierarchical NMPC implementation structure.

FC, TC, LC are flow rate controller, temperature controller and level controller, respectively.

85

5.2 Mathematical Model

In this work, a continuous a two stage MSMPR cascade crystallizer is modeled using

PBM. Both the temperature and antisolvent addition rate can be manipulated at both stages, in order to control supersaturation levels and residence times (Figure 5.2). The crystal mean size and yield of the second stage are the two controlled variables. The following assumptions are used in this work: (1) vessel is ideally well-mixed (2) the outlet flow contains exactly the same compositions as inside the vessel (3) only nucleation, growth and dissolution are considered, whereas breakage or agglomeration are negligible (4) mixture of aspirin, ethanol and water is ideal solution.

The process is described using the same mathematical model as in chapter 4. In addition, in this chapter the crystal dissolution kinetic is also considered:

k D  kexp( G2 )( c* (1  S ))kG 3 , if S  1 (5.1) GG1 RT( 273)

k k Dk exp( B2 )exp(  B3 ), if S<1 (5.2) BB1 RTS( 273) ln2 (1/ ) where S is the supersaturation ratio [-], c* is the equilibrium solubility [g/ml], T is

o temperature [ C]. DG and DB are dissolution rates. When S 1, the negative values of nucleation rate and growth rate are used to approximate dissolution rates (Qamar et al.,

2010; Nagy et al., 2011). kG1 , kG 2 , kG3 , kB1 , kB2 , kB3 are empirical parameters from literature (Lindenberg et al., 2009).

Assume Ln and Y are number based mean crystal size [ m] and yield [-] of the second stage, respectively, calculated with the following equations:

86

(2) (2) Ln  10/ (5.3)

cF(2) Y 1 2 out (5.4) cF00

In this study, and model mismatch is not considered.

Feed solution Antisolvent F1 Antisolvent F2

Temperature T1 Temperature T2

Crystal mean size Ln Yield Y

Figure 5.2. Schematic representation of the continuous CCAC process in a two stage

MSMPR cascade crystallizer.

5.3 Results and Discussion

5.3.1 Attainable region of crystal size and yield

According to Figure 5.2, the proposed control structure can be considered as a multi- input multi-output (MIMO) system, as shown in Figure 5.3. Since in this process more potential manipulating inputs exist than desired controlled variables, it is interesting to evaluate different control architectures. Five control methods are investigated based on the inputs that are manipulated, as presented in Table 5.1.

87

The basic operating scenario was defined as: F1 , F2 , T1 , T2 equal 1.05 ml/min, 2.60

o o ml/min, 25 C, 25 C, respectively. It was found that crystal mean size Ln and yield Y would reach constant values of 369 µm and 0.685 after the system reaching steady-state, by simulating the process based on equations (3.12)–(3.14), (4.7), (4.8), (4.10), (4.12), and

(5.1)–(5.4).

F1 Two stage Ln F2 MSMPR cascade T1 crystallizer Y T2

Figure 5.3. Schematic representation of investigated multi-input multi-output system.

Table 5.1. Five possible different control approaches for two-stage MSMPR cascade crystallization systems.

Control method Controlled output variable Manipulated input variable

Nucleation control F1 , T1

Growth control F2 , T2

Ln Antisolvent control F1 , F2 Y Temperature control T1 , T2

Global control F1 , F2 , T1 , T2

For the manipulated input variables, the following operating limits were used:

0FF12 , 10 mL/min (5.5)

o 10TT12 , 40 C (5.6)

88

The attainable regions of crystal size and yield of the proposed five control methods can be obtained by simulating the process considering model equations and operating limits.

Since the process is highly nonlinear and dissolution is also considered, it is possible that the maximum and minimum crystal sizes corresponding to a certain yield may not guarantee that the crystal sizes in between are also achievable. Therefore, an exhaustive search method is used based on a Monte Carlo type simulation study. The process is simulated, and the crystal size and yield are recorded repeatedly, with each variable taking a value within the operating limits and changing with a small step in the next simulation.

The hierarchical for loops were implemented in MATLAB. Finally, the boundary of all recorded crystal sizes and yields were plotted, as presented in Figure 5.4. It should be acknowledged that the attainable regions (except global control) depend on the values of the unchanged variables, however the following phenomena are always observed. Due to the largest degree of freedom, the non-square global control method has the largest attainable region. Antisolvent and temperature control methods also have relatively large attainable regions, because they can manipulate supersaturation at both stages. Since nucleation and growth control methods can only well control one of the two stages, they have relatively small attainable regions. In addition, the Pareto front of crystal size and yield is clear in global control method, indicating that generally there is a compromise between crystal size and yield. Within the operating limits, large crystal size and high yield are difficult to achieve at the same time. Different control approaches enable to operate at different combinations of size and yield, and unless the global control is used changing the product (size) and process (yield) requirements may require the change of control

89 architecture. These attainable regions are used in the development of advanced control frameworks discussed later.

1 Nucleation Control 0.9 Growth Control 6 Antisolvent Control 0.8 Temperature Control 3 4 Global Control 2 0.7 8 1 0.6 9 0.5 5

Yield[-] 7 0.4

0.3

0.2

0.1

0 100 200 300 400 500 600 700 800 Mean size [m]

Figure 5.4. Attainable regions of crystal size and yield in the case of the proposed five control methods. Operating points 1–7 were used in decentralized PID control analysis.

Operating points 1, 8, 9 were used in NMPC analysis.

5.3.2 Analysis of System Dynamics

Before proposing the control approach, it is beneficial to investigate the basic dynamic features of the studied MIMO system (e.g. whether the system is linear or nonlinear, or how inputs and outputs interact). In this work, the response of each output variable is analyzed when step changes of each input variable are performed. As shown in Figure 5.5,

starting from basic operating scenario at 0 minute, F1 , F2 , T1 , T2 are increased by a step

90 change of 0.5 ml/min or 3 oC at 100 and 400 minutes, respectively. According to the response curves for crystal size and yield, it is clear that there is strong interaction among the four inputs and the two outputs. In addition, inverse responses and oscillations are present, and the system is nonlinear and complex. For example, the crystal mean size will

finally increase if T2 is increased by a small step, however it will decrease if T2 is increased by a large step (Figure 5.5(d)). The size increase is caused by the controlled dissolution and fines removal of mild heating at the second stage, whereas large particles are dissolved and crystal mean size is reduced if the second stage is heated too much.

(a) (b) 3 2 4 3 1 2

0 [ml/min]

[ml/min] 2

1 0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700

F

F

m]

 m]

 400 400 350 350 300 250 300 250

0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700

Mean size Mean [ size Mean size Mean [ size 0.72 0.69 0.7 0.685 0.68 Yield [-] Yield 0.68 Yield [-] Yield 0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700 Time [min] Time [min]

(c) (d)

C] 30 30

C]

[

[ 1

25 2 25 T

0 100 200 300 400 500 600 700 T 0 100 200 300 400 500 600 700

m] m]   400 380 300 370 360 200 350

0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700

Mean size Mean [ size Mean [ size

0.7 0.68 0.65

0.66 0.6

Yield [-] Yield Yield [-] Yield 0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700 Time [min] Time [min]

Figure 5.5. Response of crystal mean size and yield with respect to step change in (a)

antisolvent addition rate at stage 1 ( F1 ), (b) antisolvent addition rate at stage 2 ( F2 ), (c)

temperature at stage 1 (T1 ), and (d) temperature at stage 2 (T2 ).

91

5.3.3 Feasibility of Decentralized Proportional-Integral-Derivative (PID) Control

Decentralized PID control is a general control approach for MIMO control systems. It is relatively easy to implement since only standard PID controllers are required. Using decentralized control the interacting MIMO system is controlled using two single-input single-output systems. The design of the decentralized control requires pairing each controlled variable with manipulated variable. The feasibility of this approach was evaluated using local linearization method and relative gain array (RGA) analysis. Global control method is not discussed here since RGA analysis requires equal number of inputs and outputs.

Table 5.2. Selected operating points for local linearization and RGA analysis.

o o Control method Point F1 [ml/min] F2 [ml/min] T1 [ C] T2 [ C] Ln [µm] Y [-]

1 1.05 2.60 25.0 25.0 369 0.685 Nucleation control 2 3.00 2.60 15.0 25.0 214 0.740

1 1.05 2.60 25.0 25.0 369 0.685 Growth control 3 1.05 2.50 25.0 20.0 204 0.761

1 1.05 2.60 25.0 25.0 369 0.685

Antisolvent control 4 3.00 3.00 25.0 25.0 205 0.741

5 0.50 0.50 25.0 25.0 417 0.498

1 1.05 2.60 25.0 25.0 369 0.685

Temperature control 6 1.05 2.60 15.0 15.0 217 0.836

7 1.05 2.60 30.0 35.0 487 0.479

For the proposed other four control methods, firstly two or three operating points from their corresponding attainable regions were chosen. The value, corresponding operating

92 condition and exact position of these points are indicated in Table 5.2 and Figure 5.4. It should be noticed that point 1 is the basic operating scenario.

Assume the standard mathematical representation of a linear time invariant (LTI) state- space model is:

dx A  x  B  u (5.7) dt

y C  x  D  u (5.8) where x is the model state vector, u is input and y is output. In this work, the process model was linearized at different operating points mentioned in Table 5.2 using the linearization toolbox in Simulink, in which matrices A , B , C , D can be obtained. These matrices are then used to calculate the RGAs at those operating points using the following equation:

RGA(  CA1 B  D )  ((  CA  1 B  D )  1 )T (5.9) where A , B , C , D are matrices from LTI state-space model (5.7), (5.8). For convenience,

we write F1 , F2 , T1 , T2 , Ln , Y as u1 , u2 , u3 , u4 , y1 , y2 , respectively. Since there are only two inputs and two outputs, all the RGAs are two by two matrices. Define the elements of

RGA as:

mi mj RGA   (5.10) ni nj

(/)yum i u mi  (5.11) (/)yum i y

93

RGA can be used as a tool for control pairing. The pairing which has positive ij closest to one is the most desired. The final RGAs and pairings obtained at those operating points are summarized in Table 5.3.

Table 5.3. RGAs and input-output pairings for selected operating points of the four proposed control methods.

Control method Point RGA Pairing

0.733 0.267 yu11 1  0.267 0.733 yu23 11 13 Nucleation control  21 23 0.028 1.028 yu13 2  1.028 0.028 yu21

0.265 1.265 yu14 1   1.265 0.265 yu22 Growth control 12 14  22 24 2.734 1.734 yu12 3  1.734 2.734 yu24

0.532 1.532 yu12 1  1.532 0.532 yu21

11 12 0.865 0.135 yu11 Antisolvent control 4   21 22 0.135 0.865 yu22

0.708 0.292 yu11 5  0.292 0.708 yu22

0.026 0.974 yu14 1  0.974 0.026 yu23

13 14 0.643 0.357 yu13 Temperature control 6   23 24 0.357 0.643 yu24

0.920 0.080 yu13 7  0.080 0.920 yu24

The results indicate that for a certain control method, the feasible pairing between inputs and outputs depends on the position of the specific operating point. A change-over of the

94 control pairing is necessary when the operating point is changed. Thus decentralized PID control approach is not convenient to be implemented when a relatively large operating region is required, since suitable control would require change in input-output topology and use a variable control structure. This is caused by the significant interactions among inputs and outputs and strong nonlinearity of the process. These results demonstrate that conventional control of a cascade MSMPR may not be practically feasible in none of the configurations when changing operating conditions or product quality specifications are required and the use of more advanced control approaches such as NMPC may be required.

5.3.4 Feasibility of Nonlinear Model Predictive Control (NMPC)

5.3.4.1 NMPC Problem Formulation

Compared to decentralized PID control, NMPC is a more advanced control approach since it can handle complex and highly nonlinear systems. NMPC approach uses the current measurements and nonlinear model to calculate the predicted output variables in a future time period named as prediction horizon. The future input variables in the prediction horizon are optimized to hold the output variables close to the set-point. Typically, only the first point of each optimal input variable is implemented. This optimization is then repeated when the next measurement is available. In this work, sampling time ( ) was chosen as 2 minutes. Prediction horizon ( P ) or number of prediction intervals was chosen as 25. Control horizon ( M ) or number of control moves was chosen as 5. In this work, open-loop optimal control solution approach was used, and model mismatch was not considered. The optimization problem can be formulated as:

95

min J 0.5(1  w ) J1  0.5(1  w ) J 2  wJ 3 , i  1,2,3,4 (5.12) ui( t k ), u i ( t k ),..., u i ( t k T p )

subject to:

model equations (3.12)–(3.14), (4.7), (4.8), (4.10), (4.12), and (5.1)–(5.4),

tTkp J() y y2 dt (5.13) 1 1 1set tk

tT y kp J()()1set 22 y y dt (5.14) 2y  2 2set 2set tk

y1set 2' J3 ()() trace Q  u  u (5.15) F0

u1( tk )  u 1 ( t k ),..., u 1 ( t k  M  )  u 1 ( t k  ( M  1)  )  u ... (5.16) u4( tk )  u 4 ( t k ),..., u 4 ( t k  M  )  u 4 ( t k  ( M  1)  )

1 0 0 0 0 1 0 0 Q   (5.17) 0 0 0.2 0  0 0 0 0.2

ututMi( ) i ( k  ), ttM  k  (  1) ,..., tTi k  p ,  1,2,3,4 (5.18)

0u12 ( t ), u ( t )  10,  k  tk , t k  ,..., t k  T p (5.19)

10u34 ( t ), u ( t )  40,  k  tk , t k  ,..., t k  T p (5.20)

 ui( t   )  u i ( t )   , k  t k , t k   ,..., t k  T p ,  i  3,4 (5.21)

 ui( t k )  u i ( t k1 )  , i  3,4 (5.22)

where tk is the first time point of the k th optimization, Tp is the time length of prediction

horizon (50 minutes in our case), y1set and y2set are the set-points for crystal size and yield,

96 respectively. Equations (5.13), (5.14) are cumulative errors of crystal size and yield compared to their corresponding set-points. Equations (5.15), (5.16), (5.17) are penalty term for the inputs, which can help avoid sharp change in inputs and promote smooth input responses and closed loop stability, u is a matrix that contains variations of each input,

Q is a diagonal weighting matrix, as shown in equation (5.17) for the global control

method. In the case of the other four control methods, if ui is not used as manipulated

variable, then Qii becomes zero. To reduce the number of optimization variables while maintaining long enough prediction horizon for stability, inputs after the control horizon

(M) were set equal to the last values of inputs in the control horizon, as shown in equation

(5.18). Equations (5.19), (5.20) are operating limits for the inputs. Equations (5.21) and

(5.22) are the temperature rate limits that are set as ±1 oC/min. w is the weight for penalty term, which is chosen as 0.05. As shown in the objective function (5.12), crystal size and yield were set of equal weights. SQP technique was used to solve the optimization problem.

5.3.4.2 Servo Control

A servo control scenario was designed to evaluate the feasibility of NMPC using the five proposed control methods. As presented in Table 5.4, three set-points (points 1, 8 and

9) were selected from the attainable region (Figure 5.4) in order to evaluate NMPC performance when fast change-over among these set-points are required. The operating points 1, 8 and 9 correspond to three typical desired product properties: medium size and medium yield, high yield and small size, and large size with low yield. It should be noticed that points 1 and 8 belong to the attainable regions of all five control methods, whereas

97 point 9 is outside of the attainable regions of nucleation and growth control methods (Table

5.4). Points 1, 8 and 9 have similar positions as the operating points selected in the analysis of decentralized PID control, making the evaluations of decentralized PID control and

NMPC comparable. The results indicate that the NMPC performs well even in the case when decentralized PID control would require change in the control architecture. During time 0 to 50 minutes, the system was assumed to work at steady-state with operating point

1 used as set-point. Then the set-point was changed to point 8, point 1, point 9 at 50 minutes,

200 minutes, 350 minutes, respectively, as indicated in Table 5.4.

Table 5.4. Selected set-points to evaluate NMPC feasibility.

Time [min] Set-point

Point Belonged attainable region Mean size [µm] Yield [-] 0 - 50 1 Nucleation control, Growth control, Antisolvent control, Temperature control, 369 0.685 Global control 50 - 200 8 Nucleation control, Growth control, Antisolvent control, Temperature control, 200 0.720 Global control 200 - 350 1 Nucleation control, Growth control, 369 0.685 Antisolvent control, Temperature control, Global control 350 - 500 9 Antisolvent control, Temperature control, 450 0.550 Global control

The NMPC performance of the antisolvent control method is used as an example for detailed analysis. The controlled variables, manipulated variables, concentration and supersaturation ratio profiles, and CPU time are shown in Figure 5.6.

98

(a) (b) 500

10 m]

 300

) [ml/min] ) 1 Size[ 100 5 0 100 200 300 400 500

1 0

Stage 1 antisolvent1 Stage 0 100 200 300 400 500

0.75

addition rate (F additionrate Yield[-] 0.5

0 100 200 300 400 500 10 ) [ml/min] )

1.5 2 5

1 C.V. [-] C.V. 0.5 0

0 100 200 300 400 500 Stage 2 antisolvent2 Stage Time [min] 0 100 200 300 400 500 addition rate (F additionrate Time [min]

(c) (d) 5 stage 1 0.4 4.5 stage 2 0.3 4 0.2 0.1 3.5 0 3 Concentration [g/ml] Concentration 0 100 200 300 400 500 2.5

1.8 2

stage 1 [min] CPU time 1.6 stage 2 1.5 1.4 1 1.2 0.5 1 0 Supersaturation ratio[-] Supersaturation 0 100 200 300 400 500 0 100 200 300 400 500 Time [min] Time [min]

Figure 5.6. NMPC performance of the antisolvent control method: (a) crystal size, yield and coefficient of variation (C.V.) of the CSD (red dotted lines are set-points); (b) antisolvent addition rates at stage 1 and 2; (c) concentration and supersaturation ratio; and

(d) CPU time.

For the antisolvent control method, the NMPC scheme responds fast to the set-point change among points 1, 8, and 9, and finally maintained the system at steady-state at the new set-points (Figure 5.6(a)). Although coefficient of variation is not a controlled variable, it is monitored here in order to guarantee that after the crystal mean size reaches a steady- state value, the width of size distribution is also in steady-state. Additionally, the

99 manipulated input responses are generally smooth (Figure 5.6(b)). Figure 5.6(c) indicates that high supersaturation level is desired at the first stage when high yield and small crystal size are the control objectives, whereas low supersaturation levels at both stages are preferred when large crystal size and low yield are desired. In general, for most optimizations the CPU time was less than 2 minutes, which was smaller than the sampling time (Figure 5.6(d)). This indicates that the proposed framework is promising for practical implementation. However, it has to be mentioned that the solution approach for the NMPC optimization was based on simple sequential solution technique implemented in MATLAB.

The computational time can be reduced significantly using more efficient optimization (e.g. multiple shooting) and/or other software implementation such as OptCon (Mesbah et al.,

2011, 2012; Nagy et al., 2007; Simon et al., 2009a).

The NMPC performances of the other four control methods are shown in Figure 5.7 and

Figure 5.8. Generally, the NMPC works well for the operating points that are within the attainable region of a particular control approach. Since operating point 9 is outside of the attainable region of the growth and nucleation control approaches these are unable to bring the system to this operating point. The extra degrees of freedom in the case of global control method provides faster response however requires longer CPU time due to larger number of optimization variables. The feasibilities of NMPC of all five proposed methods are summarized in Table 5.5. According to Table 5.4 and Table 5.5, the necessary condition for a certain control method and NMPC scheme to be feasible is that the set-point must be within the attainable region of that control approach. The NMPC optimization problem will be unfeasible if the set-point is out of the attainable region. Thus knowing the attainable

100 region of crystal size and yield is important for choosing and designing a certain control approach.

(a) (b) 500

10 m]  300

Size[ 100 5

0 100 200 300 400 500 [ml/min]

1 F 1 0 0 100 200 300 400 500

0.75 Yield[-] 0.5 0 100 200 300 400 500 40

1.5 30

C]

[ 1

1 T 20 C.V. [-] C.V. 0.5 0 100 200 300 400 500 10 Time [min] 0 100 200 300 400 500 Time [min]

(c) (d) 500

10 m]  300

Size[ 100 5

0 100 200 300 400 500 [ml/min]

2 F 1 0 0 100 200 300 400 500

0.75 Yield[-] 0.5 0 100 200 300 400 500 40

30

1.5 C]

[ 2

1 T 20 C.V. [-] C.V. 0.5 0 100 200 300 400 500 10 Time [min] 0 100 200 300 400 500 Time [min]

Figure 5.7. NMPC implementations: (a) controlled variables of nucleation control; (b) manipulated variables of nucleation control; (c) controlled variables of growth control; (d) manipulated variables of growth control (red dotted lines are set-points). C.V. is coefficient of variation of the CSD.

101

(a) (b)

500 40 m]  300 30

Size[ 100

0 100 200 300 400 500 [ml/min] 20

1 T 1 10 0 100 200 300 400 500

0.75 Yield[-] 0.5 0 100 200 300 400 500 40

30

1.5 C]

[ 2

1 T 20 C.V. [-] C.V. 0.5 0 100 200 300 400 500 10 Time [min] 0 100 200 300 400 500 Time [min]

(c) (d) 500 10 m] 5  300 0

[ml/min] 0 100 200 300 400 500

1 Size[ 100 F 0 100 200 300 400 500 10 5

1 0 [ml/min]

2 0 100 200 300 400 500 0.75 F

Yield[-] 40 C]

0.5 

0 100 200 300 400 500 [ 20 1 T 0 100 200 300 400 500 1.5 Time [min] 40

1 C] 

[ 20

C.V. [-] C.V. 2 0.5 T 0 100 200 300 400 500 0 100 200 300 400 500 Time [min] Time [min]

Figure 5.8. NMPC implementations: (a) controlled variables of temperature control; (b) manipulated variables of temperature control; (c) controlled variables of global control; (d) manipulated variables of global control (red dotted lines are set-points). C.V. is coefficient of variation of the CSD.

5.3.4.3 Regulatory Control

Disturbances are unavoidable in actual processes. Therefore, it is important to know how disturbances can influence the attainable regions of crystal size and yield. For example, when feed flow rate changes from 5 ml/min to 6 ml/min, it is found that the basic operating set-point (point 1) is no longer within the attainable region of nucleation, antisolvent and

102 temperature control methods, as shown in Figure 5.9. This is because the attainable regions shift due to this disturbance. As discussed in previous section, the necessary and sufficient condition for a certain control method and NMPC scheme to be feasible is that the set-point must be within the attainable region of that control approach. Therefore, in principal these three methods can no longer produce on-spec products if such a disturbance exists. Since the attainable region of global control method is the largest, and also it shifts the least, the global control method should be much more robust than the other four methods in terms of disturbance rejection.

Table 5.5. NMPC performance using the five proposed control methods.

NMPC is feasible (Y) or not (N)

Time [min] Set-point Control method

Antisolvent Temperature Nucleation Growth Global

0 - 50 1 Y Y Y Y Y

50 - 200 8 Y Y Y Y Y

200 - 350 1 Y Y Y Y Y

350 - 500 9 Y Y N N Y

Thus regulatory control was considered in this section to evaluate the disturbance rejection of the NMPC in the case of the global control method. Two scenarios were investigated for disturbance rejection, one with disturbance in the feed solution flow rate

F0 and another in the concentration c0 , respectively (Table 5.6). Step disturbances of magnitude between 10% and 20% were applied to the process after time 50 minutes. The

103

set-point was maintained at operating point 1 ( Ln 369 m , Y 0.685 ) all the time for both scenarios. The disturbance profile, controlled variables and manipulated variables for scenario 1 and 2 are presented in Figure 5.10.

1 Nucleation Control 0.9 Growth Control Antisolvent Control 0.8 Temperature Control Global Control 0.7 1

0.6

0.5 Yield[-] 0.4

0.3

0.2

0.1

0 100 200 300 400 500 600 700 800 Mean size [m]

Figure 5.9. Attainable region of crystal size and yield when feed flow rate increases from

5ml/min to 6ml/min. Operating point 1 is the basic operating set-point.

Table 5.6. Two scenarios for disturbance rejection evaluation. F0, c0 are feed solution flow rate and solute concentration.

Scenario 1 Scenario 2 Time [min] F0 [mL/min] c0 [g/mL] F0 [mL/min] c0 [g/mL] 0 - 50 5.0 0.40 50 - 200 6.0 0.44 0.40 5.0 200 - 350 5.0 0.40 350 - 500 4.0 0.36

104

(a) (b) 6 5 5

4 0

[ml/min]

[ml/min] 1

0 0 100 200 300 400 500 0 100 200 300 400 500

F F

m] 450 5  375

300 [ml/min] 0

2 Size[

0 100 200 300 400 500 F 0 100 200 300 400 500

0.75 35

C] 

0.675 [ 25

1 T Yield[-] 0.6 15 0 100 200 300 400 500 0 100 200 300 400 500 Time [min]

1.5 35

C] 

1 [ 25

2 T C.V. [-] C.V. 0.5 15 0 100 200 300 400 500 0 100 200 300 400 500 Time [min] Time [min]

(c) (d) 0.45 5 0.4

0.35 0

[g/ml]

[ml/min] 0

0 100 200 300 400 500 1 0 100 200 300 400 500

C F

m] 450 5  375

300 [ml/min] 0 Size[

0 100 200 300 400 500 2 0 100 200 300 400 500 F

0.75 35

C] 

0.675 [ 25

1 T Yield[-] 0.6 15 0 100 200 300 400 500 0 100 200 300 400 500 Time [min]

1.5 35

C] 

1 [ 25

2 T C.V. [-] C.V. 0.5 15 0 100 200 300 400 500 0 100 200 300 400 500 Time [min] Time [min]

Figure 5.10. Performance of the global control NMPC for disturbance rejection: (a) scenario 1, disturbance profile in feed flow rate (F0), crystal size, yield, and coefficient of variation (C.V.) of CSD (red dotted lines are set-points); (b) scenario 1, manipulated variables: F1, F2, T1, T2 are antisolvent addition rates and temperatures at stages 1 and 2;

(c) scenario 2, disturbance profile in feed concentration (c0), crystal size, yield, and coefficient of variation (C.V.) of the CSD (red dotted lines are set-points); (d) scenario 2 manipulated variables.

It is clear that the crystal size and yield can be well controlled at around the desired set- point, even though large disturbances from feed solution are present. The coefficient of

105 variation (or width of size distribution) can also be kept almost constant under the simulated large disturbance. Thus the proposed NMPC approach is robust in terms of disturbance rejection, if the set-point is still within the attainable region when disturbances are present.

5.4 Conclusions

A continuous two stage MSMPR cascade crystallizer was modeled in this work using population balance model, with both cooling and antisolvent addition applied to generate supersaturation. A MIMO control system was investigated, with crystal size and yield targeted as controlled variables, and antisolvent addition rates and temperatures at both stages used as manipulated variables. In addition, based on different selections of manipulated variables, five control methods (nucleation control, growth control, antisolvent control, temperature control, and global control) were discussed. The attainable regions of crystal size and yield of the proposed five control methods were obtained. The feasibility of two advanced control approaches, decentralized PID control and NMPC were evaluated. Decentralized PID control was proved to require change of control structure when a relatively large operating region is essential based on local linearization method and relative gain array (RGA) analysis, indicating that this process requires more advanced control approaches, such as NMPC, for smooth operation. It was found that the necessary condition for NMPC to be feasible is that the set-point must be within the corresponding attainable region of that particular control method. In general, when this condition is met,

NMPC scheme may be able to show good control performance for fast target product qualities (size) or process requirement (yield) change-over, as well as process disturbances

106 rejection. It is also shown that the non-square global control approach with four manipulated inputs (temperatures and flows in both stages) provides faster closed loop response and the largest attainable region. However, it should be noted that the practical performance of the NMPC highly depends on the accuracy of the model used in the controller. Another limitation of NMPC is related to sensor noise and error, which are not considered in this study either. But they can be quantified using experimental data in future works. These uncertainties would lead to variations in the boundaries of the attainable regions for control and could be incorporated in formulating the robust NMPC counterpart of the control framework presented in this work.

107

CHAPTER 6. INTEGRATED UPSTREAM AND DOWNSTREAM APPLICATION

OF WET MILLING WITH CONTINUOUS MIXED SUSPENSION MIXED

PRODUCT REMOVAL CRYSTALLIZATION10

6.1 Introduction

The application of wet mill in continuous crystallization processes is rarely reported in literature. In this work, it is used not only as a downstream unit operation for size reduction but also as an upstream continuous high shear nucleator for the first time. We proposed a novel concept of continuous wet mill seed generation (WMSG) whereby wet milling is used as an upstream unit operation before crystallization as a continuous seed generation process via high shear primary nucleation to decouple the mechanism of nucleation from growth controlled in the continuous crystallization unit. The integrated processes that use wet milling downstream as size reduction or upstream as seed generator are compared with the standard continuous MSMPR operation without wet milling using paracetamol

(acetaminophen) in ethanol solvent as the model compound in a cooling crystallization process. The effect of wet milling on the startup duration is also studied.

10 Reproduced with permission from Yang, Y., Song, L., Gao, T., Nagy, Z.K., 2015. Cryst. Growth Des. 15, 5879–5885. Copyright 2015 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/acs.cgd.5b01290

108

6.2 Experimental Section

In order to demonstrate the advantages of using wet mill on continuous MSMPRC and to prove the differences of applying wet mill upstream and downstream, three different methods and experimental setups are used in this work.

6.2.1 Method 1: MSMPRC without Wet Mill

Method 1 is the basic scenario in which only a single stage MSMPRC is used without wet mill. The experimental setup used in method 1 is presented in Figure 6.1(a). Both the feed vessel and the MSMPRC are 1 L round bottom jacketed glass reactors. Their jackets were connected to two temperature baths (Huber, USA). Each temperature bath contains a

Pt100 thermocouple that was used to measure the process temperature inside the feed vessel or MSMPRC. A FBRM S400 probe (Lasentec, USA) was installed in the MSMPRC to provide in situ particle chord length information, which was measured every 15 seconds.

The two temperature baths and FBRM were connected to a supervisory computer to record data and perform temperature control. The jacket temperatures of the feed vessel and

MSMPRC were kept constant at 55 oC and 25 oC, respectively. The agitation speeds were

350 rpm in both vessels. Paracetamol (Alfa Aesar, USA; purity > 98%) and anhydrous ethanol (KOPTEC, USA; purity > 99.5%) were used to prepare a clear 38 oC saturated feed solution (0.26 g/g solvent) (Mitchell and Frawley, 2010). A peristaltic pump (Masterflex,

USA) was installed to transfer clear solution from the feed vessel into the MSMPRC at a constant flow rate of 20 mL/min. The slurry volume inside the MSMPRC was maintained at 400 mL during operation, therefore the residence time of MSMPRC was 20 minutes. A transfer unit, which consists of a transfer zone and four valves, was used to transfer the

109 slurry out of the MSMPRC semi-continuously. This transfer unit works great for continuous crystallization process since it can avoid problems such as plugging of transfer tubing (Hou et al., 2014). In every transfer, the transfer zone was first evacuated by opening the vacuum valve for 1 second. Then the transfer zone inlet valve was opened for 1 second to draw in slurry from the MSMPRC. Finally, the transfer zone outlet valve and the nitrogen gas valve were opened simultaneously for 2 seconds to transfer out the slurry. The four valves and the transfer time interval were controlled by a control box automatically to maintain the slurry level constant in the MSMPRC. About 13.3 mL slurry was transferred every 40 seconds. The product slurry from the transfer unit was reused and dissolved for feed solution in order to save materials (Power et al., 2015). In addition, in every experiment an offline slurry sample was taken at the 11st residence time, and then filtered immediately for microscope image analysis (Kaiser Optical Systems, USA).

Since the clear startup solution in MSMPRC cannot spontaneously nucleate at 25 oC, one particular startup procedure performed in method 1 is dry seeding. For this purpose

0.10 g (0.1%) paracetamol dry powder (Figure 6.2) was added into the MSMPRC at the beginning of the experiment in order to induce nucleation. In methods 2 and 3 when wet milling is used, this dry seeding procedure is omitted because of the significantly enhanced nucleation rate under the high shear environment.

6.2.2 Method 2: MSMPRC without Wet Mill Operated Downstream in a Recycle Loop

The equipment setup for method 2 is shown in Figure 6.1(b). A magic LAB rotor-stator wet mill (IKA, USA) was installed in a recycle loop with the MSMPRC using a peristaltic pump. In this mill, a medium and a fine generator (i.e. rotor-stator pair) were used.

110

Figure 6.1. Experimental setup for (a) method 1 (without wet mill), (b) method 2 (with wet mill operated downstream) and (c) method 3 (with wet mill operated upstream). The parts shown by the solid dotted rectangles in each experimental setup are different among the three methods.

111

Wet mill tip speed is a commonly accepted scaling parameter, which is proportional to angular speed of the rotor:

vDts  r (6.1)

where vts is the rotor tip speed, Dr is the rotor outer diameter, and  is the rotor angular speed (Engstrom et al., 2013). In this work, two angular speeds, 6000 and 10000 rpm, were tested in order to investigate the influences of tip speed. In the mill recycle loop, two different flow rates (i.e. 50 and 200 mL/min) were used. Turnover number (TO) is one of the most important operating parameters, which defines the number of times the slurry is recycled through the wet mill in a batch process (Engstrom et al., 2013; Harter et al., 2013).

qt TO  mill tot (6.2) V

where V is the slurry operating volume, ttot is the total operation time (batch time), and

qmill is the recycle volumetric flow rate through wet mill.

We introduce a similar quantity for continuous crystallization coupled with wet milling, namely the turnover number per residence time (TOPRT), which defines how frequently the slurry in the MSMPRC is milled. The TOPRT (i.e. 2.5 and 10 in our studies) can be

calculated from the recycle volumetric flow rate through wet mill ( qmill ) and the volumetric

flow rate through MSMPRC ( qMSMPRC ) using the following equation:

qq TOPRT =mill = mill (6.3) VqMSMPRC where  is the MSMPRC residence time. The total operating volume in this method is also

400 mL, which consists of around 40 mL in the recycle loop and 360 mL in MSMPRC. All

112 the other operating conditions and procedures are exactly the same as in method 1. Since wet mill generates a lot of heat during operation, the wet mill jacket is cooled by cooling water to maintain the milling temperature at 25 oC.

6.2.3 Method 3: MSMPRC without Wet Mill Operated Upstream as Seed Generator

Figure 6.1(c) represents the equipment setup used in method 3. The same wet mill with same generator configuration was installed between the feeding peristaltic pump and the

MSMPRC. The clear feed solution was fed into the wet mill at a same flow rate (i.e. 20 mL/min). The wet mill was operated at 6000 and 10000 rpm, respectively. The other procedures and conditions remain the same as in methods 1 and 2.

Figure 6.2. Paracetamol powder seeds used in method 1.

6.3 Results and Discussion

The three methods discussed in section 2 are implemented in a serious of experiments, which are all monitored for around 12 residence times (RT) to capture both startup and steady-state dynamics. Three CQA of the final product (i.e. average particle size, particle

113 number, width of size distribution) and two important process qualities (i.e. yield and startup duration) are estimated from FBRM information at 11st RT. The wet mill operating conditions (i.e. angular speed, TOPRT) and corresponding process and product qualities of these experiments are summarized in Table 6.1. In upstream application, TOPRT always equals 1 since there is no recycle. This small TOPRT is suitable for upstream application since the purpose is to induce high shear primary nucleation mechanism to generate in situ seed crystals. However, in downstream application, large TOPRT and high recycling flow rate would be preferred since the purpose is to enable secondary nucleation and breakage to efficiently reduce particle size and normalize size distribution.

The MSMPRC process temperature and chord count information at different times are presented in Figure 6.3 and Figure 6.4. In general, the process temperatures in MSMPRC are well controlled and discrepancies are shown only during startup period. Since the heat generated by wet mill is very different in every experiment due to different experimental rigs and operating conditions (e.g. flow rates), there are some small differences in the final steady-state process temperatures in the MSMPRC among the different experiments. These differences are inevitable. However, their influences on nucleation rates and yield are negligible since nucleation is controlled in the wet mill, and the wet mill high shear effect on nucleation is much stronger than the supersaturation effects caused by small temperature differences in the MSMPRC (Allen et al., 2008; Gray et al., 1995; Reguera and Rubi, 2003).

The startup duration, as well as steady-state SWMCL and total chord counts can be easily identified in Figure 6.3 and Figure 6.4. The SWMCL is defined as:

3  fLii SWMCL  2 (6.4)  fLii

114

where Li represents the chord length of ith size interval, and fi is the frequency of chord length included in the i th interval (Leba et al., 2010). Figure 6.3 and Figure 6.4 clearly indicate the fast nucleation kinetics under high shear environment when wet mill is used.

In both methods 2 and 3, large amount of nuclei is formed at the beginning of the process and then washed out from the MSMPRC. Finally, the steady-state is reached when nucleation or/and breakage kinetics reach an equilibrium with the wash out speed from the

MSMPRC. The steady-state is identified based on both the total counts and mean chord length. Steady-state is assumed to be reached when both total counts and mean chord length do not show any fast increasing or decreasing trend.

Table 6.1. Summary of experimental conditions, process and product qualities using methods 1, 2 and 3.

Experiment No. 1 2 3 4 5 6

Method 1 2 3

Wet Speed [rpm] -- 10000 10000 6000 10000 6000 mill TOPRT [-] -- 10 2.5 10 1

Startup duration [RT] 6 6 >11 9 5 3.5

SWMCL [μm] 71 55 66 64 57 77

Total counts [#/s] 3700 10300 7300 4100 15100 10100

CV [-] 1.06 0.97 0.96 1.07 0.92 1.02

Yield index (YI) [-] 1.3e9 1.7e9 2.1e9 1.1e9 2.8e9 4.6e9

115

34 120 (a) Process Temperature 20000 Total Counts 100

32 Length Chord Mean Mean Chord Length, Sqr Wt 16000

] 80 C

30 [#/s] Counts

o [ 12000 Startup Steady-state 60 28 8000 40 26

[ Temperature

 4000 20 m

] 24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 No. of Residence Time [-]

34 20000 120 (b) Process Temperature 100 32 Total Counts 16000

Mean Chord Length Length Chord Mean Mean Chord Length, Sqr Wt

] 80 C

30 [#/s] Counts o

[ 12000 60 28 8000 40

26 Startup Steady-state Temperature

[

4000  20 m 24 ] 0 0 0 1 2 3 4 5 6 7 8 9 10 11 No. of Residence Time [-]

34 20000 120 (c) Process Temperature 100 32 Total Counts 16000 Length Chord Mean Mean Chord Length, Sqr Wt

] 80 C 30 [#/s] Counts o Steady-state is no reached within 11 residence time [ 12000 60 28 8000 40 26

[ Temperature

4000 m 20

] 24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 No. of Residence Time [-]

Figure 6.3. Process temperature, FBRM total counts and SWMCL in the MSMPRC in (a) experiment 1, (b) experiment 2, (c) experiment 3. Fluctuations in the temperature profile in (a) are caused by disturbances from the wet mill cooling water, but their influences on total counts and mean chord length are negligible.

116

34 20000 120 (a) Process Temperature 100 32 Total Counts 16000 Length Chord Mean Mean Chord Length, Sqr Wt

] 80 C

30 [#/s] Counts o

[ 12000 60 28 Startup Steady-state 8000 40 26

[ Temperature

4000 m 20

] 24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 No. of Residence Time [-]

34 120 (b) Process Temperature 20000 Total Counts 100

32 Length Chord Mean Mean Chord Length, Sqr Wt 16000

] 80 C

30 [#/s] Counts

o [ Startup Steady-state 12000 60 28 8000 40 26

[ Temperature

 4000 20 m

] 24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 No. of Residence Time [-]

34 120 (c) Process Temperature 20000 Total Counts 100

32 Length Chord Mean Mean Chord Length, Sqr Wt 16000

] 80 C

30 [#/s] Counts

o [ 12000 60 28 8000 40 26 Startup Steady-state

[ Temperature

 4000 20 m

] 24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 No. of Residence Time [-]

Figure 6.4. Process temperature, FBRM total counts and SWMCL in the MSMPRC in (a) experiment 4, (b) experiment 5, (c) experiment 6.

In terms of startup duration, Table 6.1, Figure 6.3 and Figure 6.4 indicate that the use of wet milling applied downstream in a recycle loop may potentially increase startup

117 duration, especially when wet mill tip speed or TOPRT is low. However, the startup duration can be significantly reduced if wet milling is used upstream as a high shear nucleator. This is because when used downstream, at the beginning of the process the supersaturation is high and particle number is small, therefore the major kinetics inside the wet mill is high shear heterogeneous primary nucleation. But as the supersaturation drops down and number of particles increases in the MSMPRC, the major kinetics inside the wet mill becomes high shear secondary nucleation and breakage. This complex kinetics change may not be favorable to startup and can result in longer duration until steady-state is achieved. On the other hand, when wet milling is used upstream, the major kinetic is high shear heterogeneous primary nucleation only. Secondary nucleation or breakage is negligible in this case, since the feed solution is clear and of high supersaturation. The high and relatively constant primary nucleation rate achieved inside the wet mill can help the system to deplete the supersaturation fast and reach steady-state quickly. Therefore the upstream application of wet milling is better than using it downstream considering startup duration.

The SWMCL shown in Table 6.1, Figure 6.3 and Figure 6.4 can be used to infer the average particle size achieved. The results indicate that the use of wet milling downstream always reduces particle size, especially when the wet mill tip speed and TOPRT are high.

This is because high shear secondary nucleation and breakage are dominant in the system, and growth is completely suppressed in this case. However, when wet mill is used upstream, small crystals are obtained when the tip speed is high, whereas large crystals can be produced when the tip speed is low. The reason is that high tip speed of wet mill results in high primary nucleation rate and large amount of small seed crystals, and therefore only

118 small amount of supersaturation is available for crystals to grow in the MSMPRC. On the contrary, low wet mill tip speed ends up with limited amount of seed crystals and relatively high supersaturation in MSMPRC for crystal growth therefore average crystal size is large.

In addition, in experiment 6 nucleation and growth kinetics are very well decoupled, which is also beneficial for achieving large average crystal size.

The total counts presented in Table 6.1, Figure 6.3 and Figure 6.4 indicate the nucleation rate and particle number in each experiment. The application of wet milling always increases nucleation rate and particle number, especially when it is used upstream and with high tip speed.

The previous conclusions related to average particle size and particle number can also be clearly identified from the square weighted chord length distributions (SWCLD), unweighted chord length distributions and relative frequency distributions (Figure 6.5) measured at steady-state (11st RT). In addition, width of the chord length distribution can be inferred by calculating the CV from the unweighted chord length distribution (Figure

6.5(b)) using the following equations:

nn  i (6.5)

nL L   ii (6.6) n

n() L L 2 SD   ii (6.7) n 1

SD CV  (6.8) L

119

(a) 14 Experiment 1 square weight Experiment 2 12 Experiment 3 Experiment 4 10 Experiment 5 Experiment 6 8

6 Counts [#/s] Counts 4

2

0

1 10 100 1000 Chord Length [m]

(b) 800 Experiment 1 700 Experiment 2 no weight Experiment 3 600 Experiment 4 Experiment 5 Experiment 6 500

400

Counts [#/s] Counts 300

200

100

0 1 10 100 Chord Length [m]

(c) 0.040 Experiment 1 0.035 Experiment 2 no weight Experiment 3 0.030 Experiment 4 Experiment 5 Experiment 6 0.025

0.020

Frequency[-] 0.015

0.010

0.005

0.000 1 10 100 Chord Length [m]

Figure 6.5. (a) SWCLDs, (b) unweighted chord length distributions and (c) unweighted relative frequency distributions obtained from FBRM at steady-state (11st RT).

120

where ni represents the chord count within chord length Li , L is the average chord length, and SD is the standard deviation. The CV summarized in Table 6.1 and the chord length and frequency distributions shown in Figure 6.5 prove that applying wet milling either upstream or downstream can help narrow down size distribution, especially when high tip speed is used. The improved crystal uniformity in downstream wet milling is because of the breakage mechanism, whereas in upstream wet milling the improved uniformity is due to the uniform seed crystals generated by the wet mill as well as the decoupled nucleation and growth mechanisms. The crystal images presented in Figure 6.6 confirm the conclusions discussed before in terms of average particle size and width of size distribution.

Yield increase is another potential advantage of using wet milling in conjunction with continuous crystallization due to the significantly enhanced nucleation rate. Since yield is a function of both particle number and particle size, a simple yield index (YI) is used to estimate the yield from the solid phase FBRM data:

Yield index (YI) = Square weighted mean chord length3  Total counts (6.9)

The yield index (YI) is a very simple and convenient measure that combines both particle size and particle number information which are inferred from chord length and chord counts. Due to the limitations of FBRM (e.g. chord length is measured instead of particle size), the approach stated above is not accurate for yield calculation. However, it is suitable in terms of qualitative comparative analysis of the proposed approaches. The results are summarized in Table 6.1 for comparison. The YI indicates that the application of rotor-stator wet milling in continuous crystallization system can improve yield under a fixed residence time due to the enhanced nucleation kinetics. This is because in continuous

121 crystallization, the process is operated at steady-state rather than equilibrium state. In our experiments, the residence time is only 20 minutes, which means that the steady-state operating point in experiment 1 is far from the solubility curve in the phase diagram and yield is governed by kinetics rather than thermodynamics.

Figure 6.6. Microscope images for crystals obtained at steady-state (11st RT) in (a) experiment 1, (b) experiment 2, (c) experiment 3, (d) experiment 4, (e) experiment 5 and

(f) experiment 6.

122

Therefore, the wet milling effect on yield improvement is significant and the yield obtained in experiment 1 is much smaller than in all the other experiments. These results actually also show an additional advantage of the wet mill by its potential to decrease equipment size, when residence time is determined by yield rather than required size, since higher yield can be achieved by lower residence time due to the increased nucleation rate achieved by the wet mill. The smaller crystallizer size for a desired yield can counteract some of the increased capital costs related to using the wet mill as an additional equipment in the process. This yield improvement is remarkable especially when wet mill is used upstream. Because in upstream application, the wet mill works at high supersaturation and induces high shear heterogeneous primary nucleation, therefore the nucleation rate enhancement is more significant than using it downstream in which secondary nucleation and breakage are dominating.

Although it is true that there are a few foreseeable downsides of integrating wet milling with continuous crystallization, such as increased capital cost, complex process design and potential risk for temperature sensitive APIs, the results in this work clearly demonstrate the benefits of the integrated continuous wet milling-crystallization (CWMC) process to achieve better control of the CSD and yield and also indicate the benefits of the novel configuration of upstream wet mill as a continuous high shear nucleator. Upstream wet mill provides an excellent approach to control primary nucleation kinetic, which is typically difficult to control due to its stochastic nature. In this work the integration of wet milling with MSMPRC is demonstrated but the results and conclusions are also valid for combined use of wet milling and continuous plug flow crystallization systems. For example,

123 wet milling could be used as a continuous seed generation unit upstream of a plug flow crystallization system.

6.4 Conclusions

A high shear rotor-stator wet mill is applied upstream and downstream of a continuous

MSMPRC. The wet mill becomes a continuous seed generator when it is used upstream because of the high shear heterogeneous primary nucleation kinetic that characterizes the system. In addition, the wet mill can be used as a continuous size reduction tool when applied downstream from MSMPRC due to the controlled secondary nucleation and breakage mechanisms. It is found that the downstream application of wet milling may increase slightly startup duration. However, startup duration can be significantly reduced if wet milling is used upstream as a continuous nucleator. In addition, using wet mill downstream can efficiently reduce average particle size and narrow size distribution. On the other hand, not only small and uniform crystals, but also large and uniform crystals can be achieved by applying wet milling upstream and properly controlling the tip speed.

Finally, wet milling used in conjunction with continuous crystallization was found to significantly enhance yield especially when used upstream as a high shear nucleator due to the improved nucleation kinetics. Therefore, in general applying rotor-stator wet milling upstream as a continuous in situ seed generator is very promising to achieve an integrated and intensified process of high yield and quick startup, and good product qualities (i.e. desired particle size and narrow size distribution) in continuous MSMPR or plug flow crystallization systems.

124

CHAPTER 7. AUTOMATED DIRECT NUCLEATION CONTROL IN CONTINUOUS

MIXED SUSPENSION MIXED PRODUCT REMOVAL COOLING

CRYSTALLIZATION11

7.1 Introduction

The works that have been done so far in literature in continuous MSMPRC are either without process control or with open-loop control only. In addition, the vast majority of works described in literature that focus on optimization and control aspects related to

MSMPRC are theoretical simulation studies only. Therefore, the experimental implementation of some practical control methodology is necessary in the development of continuous crystallization, to bring theoretical advances in the realm of possibilities.

This chapter is the first that implements a closed-loop (feedback) control in a continuous crystallization process in practice using PAT tool. In this work, a FBRM-based feedback control approach, which is modified from batch ADNC (Abu Bakar et al., 2009; Saleemi et al., 2012a, 2012b) to accommodate the requirement for steady-state operation in continuous crystallization, is proposed and implemented in single-stage and two-stage continuous MSMPRC. An API, paracetamol, is used as the model compound in ethanol solvent in a continuous cooling crystallization process.

11 Reproduced with permission from Yang, Y., Song, L., Nagy, Z.K., 2015. Cryst. Growth Des. 15, 5839– 5848. Copyright 2015 American Chemical Society. http://pubs.acs.org/doi/10.1021/acs.cgd.5b01219

125

7.2 Experimental Section

The experiments were carried out in a continuous two-stage MSMPRC (Figure 7.1).

Three 1 liter jacketed round bottom crystallizers were used as feed vessel, first stage

MSMPRC and second stage MSMPRC, respectively. The jackets of the three vessels were connected to three temperature circulators (Ministat 125 with Pilot ONE, Huber, USA), in which jacket control (internal control) model was selected. Each MSMPRC stage was equipped with a focused beam reflectance measurement (FBRM) probe (S400 and G400,

Mettler-Toledo, USA) and a Pt100 thermocouple. The particle chord counts and process temperature in each stage were measured every 15 seconds and transferred into a Labview based software Crystallization Monitoring and Control (CryMOCO) or Crystallization

Process Informatics System (CryPRINS), which was installed on a supervisory computer.

This real-time data was then used in the ADNC approach to automatically modify temperature set-points in the temperature circulators through CryMOCO. The data were communicated via RS232 interface.

A 38 oC saturated paracetamol (Purity > 98%; Alfa Aesar, USA) - ethanol (Purity >

99%; KOPTEC, USA) solution (0.26 g/g ethanol) was used as feed solution (Mitchell and

Frawley, 2010). This clear feed solution was transferred from the feed vessel into the first stage MSMPRC through a peristaltic pump (L/S digital pump, Masterflex, USA). The slurry obtained in the first stage was then transferred into the second stage semi- continuously via a transfer unit, which consists of a transfer zone, an inlet valve, an outlet valve, a nitrogen valve and a vacuum valve. The slurry could be firstly drew in from the previous stage into transfer zone by vacuum, and then transferred into the next stage via high pressure nitrogen gas. Another transfer unit was used to transfer the product slurry

126 from the second stage MSMPRC back into the feed vessel for recycle in order to save materials (Power et al., 2015). The solution in the feed vessel was kept at around 1 liter and 55 oC. The recycled crystals were observed to dissolve immediately once entered into the feed vessel. In addition, the feed solution was used as the initial solution in both

MSMPRC vessels. Therefore, theoretically the recycled slurry should have same concentration as the feed after all solids dissolve, and the influence of recycle on feed concentration was negligible.

Figure 7.1. Experimental setup of continuous two-stage MSMPRC.

The transfer intervals of the two transfer units were controlled in a controller box to maintain the slurry volumes constant in both stages (i.e. 300 mL). At the beginning of every experiment, 0.10 g paracetamol fine powders (1100 counts/sec) were introduced into the first stage to trigger secondary nucleation. ADNC was switched on at around 0.25 RT, right

127 after seed crystals were introduced. The purpose is to ensure that the ADNC approach controls secondary nucleation rather than primary nucleation, since secondary nucleation is generally dominant in MSMPRC systems.

7.3 Continuous ADNC Methodology

In the proposed continuous ADNC approach, each stage is controlled individually by an ADNC loop. For a certain MSMPRC stage, an upper bound and a lower bound of particle total chord counts set-point are defined by the user in the ADNC setup in

CryMOCO. During the continuous crystallization, once the particle total chord counts measured by the FBRM in that MSMPRC stage is found higher than the upper bound, a fast heating will be actuated automatically by the ADNC to reduce nucleation rate or trigger controlled dissolution (Figure 7.2) in that stage. The fast heating will change to slow heating once the measured total counts drops below upper bound. Similarly, fast cooling or slow cooling is implemented automatically by ADNC if the measured total chord counts is below lower bound or increases above lower bound. The CryMOCO user interface for

ADNC implementation is presented in Figure 7.3, from which the total chord counts set- point upper bound, lower bound, fast heating rate and fast cooling rate can be specified.

Slow heating and slow cooling rates are defined from heating gain ( Kh ) and cooling gain

( Kc ) (Figure 7.3) based on the following equations:

Fast heating rate Slow heating rate = (7.1) Kh

Fast cooling rate Slow cooling rate = (7.2) Kc

128

Since steady-state operation is desired in continuous crystallization, extremely large heating and cooling gains (i.e. 255) are used in this work so that slow heating and slow cooling rates become practically 0. In other words, the temperature set-point will remain constant once the measured total chord counts fall between the upper and the lower bounds.

In addition, maximum and minimum temperature set-points allowed can also be defined in

ADNC in CryMOCO (Figure 7.3).

Figure 7.2. ADNC set-point and heating/cooling algorithm.

Figure 7.3. CryMOCO user interface for ADNC implementation.

In this chapter, a series of experiments were carried out in single-stage and two-stage

MSMPRC, with and without using ADNC approach. In the two-stage MSMPRC the use of ADNC in both stages were also investigated and compared to when ADNC is only used

129 in the second stage. The detailed operating conditions and ADNC parameters are summarized in Table 7.1.

Table 7.1. Operating conditions and ADNC parameters used in every experiment.

No. No. of Residence time ADNC is used in ADNC parameters stages in each stage which stage Fast heating/cooling Minimum/maximum [min] rates [oC/min] temperatures [oC] 1 Single- 20 No ADNC -- -- 2 stage 20 First ±0.1 3 – 50 3 Two- 20 No ADNC -- -- 4 stage 20 Second ±0.1 3 – 50 5 75 First and second ±0.05 15 – 26 6 20 Second ±0.1 3 – 50

7.4 Results and Discussion

7.4.1 Startup and Control of CSD

7.4.1.1 Continuous ADNC in Single-Stage MSMPRC

Experiments 1 and 2 are from single-stage MSMPRC. In experiment 1, ADNC is turned off and the jacket temperature of MSMPRC is kept constant at 25 oC. In experiment 2, the initial jacket temperature of MSMPRC is also 25 oC, but ADNC is switched on immediately after seed crystals are introduced. The process temperature, ADNC jacket set- point, FBRM total chord counts, ADNC total counts set-point and SWMCL of these two experiments are compared in Figure 7.4. It is observed that the total counts increases slowly and 6 residence time (RT) is required to reach steady-state in experiment 1 when ADNC is not used. However, in experiment 2 when ADNC is turned on, the temperature of

130

MSMPRC decreases automatically at the beginning of the startup to force the system to nucleate faster and reach the desired total counts set-point sooner. As a result, only 4 RT is spent to reach steady-state in experiment 2 (Figure 7.4(b)). Therefore, ADNC is able to reduce significantly the startup duration in a single-stage MSMPRC. It should be noted that the steady-state is identified qualitatively based on both the total counts and mean chord length. For experiments without ADNC, steady-state is considered reached when both total counts and mean chord length do not show any clear increasing or decreasing trend. For the experiments with ADNC, steady-state is considered reached when the total counts is kept closely between the maximum and minimum values of the total counts set-points, and mean chord length does not show any clear increasing or decreasing trend. The steady-state criteria cannot be mathematically quantified here due to the nature of FBRM measurement, which shows high sensitivity to noise or disturbance, such as fouling and particles adhered on probe window, as well as variations in the mixing conditions. In our setup, the slurry is transferred using high pressure nitrogen gas, which can cause short spikes and variations in the FBRM measurements even under practically steady-state operation.

In addition, in experiment 2 the ADNC set-point is modified in CryMOCO after the system reaches steady-state 1. The purpose is to characterize the ADNC performance when a set-point change is required, as well as to compare the crystal chord length distribution

(CLD) obtained at steady-states at different set-points. It is found that only 0.7 RT is required for ADNC to move the system from steady-state 1 into steady-state 2 (Figure

7.4(b)). The CLD obtained in both steady-states are compared in Figure 7.5. The results clearly indicate that when ADNC is used, larger ADNC total counts set-point results in smaller average crystal size and more fine particles. This is probably because paracetamol

131 is a nucleation dominated compound. This phenomenon agrees with simulation study in literature which shows that there is a tradeoff between average crystal size and yield within fixed MSMPRC setup and operating constraints (Yang and Nagy, 2015a, 2015b).

(a) 30 14000 140 Process Temperature Startup (6 RT) Steady-state Jacket Set-point 12000 120 Total Counts 10000 100 [ Length Chord Mean

28 SWMCL C]

Counts [#/s] Counts o 8000 80

6000 60 26

4000 40 Temperature[

2000 20 

m]

24 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 No. of Residence Time ( ) [-] 

(b) 30 Process Temperature 14000 140 Jacket Set-point 28 Total Counts 12000 120 ADNC Set-point 26 SWMCL 10000 100 [ Length Chord Mean

Startup Steady-state 1 C]

Counts [#/s] Counts o 24 (4 RT) 8000 80 22 6000 60 20 0.7 4000 40

Temperature[ Steady 18 RT -state 2 2000 20  16 m] 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 No. of Residence Time ( ) [-] 

Figure 7.4. Process temperature, ADNC jacket set-point, FBRM total chord counts, ADNC total counts set-point and SWMCL in single-stage MSMPRC: (a) without ADNC

(experiment 1), (b) with ADNC (experiment 2).

This observation also provides an approach to control product CSD according to formulation requirements by manipulating the ADNC set-point. It is considered that once the experimental setup, feed concentration and residence times are fixed, the average particle size is a function of particle number only. In order to reach the target average

132 particle size, the first step is to find out a target particle number, which corresponds to the target average particle size. This would require a few preliminary experiments to determine a correlation or calibration relationship, which then can be used to obtain the target particle count corresponding to a desired size. Once the target particle number is found out, the target average particle size can be achieved by controlling the particle number using the

ADNC approach. In general, for paracetamol system it is observed that a relatively small

ADNC set-point should be used to avoid strong nucleation when fine crystals are not desired, and vice versa. The crystal images (Figure 7.6) confirm that more fine particles are generated in steady-state 2 than steady-state 1.

7 Steady-state 1 Steady-state 2 6

5

4

3

Counts[#/s] 2

1

0

100 Chord Length [m]

Figure 7.5. Crystal chord length distribution (CLD) at steady-state 1 and steady-state 2 in experiment 2 when ADNC is used in single-stage MSMPRC.

In addition, theoretically as long as the final operating temperatures or particle numbers are fixed, whether the continuous ADNC approach is used or not during startup, or how the steady-state is reached will not have any effect on the final steady-state CSD or average

133 crystal size, since all the crystals are continuously washed out, and once the ADNC set- point stabilized the steady-state CSD will be reached. However, under certain conditions, in practice it might be observed that using the ADNC during startup leads to larger average size than without ADNC (Figure 7.6) when steady-state is reached. This is because the

ADNC thermo cycles can dissolve fine particles and produce some very large particles

(Abu Bakar et al., 2009; Saleemi et al., 2012a, 2012b), which are very difficult to be washed out.

Figure 7.6. Crystal microscope images of (a) experiment 1, (b) experiment 2 steady-state

1, and (c) experiment 2 steady-state 2.

7.4.1.2 Continuous ADNC in Two-Stage MSMPRC

Two-stage MSMPRC is used in experiment 3, 4 and 5. Experiment 3 is the base case in which ADNC is not used and the jacket set-points of the first and the second MSMPRC are kept constant at 25 oC and 19 oC, respectively. Since the advantages of ADNC on the first stage has already been observed in section 7.4.1.1, the effects of applying ADNC on the second stage is investigated in experiment 4, in which the jacket set-point of the first stage is kept constant at 25 oC and the initial jacket set-point of the second stage is set as

134

19 oC before turning on ADNC. The second stage of these two experiments are compared in Figure 7.7.

(a) 30 14000 180 Process Temperature T2 160 28 Startup (5 RT) Steady-state Jacket Set-point T2 12000 Total Counts 140 [ Length Chord Mean 26 10000

SWMCL 120 C]

Counts [#/s] Counts o 24 8000 100

22 6000 80 60 20 4000

Temperature[ 40

 18 2000 m] 20 16 0 0 0 1 2 3 4 5 6 7 8 9 10 11 12 No. of Residence Time (  ) [-]  

Process Temperature T2 (b) Jacket Set-point T2 30 6000 500 Total Counts 28 Startup ADNC Set-point 5000 (2.5 RT) Steady-state 1 SWMCL 400 26 [ Length Chord Mean

2.3 RT 4000 [#/s] Counts C] 24 Steady-state 2 o 300 22 3000 20 200 2000

18 Temperature[

100  16 1000 m] 14 0 0 0 1 2 3 4 5 6 7 8 9 10 No. of Residence Time ( + ) [-]  

Figure 7.7. Process temperature, ADNC jacket set-point, FBRM total chord counts, ADNC total counts set-point and SWMCL in the second stage: (a) without ADNC (experiment 3),

(b) with ADNC implemented in the second stage (experiment 4). The variation in total counts between 3.5 and 4 RT is due to probe fouling and cleaning.

It is observed that the startup duration can be reduced from 5 RT to only 2.5 RT when

ADNC is used in the second stage. In addition, sometimes sharp jumps of total counts or mean chord length may be noticed, such as total counts between 3.5 RT to 4 RT in Figure

7.7(b). This part of data is actually a false signal caused by fouling and crystals adhered to

135 the window of the FBRM probe. The total counts signal returns to its original value immediately after the probe window is cleaned at 4 RT. Since the ADNC is based on the

FBRM total counts measurement only, the false signal of mean chord length will not influence the control performance, however the false signal of total counts can disrupt the control algorithm. Therefore, it is very important to ensure that the FBRM window is clean during operation. The window must be cleaned as soon as abnormal data jump is observed in the signal.

A set-point change in ADNC is also implemented in experiment 4 in the second stage

(Figure 7.7(b)). Around 2.3 RT is used for the system to move from steady-state 1 to steady-state 2. The CLD obtained from these two steady-states are compared in Figure 7.8, which indicates that in the ADNC approach, smaller ADNC set-point results in larger average crystal size and less fine particles for paracetamol.

Steady-state 1 Steady-state 2 2.5

2.0

1.5

1.0 Counts [#/s] Counts

0.5

0.0

10 100 Chord Length [m]

Figure 7.8. CLD at steady-state 1 and steady-state 2 in experiment 4 when ADNC is used in the second stage.

136

Therefore, similarly as in the case of the single-stage system, the CSD can be controlled in two-stage MSMPRC by manipulating the ADNC set-point in the second stage. Crystal images in Figure 7.9 also show that average crystal size is increased from steady-state 1 to steady-state 2 due to a decreased ADNC set-point.

Figure 7.9. Crystal microscope images of (a) experiment 3, (b) experiment 4 steady-state

1, and (c) experiment 4 steady-state 2.

In previous experiments, the RT in each stage is relatively short (i.e. 20 minutes), and

ADNC is used only in one of the two stages. Therefore, experiment 5 is performed in order to test the feasibility of applying ADNC in systems that have long residence time, as well as the possibility of using ADNC simultaneously in more than one stage. The RT of each stage in experiment 5 is 75 minutes. Two ADNC loops are used in the first and the second stage simultaneously. The FBRM and temperature data of both stages are presented in

Figure 7.10. The startup durations spent in the first and the second stage are only 2.4 RT and 3 RT, respectively. This observation once again shows the ability of the ADNC approach to provide quick startup. Two steady-state operating conditions are evaluated in experiment 5 by changing the ADNC set-point in the second stage. In this case it takes only about 0.3 RT for the system to go from steady-state 1 to steady-state 2 (Figure 7.10(b)).

137

The ability of the ADNC approach to control the CSD is also observed here based on the comparisons of CLD (Figure 7.11) and microscope images (Figure 7.12) in the two steady- states. The results clearly indicate that large and uniform crystals (Figure 7.12(a)) can be obtained by the ADNC if the set-point is properly chosen, whereas fine particles may be generated (Figure 7.12(b)) if the set-point is selected to be at larger counts/sec.

(a) Process Temperature T1 Jacket Set-point T1 6000 600 34 Total Counts FBRM1 32 ADNC Set-point 5000 500

SWMCL [ Length Chord Mean 30 Startup (2.4 RT) Steady-state

4000 400 C]

28 [#/s] Counts o 26 3000 300 24 22 2000 200

Temperature[ 20

1000 100  18 m] 16 0 0 0 1 2 3 4 No. of Residence Time ( + ) [-]  2

(b) Process Temperature T2 30 Jacket Set-point T2 4000 600 Total Counts FBRM2 ADNC Set-point 3500 500 28 SWMCL

3000 [ Length Chord Mean

Startup (3 RT) 0.3 400 C] Steady-state 1 Steady-state 2 [#/s] Counts o 26 RT 2500 2000 300

24 1500 200

Temperature[ 1000 22 100  500 m]

20 0 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 No. of Residence Time ( + ) [-]  2

Figure 7.10. Process temperature, ADNC jacket set-point, FBRM total chord counts,

ADNC total counts set-point and SWMCL in (a) first stage with ADNC (experiment 5), (b) second stage with ADNC (experiment 5).

The experiments previously discussed indicate that the length of residence time will not limit the performance of ADNC. In addition, ADNC can be used in any one of the stages

138 or in multiple stages simultaneously. Although only two-stage MSMPRC is implemented here, in principle there should be no problem to apply ANDC in MSMPRC of more than two stages.

8 Steady-state 1 Steady-state 2 7 6 5 4

3 Counts[#/s] 2 1 0 100 Chord Length [m]

Figure 7.11. CLD at steady-state 1 and steady-state 2 in experiment 5 when ADNC is used in both stages.

Figure 7.12. Crystal microscope images of (a) experiment 5 steady-state 1, (b) experiment

5 steady-state 2.

Additionally, this approach allows to control nucleation quantitatively in every stage if

ADNC is applied in multiple stages and the FBRM data are calibrated suitably. For

139 example, nucleation and the generation of fine particles in the second stage might be completely avoided by selecting a same ADNC set-point in both the first and the second stages. More generally, in a multi-stage MSMPRC with the ADNC applied and FBRM data calibrated in all stages, for a given starting solution concentration or saturation temperature, if large particle size is desired, same or similar maximum/minimum total counts set-points should be used in all stages so that nucleation after the first stage can be minimized or completely avoided. The values of these total counts set-points can be then tuned based on the yield requirement. On the other hand, if small particle size and high yield are required, the maximum/minimum total counts set-points of the second or subsequent stages should be as larger in order to maximize nucleation and minimize growth.

7.4.2 Disturbance Suppression

Since ADNC is a feedback control approach, in principle it should be able to help suppress disturbances. Experiment 6 is a follow up of experiment 3 after steady-state is reached. As shown in Figure 7.13, disturbance 1 and disturbance 2 are the same artificial pulse disturbances of adding 0.5 g paracetamol fine powders into the second stage to disrupt the steady-state. This could be for example a practical disturbance that may occur from accidental seeding from an encrust layer that may form on the walls or stirrer shaft during the continuous crystallization operation. ADNC is turned off when disturbance 1 is introduced but turned on before introducing disturbance 2. It is found that 1.7 RT is required to wash out the fine particles without ADNC. However, when ADNC is used, the

MSMPRC can automatically heat up to dissolve the fine particles therefore they are removed much quicker. Only 0.9 RT is needed to completely suppress the same disturbance

140 by using ADNC. The CLD in Figure 7.14 shows that ADNC can provide exactly same good performance as the simple washing out approach to adjust the CLD back into the original steady-state. But less time is required and less products are wasted if ADNC is used.

Disturbance 1 Disturbance 2 Process Temperature T2 30 Jacket Set-point T2 180 Total Counts 8000 160 28 ADNC Set-point 1.7 0.9 7500

SWMCL [ Length Chord Mean Steady-state 1 7000 140 26 RT RT Counts [#/s] Counts 120

C] Steady-state 2 6500 o 24 6000 100 22 5500 80 5000 60 20

Temperature[ 4500 40

 18 4000 m] 20 Without ADNC With ADNC 3500 16 0 0 1 2 3 4 5 6 7 No. of Residence Time ( ) [-] 

Figure 7.13. Process temperature, ADNC jacket set-point, FBRM total chord counts,

ADNC total counts set-point and SWMCL in experiment 6 with disturbances introduced artificially.

Initial steady-state (Before disturbance 1) 9 At Disturbance 1 At Steady-state 1 8 At Disturbance 2 7 At Steady-state 2 6 5 4

3 Counts[#/s] 2 1 0 100 Chord Length [m]

Figure 7.14. CLD at different times in experiment 6 with disturbances introduced artificially.

141

In addition, ADNC should have advantages on suppressing not only pulse disturbance but also continuous disturbance (e.g. step disturbance). For example, if a feed solution that has higher concentration than nominal concentration is used, ADNC should be able to detect the increased nucleation rate therefore increase the MSMPRC temperature accordingly to maintain the total particle chord counts within the ADNC set-point and as a result maintain the average crystal size. However, system without ADNC cannot react to the feed concentration increase. It will end up with smaller crystal size than the desired.

The operating temperature has to be re-designed manually in order to work for a different feed concentration without ADNC. Therefore, ADNC can significantly improve the robustness of continuous MSMPRC. Additionally, this study shows that the proposed closed-loop ADNC approach is stable and total counts can converge to the desired set-point under selected control parameters. However, it is observed that this closed-loop approach becomes unstable when too large heating and cooling rates are used. There are two potential causes for instability under such aggressive parameters. First of all, the actual process temperature will have significant delay compared to the calculated temperature set- point. Additionally, the temperature change will be too fast compared to the crystallization kinetics for the process to be able to respond. Therefore a few preliminary experiments are required to tune the control parameters when a new experimental setup or compound is introduced for ADNC application.

Finally, as already described in chapter 2, SSC is another widely used feedback control approach that uses concentration measurement to adjust temperature set-point in batch cooling crystallization. Since in this work ADNC is observed to have superior performance in continuous crystallization, we anticipate that SSC can be used in a similar way to reduce

142 startup duration, control CSD and suppress disturbance in continuous MSMPRC. Low cost

PAT imaging tools can also be used in the process as a supplement for FBRM or UV/Vis spectroscopy to monitor and control CSD (Simon et al., 2012).

7.5 Conclusions

In this chapter, a FBRM-based feedback control approach, named as ADNC, is implemented in single-stage and two-stage MSMPRC. Based on the in situ FBRM information, ADNC can automatically actuate heating or cooling to control the total particle chord counts within the ADNC set-point range during both startup and steady-state operations. It is found that the proposed continuous ADNC approach can significantly reduce the startup duration. In addition, the steady-state CSD can be controlled by adjusting the ADNC set-point. Large crystals can be obtained if the ADNC set-point is properly selected to avoid nucleation in the second stage. Finally, ADNC is observed to provide very quick and effective disturbance suppression in continuous MSMPRC.

143

CHAPTER 8. APPLICATION OF WET MILLING-BASED AUTOMATED DIRECT

NUCLEATION CONTROL IN CONTINUOUS COOLING CRYSTALLIZATION

PROCESSES12

8.1 Introduction

Chapter 6 reported that very uniform crystals can be obtained by integrating a wet mill unit with continuous MSMPRC (Yang et al., 2015d). The advantages of integrating wet milling (WM) into continuous crystallization, such as improved yield and narrowed size distribution, have been observed. However, there was no process control framework implemented in the study of this WM technique. In addition, Chapter 7 introduced a FBRM based feedback control approach for cascade MSMPRC systems by extending the ADNC approach developed originally for batch crystallization processes (Yang et al., 2015e). In this ADNC approach the crystal size is controlled by manipulating the MSMPRC jacket temperature and nucleation kinetics. However, one potential limitation of this ADNC technique is that the nucleation in MSMPRC cannot be separated from growth therefore this nucleation control approach will always be also influenced by the growth kinetics in the MSMPRC.

12 Reproduced with permission from Yang, Y., Song, L., Zhang, Y., Nagy, Z.K., 2016. Ind. Eng. Chem. Res. 55, 4987–4996. Copyright 2016 American Chemical Society. http://pubs.acs.org/doi/abs/10.1021/acs.iecr.5b04956

144

Therefore, in this chapter, we would like to introduce a novel approach called

WMADNC, which is a technique that combines the advantages of WM and ADNC. This approach is aligned with both the QbD and the new QbC concept in pharmaceutical manufacturing (Yang et al., 2015e). It implements the continuous ADNC closed-loop approach in a continuous wet mill unit which behaves as a perfect nucleator. The idea is to use the total particle chord counts measured by FBRM in the crystallizer to manipulate the jacket temperature of the wet mill in closed-loop. By manipulating the temperature in the wet mill the supersaturation at which nucleation happens is controlled. Because of the very short residence time in the wet mill this approach controls the nucleation kinetics directly in the wet mill without being influenced by the growth kinetics. The proposed WMADNC concept is integrated with a continuous MSMPRC in both upstream and downstream configurations to achieve closed-loop controlled primary or secondary nucleation kinetics.

This WMADNC approach should not only contain all the advantages of WM and ADNC, but also have better control performance on particle size due to decoupled nucleation and growth kinetics. Paracetamol, which is an API, and ethanol are used as solute and solvent, respectively. Performances of WMADNC under startup, steady-state, set-point change and disturbance are studied for continuous cooling crystallization in the MSMPRC.

8.2 Experimental Section

8.2.1 Upstream Application of WMADNC Approach (Process 1)

The experimental setup for process 1 is shown in Figure 8.1(a). The feed vessel and

MSMPRC were 1 liter round bottom jacketed vessel. A magic LAB rotor-stator wet mill unit (IKA) was installed between the feed vessel and MSMPRC. A medium and a fine

145 generators were used in the wet mill. The angular speed of the wet mill was set to 10,000 rpm in all the experiments. Three temperature bathes (Huber) were connected to the jackets of feed vessel, MSMPRC and wet mill. Clear feed solution was transferred from the feed vessel directly into the wet mill through a peristaltic pump (Masterflex). Both wet mill and

MSMPRC were equipped with thermocouples to measure the process temperatures. In addition, a FBRM ParticleTrack G400 probe (Mettler-Toledo) was installed in the

MSMPRC to provide real-time information of particle chord counts and CLD. The FBRM information was measured every 15 seconds and collected in the software iC FBRM 4.3, which communicated via an RS232 interface with the CryMOCO software that was installed on a supervisory computer. All temperature measurements were also collected in

CryMOCO. Slurry from the MSMPRC was transferred out semi-continuously through a transfer unit that used vacuum and nitrogen gas.

The feed solution was prepared from paracetamol (Alfa Aesar, purity > 98%) and ethanol (KOPTEC, purity > 99%) at a concentration of 0.26 g/g, which corresponds to solubility at 38 oC (Mitchell and Frawley, 2010). The feed flow rate was kept at 20 mL/min in all the experiments. Every 40 seconds around 13 mL slurry was transferred out from the

MSMPRC through the transfer unit. The total volume including wet mill and MSMPRC was maintained at around 400 mL. Therefore, the total RT was 20 minutes. The jacket temperatures of feed vessel and MSMPRC were kept at 55 oC and 25 oC, respectively, in all the experiments. The feed solution, final stage temperature and residence times are the same for all the experiments, therefore the yields are comparable in all the experiments.

The product slurry was recycled for preparing feed solutions in order to save API (Power et al., 2015; Yang et al., 2015d, 2015e). Slurry samples were taken at steady-states and

146 filtered immediately. The sample crystal images were taken using SMZ1500 optical microscope (Nikon). DXR Raman microscope (Thermo Scientific) was used for polymorph analysis with 532 nm laser wavelength and 8 mW laser power.

8.2.2 Downstream Application of WMADNC Approach (Process 2)

Figure 8.1(b) presents the experimental setup for process 2. Compared to process 1, the only difference was that the wet mill was connected to the MSMPRC through external recycling. In this case the wet mill worked as a downstream size reduction unit under high shear secondary nucleation and breakage mechanisms. All the process operating conditions

(e.g. feed flow rate, operating volume, temperatures, etc.), wet mill configurations (e.g. generators, angular speed) and measurements settings used in process 2 were exactly the same as in process 1. In addition, the recycling flow rate through the wet mill in process 2 was 200 mL/min, which can be used to calculate the TOPRT (Yang et al., 2015d):

qq TOPRT =mill = mill (8.1) VqMSMPRC

where  is the MSMPRC residence time, qmill and qMSMPRC are the recycling flow rate through wet mill and the feed flow rate through the MSMPRC, respectively. In this case a

TOPRT = 10 was used. The physical meaning of this TOPRT value is that on average, the slurry was milled 10 times before it was washed out from the MSMPRC. The TOPRT is an important operating parameter in integrated continuous crystallization and wet milling operation since it defines how frequently the slurry is milled (Yang et al., 2015d).

147

Figure 8.1. Experimental setups for (a) process 1 and (b) process 2.

148

8.2.3 WMADNC Methodology in Continuous Cooling Crystallization

WMADNC is a FBRM based closed-loop (feedback) control approach that we propose to control primary nucleation in process 1, or secondary nucleation in process 2 directly in the wet mill therefore control the final CLD. Both primary and secondary nucleation kinetics are strong functions of supersaturation:

nprim Bprim k prim c (8.2)

nsec Bsec k sec c m2 (8.3)

where Bprim and Bsec are primary and secondary nucleation rates, kprim and ksec are

primary and secondary nucleation rate constants, nprim and nsec are primary and secondary

nucleation orders, c is absolute supersaturation, and m2 is the second moment of the

CSD. Since the real-time particle total chord counts measurement from FBRM can reflect total particle number and nucleation kinetics in the system, it is used in the closed-loop control to automatically adjust the jacket temperature set-point of the wet mill therefore control supersaturation and primary or secondary nucleation rate directly inside the wet mill. Similar as the continuous ADNC algorithm (Yang et al., 2015e), first of all an upper and a lower bounds for desired particle total chord counts set-point must be defined in

WMADNC (Figure 8.2). During the operation, if the measured total counts was found smaller than the lower bound, a pre-defined cooling rate would be actuated automatically by CryMOCO to cool down the wet mill jacket therefore increase supersaturation and hence nucleation rate in the wet mill, and vice versa. If the measured total counts fell between the upper and the lower bounds, then the wet mill jacket temperature would be kept constant by CryMOCO in order to maintain total counts and steady-state operation. In

149 this work, heating and cooling rates were set as ± 0.1 oC/min in all the experiments that used WMADNC. The initial wet mill jacket temperature was set to 20 oC in all the experiments.

Figure 8.2. WMADNC methodology.

8.3 Results and Discussion

8.3.1 Upstream Application of the WMADNC Approach to Achieve Closed-Loop

Controlled Primary Nucleation (Process 1)

In process 1 (Figure 8.1(a)), the wet mill behaves as an upstream in situ seed generator under high shear primary nucleation mechanism since the feed solution is clear. It produces seed crystals for the MSMPRC, which acts as a growth vessel. WMADNC approach is used upstream to achieve closed-loop controlled primary nucleation kinetics. The dynamics of process 1 applied with WMADNC approach under startup, steady-state and disturbance were studied.

150

8.3.1.1 Startup and Steady-State Dynamics

The startup and steady-state dynamics of process 1 with the WMADNC approach applied is plotted in Figure 8.3(a), which includes milling outlet temperature, milling jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point and FBRM SWMCL. The upper and lower bounds of the WMADNC total counts set-point are 18,500 and 16,000, respectively.

(a) Mill outlet temperature Mill jacket setpoint 52 Total counts 24000 140 WMANDC set-point 48 120 [ Length Chord Mean SWMCL 20000 44

100 C]

Counts [#/s] Counts o 40 16000 80 36 12000 32 60 8000

28 40 Temperature[

 24 m] 4000 20 20 0 0 0 1 2 3 4 5 6 7 8 9 No. of Residence Time (RT) [-]

(b) 30000 Total Counts Without control 25000 Total Counts With WMADNC

20000

15000

Counts[#/s] 10000

5000

0 0 1 2 3 4 5 6 7 8 9 10 No. of Residence Time (RT) [-]

Figure 8.3.(a) Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL during startup and steady-state operation in process 1 with WMADNC approach. (b) Comparison of total particle chord counts dynamics in process 1 without control and with WMADNC approach.

151

The fast increase of total counts indicates large primary nucleation rate at the beginning of the process. Then once the total counts overshoots above the set-point, it can be seen that the WMADNC approach automatically heats up the wet mill jacket to reduce the primary nucleation rate. Finally, two oscillations of total counts are observed during startup before it converges to the desired WMADNC set-point and steady-state operation. In this work, steady-state is assumed reached once the measured total counts are around the

WMADNC set-point for 2 RT or longer (Yang et al., 2015e). Figure 8.3(b) presents the comparison of the total counts dynamics for process 1 with WMADNC approach implemented and case without control when mill jacket and outlet temperatures are constant. It is clear that both experiments end up with two oscillations during startup and reach steady-state at a similar time. Therefore, the WMADNC approach in this case does not show any improvement in the startup duration. Because the primary nucleation kinetics in wet mill is too fast and too sensitive to supersaturation therefore change in milling temperature is unfavorable to stabilize startup. This observation proves that closed-loop control for upstream wet mill which works as in situ nucleator is redundant during startup operation.

8.3.1.2 Set-Point Change and Control of CLD

After steady-state 1 is reached in process 1 with WMADNC approach, the upper and lower bounds of WMADNC total counts set-point are reduced to 12,500 and 10,000. As shown in Figure 8.4, when the total counts set-point is reduced, the wet mill jacket temperature automatically increases to decrease supersaturation and hence nucleation rate in the wet mill. The process is able to reach the new total counts set-point and steady-state

152

2 very quickly. The SWMCL in steady-state 2 is much larger than in steady-state 1 due to the reduced nucleation rate and increased supersaturation level in the MSMPRC for crystal growth.

Mill outlet temperature 56 Mill jacket setpoint 20000 100 Total counts 52 WMADNC set-point SWMCL 16000 90 48 Steady-state 1 [ Length Chord Mean 44 12000 80

C] Steady-state 2

Counts [#/s] Counts o 40 8000 70 36 32 4000 60

Temperature[ 28

0 50 m] 24 20 -4000 40 7 8 9 10 11 12 13 14 No. of Residence Time [-]

Figure 8.4. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 1 with

WMADNC approach under set-point change.

SWCLDs of these two steady-states shown in Figure 8.5 confirm the increase of average crystal size. It is noticed that many more large crystals and less fines can be obtained by decreasing the total counts set-point in process 1. Therefore, this WMADNC approach provides an efficient and novel way to directly control CLD by adjusting the WMADNC total chord counts set-point. Additionally, it can be noticed that at the end of steady-state 1 zone in Figure 8.4, the total counts drops slightly below set-point lower bound and shows a downward trend. This extent of variation of FBRM data is very common in continuous

MSMPRC experiments in literature (Hou et al., 2014; Peña and Nagy, 2015; Yang et al.,

2015d, 2015e) because of unavoidable disturbances in continuous crystallization experiments, including variation of slurry volume due to semi-continuous transferring,

153 non-ideal transferring due to not well-mixed slurry, among others. Therefore, perfect steady-state operation and FBRM data is never achievable in real experiments, and only quasi steady-state is achievable. However, for simplification the term “steady-state” will be used in this chapter instead (Yang et al., 2015e). Since the FBRM data has already been stabilized for more than 2 RT in steady-state 1, and total counts stopped decreasing right after steady-state 1 zone, it is reasonable to assume steady-state 1 is reached and maintained.

In addition, in Figure 8.4 as soon as the total counts drops below the set-point lower bound, it can be seen that the controller actuates a cooling rate to counteract the disturbance, and very soon the total counts stopped decreasing. This shows the ability of WMADNC to counteract disturbance. More detailed study of WMADNC’s ability to reject disturbance will be introduced in the next section.

Steady-state 1 Steady-state 2 14

12

10

8

6 Counts[#/s] 4

2

0 10 100 Chord Length [m]

Figure 8.5. SWCLDs of steady-state 1 and steady-state 2 in process 1 with WMADNC approach.

154

8.3.1.3 Disturbance Rejection

Since the WMADNC is a closed-loop nucleation control approach, its performance to reject disturbance is also evaluated. According to Figure 8.6, at first, process 1 is at steady- state 2 with the WMADNC approach enabled. Then suddenly a step disturbance is introduced on purpose by changing the angular speed of the wet mill from 10,000 rpm to

6,000 rpm. Decrease of the wet mill angular speed can lead to two opposite effects. On one hand, the reduced shear rate decreases nucleation rate. On the other hand, the reduced angular speed decreases the energy introduced therefore decreases milling temperature and increases supersaturation and primary nucleation rate.

Mill outlet temperature 60 Mill jacket setpoint Mill@ Mill@ Total counts 84 56 14000 10k rpm 6k rpm WMADNC set-point 52 SWMCL 81 [ Length Chord Mean 12000

48 [#/s] Counts C]

o Steady-state 3 78 44 Steady-state 2 40 10000 75 36 8000 72

32 Temperature[

 28 69 m] 6000 24 66 16.0 16.5 17.0 17.5 18.0 18.5 No. of Residence Time (RT) [-]

Figure 8.6. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 1 under disturbance with WMADNC approach.

It is observed that the total counts and SWMCL start to decrease and increase, respectively, after the step disturbance is introduced. Therefore, in this case the first effect is dominating and primary nucleation rate is reduced. But once the measured total counts drop below the set-point lower bound, the disturbance is noticed by the WMADNC

155 approach which then cools down the wet mill automatically to enhance supersaturation and nucleation rate in the wet mill. Finally, both total counts and SWMCL are able to return to their original values although the disturbance still exists, and steady-state 3 is reached. The

CLDs of steady-state 2 and steady-state 3 in this test (Figure 8.7) demonstrate that the disturbance is successfully rejected, and same CLD is obtained before and after this disturbance.

Steady-state 2 12 Steady-state 3

10

8

6

Counts[#/s] 4

2

0 Chord Length [m]

Figure 8.7. SWCLDs at steady-state 2 and steady-state 3 in process 1 with WMADNC implemented upstream.

The microscope images and Raman spectra of crystals obtained in steady-state 1, steady-state 2 and steady-state 3 during process 1 servo control (i.e. set-point change) and regulatory control (i.e. disturbance change) tests are presented in Figure 8.8 and Figure 8.9, respectively. Figure 8.8 confirms that CLD shifts towards larger size from steady-state 1 to steady-state 2 due to reduced primary nucleation rate, and CLD does not change from steady-state 2 to steady-state 3 though disturbance exists. In addition, the Raman spectra

156 of samples presented in Figure 8.9 indicate that monoclinic paracetamol (Form I, stable form, octahedrons) rather than orthorhombic paracetamol (Form II, metastable form, needle-like) crystals are obtained in all three steady-states (Barthe et al., 2008; Nanubolu and Burley, 2012; Thorley et al., 2006). This observation agrees with literature in which wet mill was found to facilitate the formation of stable form crystals (Lo et al., 2012).

Figure 8.8. Microscope images of crystals obtained in (a) steady-states 1, (b) steady-state

2 and (c) steady-state 3 in process 1 with WMADNC implemented upstream.

12000

9000 Steady-state 1

6000

Steady-state 2 Counts[a.u.] 3000 Steady-state 3

Form I 0 400 800 1200 1600 2000 2400 2800 3200 Raman shift [cm-1]

Figure 8.9. Raman spectra of crystals obtained in steady-state 1, steady-state 2 and steady- state 3 in process 1 with WMADNC implemented upstream.

157

8.3.2 Downstream Application of the WMADNC Approach to Achieve Closed-Loop

Controlled Secondary Nucleation (Process 2)

In process 2 (Figure 8.1(b)), the wet mill works as a downstream size reduction unit under high shear secondary nucleation and breakage mechanisms. It can efficiently reduce crystal size. The WMADNC approach is used downstream to achieve closed-loop controlled secondary nucleation kinetics. The application of the WMADNC approach in process 2 under startup, steady-state and disturbance are investigated next.

8.3.2.1 Startup and Steady-State Dynamics

Process variables during operation of process 2 with the WMADNC approach enabled are presented in Figure 8.10(a). The upper and lower bounds of the WMADNC set-point are selected as 13,000 and 10,000, respectively. Within the first RT, the particle total chord counts increases very fast due to the high shear enhanced secondary nucleation kinetics, which leads to an overshoot of the total counts. As a result, the WMADNC approach heats up the wet mill jacket to reduce supersaturation and secondary nucleation rate. Finally, the total counts automatically converges to the selected set-point after one oscillation. Figure

8.10(b) presents a comparison of total counts in process 2 without control (i.e. constant mill temperature) and with WMADNC approach. Similar as in process 1, WMADNC approach does not improve startup in process 2 either. This again confirms that change in milling temperature is unfavorable to stabilize the startup dynamics since the nucleation kinetics in the wet mill is too fast and too sensitive to the milling temperature.

158

(a) Mill outlet temperature Mill jacket setpoint 24000 100 55 Total counts WMADNC set-point 20000

50 SWMCL [ Length Chord Mean 16000 80

45 C]

12000 [#/s] Counts o 40 8000 60 35 4000 30 0

Temperature[ 40 -4000  25 m] -8000 20 -12000 20 0 1 2 3 4 5 6 7 8 9 10 No. of Residence Time (RT) [-]

(b) 30000 Total Counts Without control 25000 Total Counts With WMADNC

20000

15000

Counts[#/s] 10000

5000

0 0 1 2 3 4 5 6 7 8 9 10 11 No. of Residence Time (RT) [-]

Figure 8.10. (a) Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL during startup and steady-state operation in process 2 with WMADNC approach. (b) Comparison of total particle chord counts dynamics in process 2 without control and with WMADNC approach.

8.3.2.2 Set-Point Change and Control of CLD

As shown in Figure 8.11, the upper and lower bounds of the WMADNC set-point are reduced to 8,000 and 6,000, respectively, when the process is at steady-state 1. This set- point change results in the WMADNC adapting the heating of the wet mill jacket to decrease supersaturation and secondary nucleation rate. The total counts is able to reach the new set-point after a few RT. At the same time, it can be seen from Figure 8.11 that the

159

SWMCL increases from steady-state 1 to steady-state 2. In addition, the SWCLDs of both steady-states are compared in Figure 8.12.

Mill outlet temperature Mill jacket setpoint 100 56 Total counts 12000 52 Steady-state 4 WMADNC set-point SWMCL 90 48 [ Length Chord Mean

Steady-state 5 8000 80 C]

44 [#/s] Counts o 40 70 36 4000 32 60

Temperature[ 28

0 m] 24 50 20 40 8 9 10 11 12 13 14 15 16 17 18 No. of Residence Time (RT) [-]

Figure 8.11. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 2 with

WMADNC approach under set-point change.

Steady-state 4 6 Steady-state 5

5

4

3

Counts[#/s] 2

1

0 10 100 Chord Length [m]

Figure 8.12. SWCLDs of steady-state 4 and steady-state 5 in process 2 with WMADNC approach.

160

It is clear that the number of fine crystals decreases significantly from steady-state 1 to steady-state 2. This also provides an excellent approach to tailor CLD and control the number of fine crystals in the system by adjusting the WMADNC set-point.

8.3.2.3 Disturbance Rejection

The closed-loop disturbance rejection performance of the WMADNC approach in process 2 is also tested. As shown in Figure 8.13, the wet mill angular speed is suddenly changed from 10,000 rpm to 13,000 rpm when the system is at steady-state. The two opposite effects of changing wet mill angular speed described in process 1 also exist in this process.

Mill outlet temperature 72 Mill jacket setpoint 100 Total counts 68 Mill@ Mill@ WMADNC set-point 10000 64 10k rpm 13k rpm SWMCL 90 60 [ Length Chord Mean 8000 56 Steady-state 6

C] Steady-state 5

Counts [#/s] Counts o 52 80 48 6000 44 70 40 4000 36

Temperature[ 32 60  28 2000 m] 24 20 0 50 22 23 24 25 26 27 No. of Residence Time (RT) [-]

Figure 8.13. Milling outlet temperature and jacket temperature set-point, FBRM total particle chord counts, WMADNC total counts set-point, and SWMCL in process 2 with

WMADNC approach under disturbance.

It is clear that in this process the temperature effect of wet milling is dominating, and total counts and SWMCL begin to decrease and increase respectively due to increased milling temperature and reduced secondary nucleation rate. The decrease in total counts

161 automatically activates the WMADNC approach, which cools down the wet mill jacket immediately and finally brings total counts and SWMCL back to their initial steady-state values. The SWCLDs shown in Figure 8.14 confirm that the step disturbance is completely rejected by WMADNC approach and same CLD is achieved consistently under disturbance.

Steady-state 5 6 Steady-state 6

5

4

3

Counts[#/s] 2

1

0 Chord Length [m]

Figure 8.14. SWCLDs of steady-state 5 and steady-state 6 in process 2 with WMADNC implemented downstream.

Figure 8.15 and Figure 8.16 present the microscope images and Raman spectra of crystals obtained in steady-state 4, steady-state 5 and steady-state 6 in process 2, respectively. The results confirm that less fine particles are present in steady-state 5 than steady-state 4 due to reduced secondary nucleation rate, and the disturbance introduced does not change the CLD from steady-state 5 to steady-state 6. In addition, the stable form of paracetamol (Form I) is obtained in all three steady-states, indicated by the identical

Raman spectra of all samples, shown in Figure 8.16.

162

Figure 8.15. Microscope images of crystals obtained in (a) steady-states 4, (b) steady-state

5 and (c) steady-state 6 in process 2 with WMADNC implemented downstream.

15000

12000 Steady-state 4

9000

Steady-state 5

6000 Counts[a.u.]

3000 Steady-state 6

Form I 0 400 800 1200 1600 2000 2400 2800 3200 Raman shift [cm-1]

Figure 8.16. Raman spectra of crystals obtained in steady-state 4, steady-state 5 and steady- state 6 in process 2 with WMADNC implemented downstream.

The FBRM based closed-loop WMADNC approach described in this work should be applicable to not only continuous MSMPRC, but also other types of continuous crystallizers such as continuous plug flow crystallizer. It can be envisioned that both upstream and downstream applications of WMADNC can be integrated with continuous

163 plug flow crystallization processes to achieve closed-loop controlled primary or secondary nucleation kinetics, therefore provide superior control on CLD and reject disturbances. In addition, the proposed WMADNC approach does not require any modelling effort or knowledge about solubility or crystallization kinetics of the compound. The only requirement is a few experiments to tune the control parameters such as heating and cooling rates. Therefore, the WMADNC approach provides a very fast approach for highly automated recipe design for continuous crystallization processes.

8.4 Conclusions

A closed-loop WMADNC approach is proposed and implemented in laboratory scale integrated continuous MSMPRC-wet mill system. The proposed approach is demonstrated to provide closed-loop controlled primary or secondary nucleation kinetics by using it upstream or downstream in continuous cooling crystallization. Closed-loop WMADNC approach does not show any improvement in the startup of either applications, which indicates that closed-loop control is redundant for wet milling during startup. However, during steady-state operation, the proposed WMADNC is proved to be an excellent approach to tailor CLD by adjusting the total chord counts set-point in both applications therefore can directly control the CQA. In addition, WMADNC can provide very fast and effective disturbance rejections. Same CLD can be achieved consistently under disturbances in both processes using WMADNC upstream or downstream.

164

CHAPTER 9. APPLICATION OF ULTRA-PERFORMANCE LIQUID

CHROMATOGRAPHY AS AN ONLINE PROCESS ANALYTICAL

TECHNOLOGY TOOL IN PHARMACEUTICAL CRYSTALLIZATION13

9.1 Introduction

This chapter presents the initial application of UPLC combined with automated online process sampling and dilution as a PAT tool for the systematic study of pharmaceutical crystallization. Crystallization of two APIs, including caffeine and acetylsalicylic acid

(ASA, or aspirin) were investigated. The first section demonstrates how online UPLC can be implemented as a method for concentration measurement in crystallization and degradation of APIs. The second section compares online UPLC with common spectroscopy techniques (i.e. ATR-UV/vis and Raman) and introduces a novel method that uses online UPLC data to simplify the spectroscopy calibration process. The new method can be fully automated and can serve as a continuous probe validation and recalibration procedure. The last section describes development and verification of online UPLC for real-time crystallization product purity monitoring.

13 Reproduced with permission from Cryst. Growth Des., submitted for publication. Unpublished work copyright 2016 American Chemical Society.

165

9.2 Experimental Section

9.2.1 Materials

Caffeine and aspirin (ASA) were studied as model APIs in this work. Salicylic acid

(SA) or paracetamol (PCM) were spiked as impurities in the aspirin crystallization process.

Water, acetonitrile (ACN), methanol (MeOH), and acetic acid (AA) were used in mobile phases in UPLC methods for different compounds. The manufacturer and purity of materials used in this work are summarized in Table 9.1.

Table 9.1. Manufacturer and purity of materials used.

Name Manufacturer Purity or Grade Caffeine Fisher Chemical ≥98.5% Aspirin (ASA) Alfa Aesar ≥99% Salicylic acid (SA) Fisher Chemical ≥99.5% Paracetamol (PCM) ACROS Organics ≥98% Water Fisher Chemical Optima LC/MS grade Acetonitrile (ACN) Fisher Chemical Optima LC/MS grade Methanol (MeOH) Fisher Chemical Optima LC/MS grade Acetic acid (AA) Fisher Chemical ≥99.5%

9.2.2 Experimental Setup

The experimental setup is presented in Figure 9.1. The crystallizer was a one-liter jacketed glass vessel (LABTEX) with overhead stirring, which was connected to a refrigerated/heated circulating temperature bath (Huber, Ministat 125). FBRM (Mettler-

Toledo, G400, software IC FBRM 4.3), ATR-UV/vis (Zeiss, MCS621, software

ProcessXplorer 1.3), Raman immersion optic probe (Kaiser Optical Systems, RXN1-785,

166 software IC Raman 4.1), a Pt100 thermocouple and a condenser were installed on the crystallizer. The in situ measurements from the above tools were communicated to

CryMOCO, also called CryPRINS via an RS232 interface (Yang et al., 2015d, 2015e,

2016). The temperature bath was controlled by CryMOCO in closed-loop. The online

UPLC system with integral process sampling and dilution used in this work (Waters,

PATROL UPLC®, with Empower 3 software) consists of binary solvent manager, process sample manager, column manager and an UV detector. A stainless steel inlet solvent filter with 2 µm pore size (Upchurch Scientific) was installed at the end of the PATROL UPLC® process sample line immersed in the crystallizer to enable sampling of just the mother liquor from the crystallization slurry. The PATROL UPLC® process sample manager outline was routed back to the vessel to recycle and conserve unused sample solution.

Robust sampling for online UPLC from hot supersaturated crystallization solutions was challenging due to the strong potential for APIs to precipitate and clog the sampling loop.

To avoid these problems, the process sample line, recycle line and all of the process sample manager sample tubing were enwrapped by specially designed tubing heaters that were connected to a power supply providing heating current. Additionally, hot air was blown into the process sample manager to keep the sample valve and injection valve hot. The temperatures of tubing heaters and hot air were controlled as 10–15 oC higher than the maximum saturation temperature of the solution. To avoid the influence of dead volume during each measurement, the process sample line was primed by 2 mL sample before auto- dilution in the process sample manager. The UPLC data was processed with Empower 3 software (Waters).

167

Figure 9.1. Schematic of experimental setup.

9.2.3 UPLC Methods

The UPLC methods used in this work are summarized in Table 9.2. Representative chromatograms of the three APIs are presented in Figure 9.2. The minimum run time required in all three methods was less than 2 minutes. Combined with the one and half minute sampling and dilution of the PATROL UPLC®, a minimum total cycle time of 3.5 minutes can be achieved. This fast measurement time enabled the use of UPLC as an online

PAT to achieve real-time process monitoring and control. If the concentration changes very slowly or frequent measurement is not required, one can use longer measurement intervals by setting up a delay between each measurement in Empower 3 in order to save materials.

In each method, calibration curves were calculated based on measured peak areas of five

168 to seven standard solutions of various concentrations and linear least square regression analysis (Pickering and Brown, 2012).

Table 9.2. Isocratic UPLC methods used in this work for different compounds.

Caffeine Aspirin + impurities Column Waters BEH C18 Waters BEH C18 1.7 µm particles 1.7 µm particles 2.1*50 mm 2.1*100 mm Column temperature 40 oC 40 oC Mobile phase A Water 1.5% AA in water Mobile phase B ACN MeOH Ratio A:B 92:8 55:45 Flow rate 0.6 mL/min 0.5 mL/min Diluent Mobile phase Mobile phase UV wavelength 273 nm 275 nm Injection volume 2 µL 2 µL Dilution factor 100 100 Run time 1.7 minutes 2.5 minutes

0.7 0.10 (a) (b) 0.6 Caffeine 0.08 Aspirin 0.5 0.06

0.4

AU AU 0.3 0.04 Salicylic acid

0.2 0.02 0.1

0.0 0.00

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0.0 0.5 1.0 1.5 2.0 2.5 Minutes Minutes

Figure 9.2. Example chromatogram of (a) caffeine and (b) aspirin.

169

9.3 Results and Discussion

9.3.1 Online UPLC-Based Crystallization and Degradation Monitoring

Caffeine crystallization from water was carried out to demonstrate the use of online

UPLC for real-time concentration measurement. Caffeine was chosen as the first model compound because it is known to have a strong tendency to create fouling (Wu and Bird,

2007), making it challenging for many PAT tools such as FBRM, PVM and ATR-UV/vis.

Temperature Concentration from online UPLC 0.040 80000 30 Total counts from FBRM 30 Mean chord length (no wt.) from FBRM 70000 0.035 25

Mean Chord Length [ Length Chord Mean 60000 25 0.030 [g/g] Concentration 20

Counts [#/s] Counts C]

o 50000 0.025 20 Fouling 40000 15 0.020 30000 10 15

0.015 20000 Temperature[

Seeding m] 0.010 10000 5 10 0.005 0 0 0 1 2 3 4 5 6 Time [h]

Figure 9.3. Temperature, concentration, total counts and unweighted mean chord length during caffeine crystallization process.

The process parameters, including temperature, concentration measured by online

UPLC, total chord counts and unweighted mean chord length measured by FBRM, are plotted in Figure 9.3. The profile for online UPLC measured concentration agrees well with the total chord counts measured from FBRM in terms of the nucleation time, indicating that there is minimal delay in the online UPLC measurement. After 5 hours, significant caffeine fouling started to occur on the crystallizer wall and the FBRM measured total

170 chord counts and mean chord length also started to decrease and increase, respectively, due to probe fouling. However, Figure 9.3 indicates that the online UPLC continued to provide accurate and robust concentration measurement, even in the presence of significant fouling.

Because UPLC can measure concentrations for multiple components, it is very useful for monitoring degradation of unstable APIs. The hydrolysis of aspirin (ASA) to salicylic acid (SA) is a common issue in aspirin crystallization, especially at high temperature

(Bakar and Niazi, 1983), making crystallization of ASA in water an ideal case study.

70 1.1 Temperature 16 60 ASA concentration from online UPLC 1.0

SA concentration from online UPLC 14 [mg/ml] concentration ASA 0.9 50 [mg/ml] concentration SA 12 0.8 40

C] 0.7

o 10 30 0.6 8 0.5 20 6 0.4 10 ASA added 4 0.3

Temperature[ 0 2 0.2 0.1 -10 0 0.0 -20 0.0 0.5 1.0 1.5 2.0 2.5 Time [h]

Figure 9.4. Temperature and concentrations of aspirin (ASA) and degradation byproduct salicylic acid (SA), during crystallization of aspirin in water.

As shown in Figure 9.4, a 55 oC saturated ASA solution was prepared and crystallized by cooling from 60 oC to 15 oC. Both ASA and SA concentration were monitored successfully during the ASA crystallization process. No sample line clogging was observed

171 even though a hot solution was sampled and tested. Figure 9.4 indicates that SA concentration starts to increase at 0.6 hours due to hydrolysis of ASA, but it begins to decrease at 2.3 hours because of its absorption into ASA crystals (Borsos et al., 2016).

9.3.2 Online UPLC-Based Quick Calibration for UV/vis and Raman Spectroscopy

Although spectroscopy-based concentration measurements have some disadvantages, as described in the introduction, spectroscopic techniques such as UV/vis, IR and Raman can provide much faster measurements (i.e. a few seconds) than are possible with online

UPLC. Therefore there are still many situations in which spectroscopy cannot be replaced by online UPLC, for example when fast closed-loop process control is required. However, as described in the introduction, one major disadvantage of spectroscopy-based concentration measurement is its relatively complicated and time-consuming standard calibration procedure, which requires spectroscopy measurements at different concentrations, temperatures, etc (Saleemi et al., 2012b). Due to the difficulty of measuring concentration in a slurry, clear solutions are widely used in the standard calibration approach because their concentrations can be simply calculated from mass balance

(Saleemi et al., 2012b). As a result, only a limited metastable zone can be covered in this type of standard calibration approach because concentration will become unknown once the solute starts to crystallize at high supersaturation. Because of the ability of online UPLC to measure solute concentration in slurries, this section introduces an online UPLC-based quick calibration approach for UV/vis and Raman spectroscopy that overcomes these limitations. Simultaneous online UPLC and ATR-UV/vis (or Raman) measurements were made for crystallization and dissolution experiments to cover the boundary of a large

172 metastable zone, and then ATR-UV/vis or Raman signals were fitted with online UPLC measured concentration, and finally the model was validated in another validation run by comparing the calibration calculated and UPLC measured concentrations. In this quick calibration approach, the model was trained and used in similar conditions (e.g. slurry density, supersaturation range), so it is expected to be robust.

The first case study employed aspirin in ethanol as a model system. The aspirin UV/vis first derivative spectrum and Raman raw spectrum are presented in Figure 9.5. The peak heights at 267 nm (UV/vis) and 873 cm-1 (Raman) of the two spectra were used to create univariate calibration models. The peak locations were selected by trial and error to minimize the influences from particles.

16000 (a) Aspirin UV/vis 1st derivative spectrum (b) Aspirin Raman spectrum 0.010 14000

267nm peak 12000 0.005 -1 873cm peak 10000 0.000

8000

-0.005

6000

1st Derivative [-] Intensity [counts] -0.010 4000

2000 -0.015 0 200 300 400 500 600 700 2800 2400 2000 1600 1200 800 400 Wavelength [nm] Raman shift [cm-1]

Figure 9.5. (a) UV/vis first derivative spectrum, and (b) Raman spectrum of a 25 oC saturated aspirin-ethanol solution.

The calibration data are plotted in Figure 9.6. In the first 15 hours, an unseeded fast cooling crystallization was carried out followed by a slow heating experiment. This part of data was used to train the calibration model. The purpose of running fast cooling and slow

173 heating during model training is to make the calibration trajectory cover a large metastable zone in order to avoid extrapolation when using the calibration model in crystallization experiments. The following equation was used to fit the UV/vis or Raman calibration model:

Concentration [mg/g] = aT bp c (9.1) where T is temperature [oC], p is peak height at 267 nm in UV/vis first derivative spectrum or 873 cm-1 in Raman spectrum of aspirin, a, b , c are fitted coefficients. The

UV/vis first derivative spectrum was used to minimize the baseline offsets. From 15 hours to 28 hours, further cooling and heating experiments were implemented using different seed loadings, cooling and heating rates to validate the calibration model. The coefficients were fitted using a least square regression method in software MiniTab 17. The fitted coefficients, root mean square error (RMSE) as well as maximum and average errors of fitted data in the validation run are summarized in Table 9.3. Although liquid chromatography itself is an accurate tool, the online sampling of UPLC may still introduce errors. For example, every sampling would take about one and half minutes, therefore it is not accurate if concentration changes very quickly within the sampling window, such as when nucleation occurs. The errors shown in Table 9.3 are at the similar level as the standard approach reported in literature (Saleemi et al., 2012b). But the proposed online

UPLC-based quick calibration approach is much simpler and more time efficient (1 day vs. 1 week) than the standard approach used in literature.

In Figure 9.6 one may notice that at around 2 hours, right after primary nucleation, there was a sharp UV/vis concentration dip that significantly underestimates the aspirin concentration. This is a common phenomenon in UV/vis measurement due to the light

174 absorption and scattering of fine particles or small nuclei (Van Eerdenbrugh et al., 2011).

The measurement returned to normal values after a while when the nuclei grew larger. The

UPLC measurements are unaffected by these interferences, providing additional evidence that online UPLC is a more robust PAT tool than spectroscopy in these slurry systems.

Temperature Measured concentration from online UPLC Predicted concentration from UV/Vis Predicted concentration from Raman 300 Unseeded fast cooling -1 oC/min Seeded cooling 5% seeds, -0.1 oC/min 275 20 Slow heating o Heating 0.03 C/min [mg/g] Concentration

0.05 oC/min 250 C] o 0 225

200 -20

175 Temperature[ 150 -40 Model training Model validation 125 0 4 8 12 16 20 24 UV/vis dip due to Time [h] significant nucleation

Figure 9.6. UPLC-based UV/vis and Raman quick calibration for aspirin.

Table 9.3. Fitted coefficients and errors of UV/vis and Raman quick calibration model for aspirin.

UV/vis Raman a [-] 0.6108 1.388 b [-] 1.332e5 4.071e-2 c [-] 31.18 13.98 RMSE [mg/g] 6.26 5.97 Maximum error 5.38% 6.65% Average error 3.51% 3.35%

175

9.3.3 Online UPLC for Real-Time Product Purity Monitoring

Based on the online UPLC measurement in the mother liquor, theoretically one can calculate the purity of product in real-time using the following mass balance equation:

c()() API c API Purity( t)  0 t (9.2) c00()()( API ctt API  c imp))  c( imp

where c0 () API and ct () API are API concentration at time 0 and t , c0 () imp and ct () imp are impurity concentration at time 0 and t . Two case studies were carried out using aspirin

(ASA) as API to verify the proposed approach.

In the first case study, salicylic acid (SA), which is the degradation product and common impurity for aspirin, was used as impurity. About 5% SA was spiked into aspirin raw materials on purpose. Ethanol was the crystallization solvent. As presented in Figure 9.7, a seeded cooling crystallization was implemented. Predicted ASA concentration was calculated from the UV/vis and Raman spectra based on the calibration models developed in the previous section. The measured ASA and SA concentrations were obtained from online UPLC. The predicted ASA product purity was calculated based on equation (9.2).

The measured ASA product purity was obtained by manually taking slurry samples, filtering, washing, dissolving and measuring offline in UPLC. It can be noticed that the SA concentration in mother liquor did not change much from initial value during the cooling process, indicating the ASA crystallization product purity should be near 100% throughout the crystallization according to equation (9.2). The final offline ASA crystallization product purity measurement agreed with this prediction. In addition, there was about a 10% offset between the predicted ASA concentration from UV/vis or Raman and the measured

ASA concentration from online UPLC. This observation indicates that UV/vis and Raman

176 are no longer accurate when the impurity is present due to the overlapped spectra. These data further illustrate how online UPLC is especially useful for multi-component systems because it can provide accurate concentration measurement for each component simultaneously.

Temperature Predicted ASA concentration from UV/vis 50 Predicted ASA concentration from Raman 600 1.5 Measured ASA concentration from online UPLC 40 Measured SA concentration from online UPLC 1.4 Offline measured ASA product purity 500 30 Predicted ASA product purity 1.3

0.1% ASA seeds [mg/g] Concentration

20 1.2 [-] purity ASA Crystal

400 C] o 10 1.1 0 300 1.0 -10 0.9 200

-20 0.8 Temperature [ -30 0.7 100 -40 0.6 -50 0 0.5 0.0 0.5 1.0 1.5 2.0 Time [h]

Figure 9.7. Real-time aspirin (ASA) product purity monitoring with salicylic acid (SA) present as impurity.

The second case study used PCM as the impurity because its structure is also similar to aspirin. Approximately 100% PCM was spiked into ASA and the two components were crystallized together at the same time. As shown in Figure 9.8, at about 1.4 hours ASA concentration started to drop when ASA seed crystals were introduced. When ASA finished crystallization at about 2.5 hours, PCM started to crystallize and its concentration started to drop. The predicted ASA product purity during this process was calculated using

177 equation (9.2) along with the online UPLC measurement of the mother liquor. Again, the predicted purity agreed well with the measured product purity obtained from manual sampling and offline analysis of the solid product. Therefore the proposed online UPLC- based real-time product purity monitoring approach was successfully verified.

Temperature Measured ASA concentration from online UPLC Measured PCM concentration from online UPLC 800 1.4 30 Offline measured ASA product purity Predicted ASA product purity 1.3 20 700 1.2 10 600 Concentration [mg/g] Concentration 1.1 0 [-] purity ASA Crystal

C] 500 1.0 o -10 0.9 400 -20 0.8 -30 0.1% ASA seeds 300 0.7 -40 0.6 Temperature [ 200 -50 0.5 100 -60 0.4 -70 0 0.3 0 1 2 3 4 5 6 7 Time [h]

Figure 9.8. Real-time aspirin (ASA) product purity monitoring with paracetamol (PCM) present as impurity.

9.4 Conclusions

In this work the PATROL UPLC® system with automated process sampling and dilution was implemented for online monitoring of pharmaceutical crystallizations. Three online

UPLC-based applications were proposed and implemented: (1) solute concentration monitoring during crystallizations and associated degradation analyses, (2) quick calibration of spectroscopic methods such as UV/vis and Raman spectroscopy, and (3) real-

178 time crystallization product purity monitoring. The results show that online UPLC is a powerful online PAT tool that can replace or improve the commonly used spectroscopy approaches, especially for crystallization processes in impure media. The advantages of online UPLC include simple calibration, and accuracy and robustness in the presence of impurities, fouling or fine particles. UPLC was also verified as an online PAT tool to monitor product purity in real-time for pharmaceutical crystallization.

In addition to the applications demonstrated here, online UPLC may also be used as a tool to control crystallization process in closed-loop. For example, one may have the online

UPLC communicate the measured solute concentration to crystallization control software

(e.g. CryMOCO or CryPRINS) so that the crystallization temperature can be adjusted dynamically to achieve supersaturation control (SSC) of the crystallization, similar to the

SSC approach reported in literature using ATR-UV/vis (Saleemi et al., 2012b). The advantage of online UPLC-based SSC is that it can work well even if impurities are present, whereas ATR-UV/vis-based SSC will likely have errors in the concentration measurements caused by overlapping peaks.

179

CHAPTER 10. CONCLUSIONS AND FUTURE WORK

10.1 Conclusions

The design and control of integrated crystallization operation and systems are important to meet the needs of improved CQAs, simplified process design and automated process control in the manufacturing of APIs. In this dissertation, three integrated crystallization operation and systems are investigated, including CCAC systems, integrated crystallization and wet milling operation, and crystallization systems in the presence of impurities.

CCAC systems implemented in both batch and continuous processes can be designed and controlled using PBM. In batch processes, it was found that based on the sensitivity of the solubility surface, the linear cooling and antisolvent addition profiles should either belong to cooling first or antisolvent first strategy to maximize crystal size and minimize batch time, depending on whether the initial solution locates at temperature sensitive region or antisolvent concentration sensitive region. The proposed PBM based design approach can be used to produce operating surfaces based on which the performance of CCAC systems can be understood and operating strategies rapidly designed depending on the initial concentration, solvent composition and desired yield. In continuous MSMPR processes, the optimal operating trajectory to maximize crystal size was proved to be similar as in batch processes and the knowledge was found transferrable using PBM based simulation and optimization techniques. The startup duration and waste amount of CCAC

180 in continuous MSMPRC can also be reduced significantly by implementing dynamic temperature or/and antisolvent addition profiles. In addition, simple decentralized PID control framework was demonstrated infeasible for such integrated CCAC-MSMPR systems, whereas NMPC framework was promising to control crystal size and yield simultaneously under disturbances or when set-point change was required.

Wet milling is an unit operation that can induce nucleation and break crystals, therefore improved control of CSD can be achieved by integrating it with crystallization. Integrated upstream and downstream applications of wet milling in continuous MSMPR cooling crystallization was investigated experimentally in this dissertation. The upstream wet mill can be used as a continuous in situ seed generator, and the downstream wet mill can efficiently reduce crystal size. Improved process yield and reduced width of CSD was also observed in such integrated operation. In addition, a FBRM-based closed-loop control approach, ADNC, was developed for continuous MSMPR cooling processes to achieve automated control of particle number and average crystal size under requirement of set- point change and disturbance rejection. This ADNC approach was then combined with wet milling technique, and a WMADNC approach was developed and implemented experimentally to achieve closed-loop controlled primary or secondary nucleation kinetics, by using it upstream or downstream. This combined WMADNC approach introduces the advantages of both WM and ADNC to continuous crystallization and superior control on

CSD was observed.

The characterization of crystallization systems in the presence of impurities is tough due to the limitation of PAT currently exists. In this dissertation, a PATROL UPLC® system with automated process sampling and dilution was implemented for online monitoring of

181 pharmaceutical crystallizations in impure media. This novel online PAT provides an efficient and automated approach to monitor concentrations of all components in the mother liquor in real-time. It can be used to replace the commonly used inline spectroscopy techniques (e.g. UV/vis, IR, Raman) which typically only work for single component systems, or to simplify the calibration procedures of inline spectroscopy techniques. A real- time crystallization product purity monitoring approach was also developed based on online UPLC and validated experimentally.

The modeling and experimental works of this dissertation indicate how PBM and PAT can be used to design and control integrated crystallization operation and systems, including CCAC systems, integrated crystallization and wet milling operation, and crystallization systems in the presence of impurities.

10.2 Future Work

Some recommendations for future work are as follows:

1) One limitation of the current model-based control approaches studied is the neglect of the uncertainties of models and crystallization kinetics. These uncertainties should be considered in the future works to achieve more robust design and control.

2) The NMPC control framework proposed in Chapter 5 can be implemented in the laboratory scale MSMPRC using FBRM, ATR-UV/vis, and software such as OptCon. The implementation of NMPC in continuous crystallization will be a milestone for the development of advanced control of continuous crystallization processes.

3) The integrated upstream and downstream applications of wet milling in continuous

MSMPRC can be expanded to continuous PFC in a similar way, to generate seed crystals

182 in situ or to reduce crystal size. The influences of wet milling on PFC startup duration, yield, average crystal size and width of size distribution can be investigated and recommendations of how to integrate wet milling with PFC can be proposed.

4) The closed-loop control approaches (i.e. ADNC and WMADNC) developed in

Chapter 7 and Chapter 8 can be expanded to continuous PFC to achieve closed-loop controlled primary or secondary nucleation kinetics, and automated control under disturbances or set-point change-over.

5) The online UPLC technique developed in Chapter 9 can be used to implement online UPLC-based SSC approach, which should be more robust than the commonly used

ATR-UV/vis-based SSC approach in impure media. In this feedback control approach, a crystallization control software (e.g. CryMOCO) can be used to analyze the real-time concentration measurement from online UPLC and to implement a dynamic cooling profile to maintain supersaturation at a desired level in closed-loop.

6) A systematic study of batch versus continuous crystallizations (i.e. MSMPRC, PFC) can be carried out from both technical and economic points of view to achieve some general recommendations on the selection of the most efficient and economical crystallization process. For example, one can anticipate that MSMPRC and PFC are not good options for compounds that have very slow growth kinetics due to the requirement of long residence times and large operating volume. Batch crystallization, on the other hand, might be able to achieve large crystals even in small operating volume by simply increasing the batch time. In addition, it is more cost effective to use continuous crystallization for API that is of high demand, and vice versa, from the economic point of view.

8

REFERENCES

183

REFERENCES

Aamir, E., Nagy, Z.K., Rielly, C.D., 2010. Optimal seed recipe design for crystal size distribution control for batch cooling crystallisation processes. Chem. Eng. Sci. 2010, 65,

3602–3614.

Abbou Oucherif, K., Raina, S., Taylor, L.S., Litster, J.D., 2013. Quantitative analysis of the inhibitory effect of HPMC on felodipine crystallization kinetics using population balance modeling. CrystEngCommun 15, 2197–2205.

Abu Bakar, M.R., Nagy, Z.K., Saleemi, A.N., Rielly, C.D., 2009. The impact of direct nucleation control on crystal size distribution in pharmaceutical crystallization processes.

Cryst. Growth Des. 9, 1378–1384.

Acevedo, D., Nagy, Z.K., 2014. Systematic classification of unseeded batch crystallization systems for achievable shape and size analysis. J. Cryst. Growth 394, 97–105.

Acevedo, D., Peña, R., Yang, Y., Barton, A., Firth, P., Nagy, Z.K., 2016. Evaluation of mixed suspension mixed product removal crystallization processes coupled with a continuous filtration system. Chem. Eng. Process. Process Intensif. 108, 212–219.

184

Acevedo, D., Tandy, Y., Nagy, Z.K., 2015. Multiobjective optimization of an unseeded batch cooling crystallizer for shape and size manipulation. Ind. Eng. Chem. Res. 54, 2156–

2166.

Allen, R.J., Valeriani, C., Tanase-Nicola, S., Wolde, P.R., Frenkei, D., 2008.

Homogeneous nucleation under shear in a two-dimensional Ising model: cluster growth, coalescence, and breakup. J. Chem. Phys. 129, 134704.

Alvarez, A.J., Myerson, A.S., 2010. Continuous plug flow crystallization of pharmaceutical compounds. Cryst. Growth Des. 10, 2219–2228.

Alvarez, A.J., Singh, A., Myerson, A.S., 2011. Crystallization of cyclosporine in a multistage continuous MSMPR crystallizer. Cryst. Growth Des. 11, 4392–4400.

Bakar, S.K., Niazi, S., 1983. Stability of aspirin in different media. J. Pharm. Sci. 72, 1024–

1026.

Barba, L., Arrighetti, G., Calligaris, S., 2013. Crystallization and melting properties of extra virgin olive oil studied by synchrotron XRD and DSC. Eur. J. Lipid Sci. Technol.

115, 322–329.

Barrett, M., McNamara, M., Hao, H., Barrett, P., Glennon, B., 2010. Supersaturation tracking for the development, optimization and control of crystallization processes. Chem.

Eng. Res. Des. 88, 1108–1119.

Barrett, P., Glennon, B., 2002. Characterizing the metastable zone width and solubility curve using lasentec FBRM and PVM. Chem. Eng. Res. Des. 80, 799–805.

185

Barthe, S.C., Grover, M.A., Rousseau, R.W., 2008. Observation of polymorphic change through analysis of FBRM data: transformation of paracetamol from form II to form I.

Cryst. Growth Des. 8, 3316–3322.

Borsos, A., Majumder, A., Nagy, Z.K., 2016. Multi-impurity adsorption model for modeling crystal purity and shape evolution during crystallization processes in impure media. Cryst. Growth Des. 16, 555–568.

Braatz, R.D., 2002. Advanced control of crystallization processes. Annu. Rev. Control 26,

87–89.

Burcham, C.L., Collins, P.C., Jarmer, D.J., Seibert, K.D., 2008. Formulation and analytical development for low-dose oral drug products; Zheng, J., Eds.; John Wiley & Sons Inc.:

Hoboken, NJ; Chapter 8, pp 205–222.

Chen, J., Sarma, B., Evans, J.M.B., Myerson, A.S., 2011. Pharmaceutical crystallization.

Cryst. Growth Des. 11, 887–895.

Chew, J.W., Black, S.N., Chow, P.S., Tan, R.B.H., 2007. Comparison between open-loop temperature control and closed-loop supersaturation control for cooling crystallization of glycine. Ind. Eng. Chem. Res. 46, 830–838.

Cornel, J., Lindenberg, C., Mazzotti, M., 2008. Quantitative application of in situ ATR-

FTIR and Raman spectroscopy in crystallization processes. Ind. Eng. Chem. Res. 47,

4870–4882.

186

Davey, R.J., Garside, J., 2000. From molecules to crystallizers: an introduction to crystallization, first ed., Oxford university press.

Eder, R.J., Radl, S., Schmitt, E., Innerhofer, S., Maier, M., Gruber-Woelfler, H., Khinast,

J.G., 2010. Continuously seeded, continuously operated tubular crystallizer for the production of active pharmaceutical ingredients. Cryst. Growth Des. 10, 2247–2257.

Engstrom, J., Wang, C., Lai, C., Sweeney, J., 2013. Introduction of a new scaling approach for particle size reduction in toothed rotor-stator wet mills. Int. J Pharm. 456, 261–268.

Ferguson, S., Morris, G., Hao, H., Barrett, M., Glennon, B., 2012. In-situ monitoring and characterization of plug flow crystallizers. Chem. Eng. Sci. 77, 105–111.

Frawley, P.J., Mitchell, N.A., Ó’Ciardhá, C.T., Hutton, K.W., 2012. The effects of supersaturation, temperature, agitation and seed surface area on the secondary nucleation of paracetamol in ethanol solutions. Chem. Eng. Sci. 75, 183–197.

Fujiwara, M., Nagy, Z.K., Chew, J.W., Braatz, R.D., 2005. First-principles and direct design approaches for the control of pharmaceutical crystallization. J. Process Control 15,

493–504.

Gray, R.A., Warren, P.B., Chynoweth, S., Michopoulos, Y., Pawley, G.S., 1995.

Crystallization of molecular liquids through shear-induced nucleation. Proc. R. Soc. Lond.

A 448, 113–120.

187

Gron, H., Borissova, A., Roberts, K.J., 2003. ATR-FTIR spectroscopy for closed-loop supersaturation control of a batch crystallizer producing monosodium glutamate crystals of defined size. Ind. Eng. Chem. Res. 42, 198–206.

Hacherl, J.M., Paul, E.L., Buettner, H.M., 2003. Investigation of impinging-jet crystallization with a calcium oxalate model system. AIChE J. 49, 2352–2362.

Harter, A., Schenck, L., Lee, I., Cote, A., 2013. High-shear rotor-stator wet milling for drug substances: expanding capability with improved scalability. Org. Process Res. Dev.

17, 1335–1344.

Hassell, D.C., Bowman, E.M., 1998. Process analytical chemistry for spectroscopists. Appl.

Spectrosc. 52, 18A–29A.

Hermanto, M.W., Chiu, M., Braatz, R.D., 2009. Nonlinear model predictive control for the polymorphic transformation of L-Glutamic acid crystals. AICHE J. 55, 2631–2645.

Hermanto, M.W., Chow, P.S., Tan, R.B.H., 2010. Implementation of focused beam reflectance measurement (FBRM) in antisolvent crystallization to achieve consistent product quality. Cryst. Growth Des. 10, 3668–3674.

Hermanto, M.W., Chow, P.S., Tan, R.B.H., 2012. Operating strategy to produce consistent

CSD in combined antisolvent-cooling crystallization using FBRM. Ind. Eng. Chem. Res.

51, 13773−13783.

188

Hou, G., Power, G., Barrett, M., Glennon, B., Morris, G., Zhao, Y., 2014. Development and characterization of a single stage mixed-suspension, mixed-product-removal crystallization process with a novel transfer unit. Cryst. Growth Des. 14, 1782–1793.

Johnson, B.K., Prud’homme, R.K., 2003. Chemical processing and micromixing in confined impinging jets. AIChE J. 49, 2264–2282.

Kalbasenka, A.N., Spierings, L.C.P., Huesman, A.E.M., Kramer, H.J.M., 2007.

Application of seeding as a process actuator in a model predictive control framework for fed-batch crystallization of ammonium sulphate. Part. Part. Syst. Charact. 24, 40–48.

Kamahara, T., Takasuga, M., Tung, H.H., Hanaki, K., Fukunaka, T., Izzo, B., Nakada, J.,

Yabuki, Y., Kato, Y., 2007. Generation of fine pharmaceutical particles via controlled secondary nucleation under high shear environment during crystallization: process development and scale-up. Org. Process Res. Dev. 11, 699–703.

Kempkes, M., Eggers, J., Mazzotti, M., 2008. Measurement of particle size and shape by

FBRM and in situ microscopy. Chem. Eng. Sci. 63, 4656–4675.

Khan, S., Ma, C.Y., Mahmud, T., Penchev, R.Y., Roberts, K.J., Morris, J., Özkan, L.,

White, G., Grieve, B., Hall, A., Buser, P., Gibson, N., Keller, P., Shuttleworth, P., Price,

C.J., 2011. In-process monitoring and control of supersaturation in seeded batch cooling crystallization of L-glutamic acid: from laboratory to industrial pilot plant. Org. Process

Res. Dev. 15, 540–555.

189

King, J.C., Li, H., Grover, M.A., Kawajiri, Y., Rousseau, R.W., 2015. Optimization of two- stage cooling profile in unseeded batch crystallization. IFAC-PapersOnLine 2015, 48,

297–302.

Kougoulos, E., Smales, I., Verrier, H.M., 2011. Towards integrated drug substance and drug product design for an active pharmaceutical ingredient using particle engineering.

AAPS PharmSciTech 12, 287–294.

Lawton, S., Steele, G., Shering, P., Zhao, L., Laird, I., Ni, X.W., 2009. Continuous crystallization of pharmaceuticals using a continuous oscillatory baffled crystallizer. Org.

Process Res. Dev. 13, 1357–1363.

Leba, H., Cameirao, A., Herri, J., Darbouret, M., Peytavy, J., Glenat, P., 2010. Chord length distribution measurements during crystallization and agglomeration of gas hydrate in a water-in-oil emulsion: simulation and experimentation. Chem. Eng. Sci. 65, 1185–1200.

Li, H., Grover, M.A., Kawajiri, Y., Rousseau, R.W., 2013. Development of an empirical method relating crystal size distributions and FBRM measurements. Chem. Eng. Sci. 89,

142–151.

Li, H., Kawajiri, Y., Grover, M.A., Rousseau, R.W., 2014. Application of an empirical

FBRM model to estimate crystal size distributions in batch crystallization. Cryst. Growth

Des. 14, 607–616.

190

Li, J., Trout, B.L., Myerson, A.S., 2016. Multistage continuous mixed-suspension, mixed- product removal (MSMPR) crystallization with solids recycle. Org. Process Res. Dev. 20,

510–516.

Lindenberg, C., Krattli, M., Cornel, J., Mazzotti, M., 2009. Design and optimization of a combined cooling/antisolvent crystallization process. Cryst. Growth Des. 9, 1124–1136.

Lo, E., Mattas, E., Wei, C., Kacsur, D., Chen, C., 2012. Simultaneous API particle size reduction and polymorph transformation using high shear. Org. Process Res. Dev. 16, 102–

108.

Ma, C.Y., Wang, X.Z., 2012. Closed-loop control of crystal shape in cooling crystallization of L-glutamic acid. J. Process Control 22, 72–81.

McGlone, T., Briggs, N.E.B., Clark, C.A., Brown, C.J., Sefcik, J., Florence, A.J., 2015.

Oscillatory flow reactors (OFRs) for continuous manufacturing and crystallization. Org.

Process Res. Dev. 19, 1186–1202.

Mesbah, A., Huesman, A.E.M., Kramer, H.J.M., Nagy, Z.K., Van den Hof, P.M.J., 2011.

Real-time control of seeded batch crystallization processes. AIChE J. 57, 1557–1569.

Mesbah, A., Nagy, Z.K., Huesman, A.E.M., Kramer, H.J.M., Van den Hof, P.M.J., 2012.

Real-time control of industrial batch crystallization processes using a population balance modeling framework. IEEE Transaction on Control Systems Technologies 20, 1188–1201.

Mitchell, N.A., Frawley, P.J., 2010. Nucleation kinetics of paracetamol-ethanol solutions from metastable zone widths. J. Cryst. Growth 312, 2740–2746.

191

Mullin, J.W., 2001. Crystallization, fourth edition, Butterworth Heinemann.

Mullin, J.W., Nyvlt, J., 1971. Programmed cooling of batch crystallizers. Chem. Eng. Sci.

26, 369–377.

Nagy, Z.K., Aamir, E., 2012. Systematic design of supersaturation controlled crystallization processes for shaping the crystal size distribution using an analytical estimator. Chem. Eng. Sci. 84, 656–670.

Nagy, Z.K., Aamir, E., Rielly, C.D., 2011. Internal fines removal using population balance model based control of crystal size distribution under dissolution, growth and nucleation mechanisms. Cryst. Growth Des. 11, 2205–2219.

Nagy, Z.K., Braatz, R.D., 2003. Robust nonlinear model predictive control of batch processes. AICHE J. 49, 1776–1786.

Nagy, Z.K., Braatz, R.D., 2012. Advances and new directions in crystallization control. Annu. Rev. Chem. Biomol. Eng. 3, 55–75.

Nagy, Z.K., Chew, J.W., Fujiwara, M., Braatz, R.D., 2008a. Comparative performance of concentration and temperature controlled crystallizations. J. Process Control 18, 399–407.

Nagy, Z.K., Fevotte, G., Kramer, H., Simon, L.L., 2013. Recent advances in the monitoring, modelling and control of crystallization systems. Chem. Eng. Res. Des. 91,

1903–1922.

192

Nagy, Z.K., Fujiwara, M., Braatz, R.D., 2008b. Modelling and control of combined cooling and antisolvent crystallization processes. J. Process Control 18, 856–864.

Nagy, Z.K., Mahn, B., Franke, R., Allgower, F., 2007. Efficient output feedback nonlinear model predictive control for temperature control of industrial batch reactors. Control Eng.

Pract. 15, 839–859.

Nanubolu, J.B., Burley, J.C., 2012. Investigating the recrystallization behavior of amorphous paracetamol by variable temperature Raman studies and surface Raman mapping. Mol. Pharmaceutics 9, 1544–1558.

Nowee, S.M., Abbas, A., Romagnoli, J.A., 2008. Antisolvent crystallization: model identification, experimental validation and dynamic simulation. Chem. Eng. Sci. 63, 5457–

5467.

Park, K., Kim, D.Y., Yang, D.R., 2016. Operating strategy for continuous multistage mixed suspension and mixed product removal (MSMPR) crystallization processes depending on crystallization kinetic parameters. Ind. Eng. Chem. Res. 55, 7142–7153.

Peña, R., Nagy, Z.K., 2015. Process intensification through continuous spherical crystallization using a two-stage mixed suspension mixed product removal (MSMPR) system. Cryst. Growth Des. 15, 4225–4236.

Pickering, M., Brown, S., 2012. Quantification and validation of HPLC-UV and LC-MS assays for therapeutic drug monitoring of ertapenem in human plasma. Biomed.

Chromatogr. 27, 568–574.

193

Power, G., Hou, G., Kamaraju, V.K., Morris, G., Zhao, Y., Glennon, B., 2015. Design and optimization of a multistage continuous mixed suspension, mixed product removal crystallizer. Chem. Eng. Sci. 133, 125–139.

Qamar, S., Mukhtar, S., Seidel-Morgenstern, A., 2010. Efficient solution of a batch crystallization model with fines dissolution. J. Cryst. Growth 312, 2936–2945.

Quon, J.L., Zhang, H., Alvarez, A., Evans, J., Myerson, A.S., Trout, B.L., 2012.

Continuous crystallization of aliskiren hemifumarate. Cryst. Growth Des. 12, 3036–3044.

Ramkrishna, D., 2000. Population balances: theory and applications to particulate systems in engineering. Academic Press, San Diego.

Randolph, A.L., 1988, Theory of particulate processes. Academic Press, San Diego, second edition.

Rawlings, J.B., Miller, S.M., Witkowski, W.R., 1993. Model identification and control of solution crystallization processes: a review. Ind. Eng. Chem. Res. 32, 1275–1296.

Reguera, D., Rubi, J.M., 2003. Homogeneous nucleation in inhomogeneous media. II.

Nucleation in a shear flow. J. Chem. Phys. 119, 9888–9893.

Saleemi, A.N., Rielly, C.D., Nagy, Z.K., 2012a. Automated direct nucleation control for in situ dynamic fines removal in batch cooling crystallization. CrystEngComm 14, 2196-

2203.

194

Saleemi, A.N., Rielly, C.D., Nagy, Z.K., 2012b. Comparative investigation of supersaturation and automated dicrect nucleation control of crystal size distributions using

ATR-UV/vis spectroscopy and FBRM. Cryst. Growth Des. 12, 1792–1807.

Sandwell, D.T., 1987. Biharmonic spline interpolation of GEOS-3 and SEASAT altimeter data. Geophys. Res. Lett. 14, 139–142.

Sheikhzadeh, M., Trifkovic, M., Rohani, S., 2008. Adaptive MIMO neuro-fuzzy logic control of a seeded and an unseeded anti-solvent semi-batch crystallizer. Chem. Eng. Sci.

63, 1261–1272.

Simon, L.L., Merz, T., Dubuis, S., Lieb, A., Hungerbuhler, K., 2012. In-situ monitoring of pharmaceutical and specialty chemicals crystallization processes using endoscopy- stroboscopy and multivariate image analysis. Chem. Eng. Res. Des. 90, 1847–1855.

Simon, L.L., Nagy, Z.K., Hungerbuhler, K., 2009a. Model based control of a liquid swelling constrained batch reactor subject to recipe uncertainties. Chem. Eng. J. 153, 151–

158.

Simon, L.L., Nagy, Z.K., Hungerbuhler, K., 2009b. Endoscopy-based in situ bulk video imaging of batch crystallization processes. Org. Process Res. Dev. 13, 1254–1261.

195

Simon, L.L., Pataki, H., Marosi, G., Meemken, F., Hungerbühler, K., Baiker, A., Tummala,

S., Glennon, B., Kuentz, M., Steele, G., Kramer, H.J.M., Rydzak, J.W., Chen, Z., Morris,

J., Kjell, F., Singh, R., Gani, R., Gernaey, K.V., Louhi-kultanen, M., O’Reilly, J., Sandler,

N., Antikainen, O., Yliruusi, J., Frohberg, P., Ulrich, J., Braatz, R.D., Leyssens, T., Stosch,

M.V., Oliveira, R., Tan, R.B.H., Wu, H., Khan, M., O’Grady, D., Pandey, A., Westra, R.,

Delle-case, E., Pape, D., Angelosante, D., Maret, Y., Steiger, O., Lenner, M., Abbou-

Oucherif, K., Nagy, Z.K., Litster, J.D., Kamaraju, V.K., Chiu, M., 2015. Assessment of recent process analytical technology (PAT) trends: a multiauthor review. Org. Process Res.

Dev. 19, 3–62.

Simone, E., Nagy, Z.K., 2015a. A link between the ATR-UV/Vis and Raman spectra of zwitterionic solutions and the polymorphic outcome in cooling crystallization.

CrystEngComm 17, 6538–6547.

Simone, E., Saleemi, A.N., Nagy, Z.K., 2014a. Application of quantitative Raman spectroscopy for the monitoring of polymorphic transformation in crystallization processes using a good calibration practice procedure. Chem. Eng. Res. Des. 92, 594–611.

Simone, E., Saleemi, A.N., Nagy, Z.K., 2014b. Raman, UV, NIR, and mid-IR spectroscopy with focused beam reflectance measurement in monitoring polymorphic transformations.

Chem. Eng. Tech. 8, 1305–1313.

196

Simone, E., Saleemi, A.N., Nagy, Z.K., 2015b. In situ monitoring of polymorphic transformations using a composite sensor array of Raman, NIR, and ATR-UV/vis spectroscopy, FBRM, and PVM for an intelligent decision support system. Org. Process

Res. Dev. 19, 167–177.

Simone, E., Saleemi, A.N., Tonnon, N., Nagy, Z.K., 2014c. Active polymorphic feedback control of crystallization processes using a combined Raman and ATR-UV/Vis spectroscopy approach. Cryst. Growth Des. 14, 1839–1850.

Simone, E., Zhang, W., Nagy, Z. K., 2015c. Application of process analytical technology- based feedback control strategies to improve purity and size distribution in biopharmaceutical crystallization. Cryst. Growth Des. 15, 2908–2919.

Steen, L., Rigolle, A., Glorieux, S., Paelinck, H., Fraeye, I., Goderis, B., Foubert, I., 2015.

Isothermal crystallization behavior of lard at different temperatures studied by DSC and real-time XRD. Food Res. Int. 69, 49–56.

Su, Q., Nagy, Z.K., Rielly, C.D., 2015. Pharmaceutical crystallization processes from batch to continuous operation using MSMPR stages: modelling, design, and control. Chem. Eng.

Process. Process Intensif. 89, 41–53.

Su, W., Hao, H., Barrett, M., Glennon, B., 2010. The impact of operating parameters on the polymorphic form transformation of D-mannitol characterized in situ with raman spectroscopy, FBRM, and PVM. Org. Process Res. Dev. 14, 1432–1437.

197

Swartz, M.E., 2005. UPLCTM: an introduction and review. J. Liq. Chromatogr. Relat.

Technol. 28, 1253–1263.

Thompson, D.R., Kougoulos, E., Jones, A.G., Wood-Kaczmar, M.W., 2005. Solute concentration measurement of an important organic compound using ATR-UV spectroscopy. J. Cryst. Growth 276, 230–236.

Thorley, F.C., Baldwin, K.J., Lee, D.C., Batchelder, D.N., 2006. Dependence of the Raman spectra of drug substances upon laser excitation wavelength. J. Raman Spectrosc. 37, 335–

341.

Trifkovic, M., Sheikhzadeh, M., Rohani, S., 2009. Multivariable real-time optimal control of a cooling and antisolvent semibatch crystallization process. AIChE J. 55, 2591–2602.

U.S. Food and Drug Administration, 2004. Guidance for industry PAT - A framework for innovative pharmaceutical development, manufacturing and quality assurance. FDA:

Rockville, MD.

Van Eerdenbrugh, B., Alonzo, D.E., Taylor, L.S., 2011. Influence of particle size on the ultraviolet spectrum of particle-containing solutions: implications for in-situ concentration monitoring using UV/Vis fiber-optic probes. Pharm. Res. 28, 1643–1652.

Vetter, T., Burcham, C.L., Doherty, M.F., 2014. Regions of attainable particle size in continuous and batch crystallization processes. Chem. Eng. Sci. 106, 167–180.

Ward, J.D., Mellichamp, D.A., Doherty, M.F., 2006. Choosing an operating policy for seeded batch crystallization. AIChE J. 52, 2046–2054.

198

Wong, S.Y., Tatusko, A.P., Trout, B.L., Myerson, A.S., 2012. Development of continuous crystallization processes using a single-stage mixed-suspension, mixed-product removal crystallizer with recycle. Cryst. Growth Des. 12, 5701–5707.

Woo, X.Y., Nagy, Z.K., Tan, R.B.H., Braatz, R.D., 2009. Adaptive concentration control of cooling and antisolvent crystallization with laser backscattering measurement. Cryst.

Growth Des. 9, 182–191.

Worlitschek, J., Mazzotti, M., 2004. Model-based optimization of particle size distribution in batch-cooling crystallization of paracetamol. Cryst. Growth Des. 4, 891–903.

Wu, D., Bird, M.R., 2007. The fouling and cleaning of membranes during the filtration of model tea component solutions. J. Food Process Eng. 30, 293–323.

Xie, W., Rohani, S., Phoenix A., 2001. Dynamic modelling and operation of a seeded batch cooling crystalliser. Chem. Eng. Commun. 187, 229–249.

Yang, Y., Nagy, Z.K., 2014. Model-based systematic design and analysis approach for unseeded combined cooling and antisolvent crystallization (CCAC) systems. Cryst.

Growth Des. 14, 687–698.

Yang, Y., Nagy, Z.K., 2015a. Advanced control approaches for combined cooling/antisolvent crystallization in continuous mixed suspension mixed product removal cascade crystallizers. Chem. Eng. Sci. 127, 362–373.

199

Yang, Y., Nagy, Z.K., 2015b. Application of nonlinear model predictive control in continuous crystallization systems. Proceedings of the American Control Conference,

4282–4287.

Yang, Y., Nagy, Z.K., 2015c. Combined cooling and antisolvent crystallization in continuous mixed suspension, mixed product removal cascade crystallizers: steady-state and startup optimization. Ind. Eng. Chem. Res. 54, 5673–5682.

Yang, Y., Song, L., Gao, T., Nagy, Z.K., 2015d. Integrated upstream and downstream application of wet milling with continuous mixed suspension mixed product removal crystallization. Cryst. Growth Des. 15, 5879–5885.

Yang, Y., Song, L., Nagy, Z.K., 2015e. Automated direct nucleation control in continuous mixed suspension mixed product removal cooling crystallization. Cryst. Growth Des. 15,

5839–5848.

Yang, Y., Song, L., Zhang, Y., Nagy, Z.K., 2016. Application of wet milling-based automated direct nucleation control in continuous cooling crystallization processes. Ind.

Eng. Chem. Res. 55, 4987−4996.

Yu, Z.Q., Chew, J.W., Chow, P.S., Tan, R.B.H., 2007. Recent advances in crystallization control - an industrial perspective. Chem. Eng. Res. Des. 85, 893–905.

Zhang, G.P., Rohani, S., 2003. On-line optimal control of a seeded batch cooling crystallizer. Chem. Eng. Sci. 58, 1887–1896.

200

Zhang, H., Quon, J., Alvarez, A.J., Evans, J., Myerson, A.S., Trout, B., 2012. Development of continuous anti-solvent/cooling crystallization process using cascaded mixed suspension, mixed product removal crystallizers. Org. Process Res. Dev. 16, 915–924.

Zhang, Y., Jiang, Y., Zhang, D., Qian, Y., Wang, X.Z., 2015. Metastable zone width, crystal nucleation and growth kinetics measurement in anti-solvent crystallization of β- artemether in the mixture of ethanol and water. Chem. Eng. Res. Des. 95, 187–194.

Zhao, J., Wang, B., Dong, B., Feng, Q., Xu, C., 2013. Monitoring the polymorphic transformation of imidacloprid using in situ FBRM and PVM. Org. Process Res. Dev. 17,

375–381.

Zhou, G.X., Fujiwara, M., Woo, X.Y., Rusli, E., Tung, H., Starbuck, C., Davidson, O., Ge,

Z., Braatz, R.D, 2006. Direct design of pharmaceutical antisolvent crystallization through concentration control. Cryst. Growth Des. 6, 892–898.

13

VITA

201

VITA

EDUCATION

 Doctor of Philosophy, December 2016

Purdue University, West Lafayette, IN

Major in

 Bachelor of Science, July 2012

Peking University, Beijing, China

Major in Chemistry

INDUSTRIAL EXPERIENCE

 Science intern, AbbVie, North Chicago, IL, May 2014 – August 2014

TEACHING EXPERIENCE

 Teaching assistant for Statistical Modeling and Quality Enhancement, Spring 2015

 Teaching assistant for Process Dynamics and Control, Fall 2013

CERTIFICATION & AWARDS

 Certified LABVIEW Associate Developer (CLAD), National Instruments, 05/18/2015

– 05/17/2017

202

 Eastman Travel Award, Purdue University, 2014

 GAW Scholarship, Purdue University, 2012

ORAL AND POSTER PRESENTATIONS

 Yang, Y., Nagy, Z.K. "Monitoring and Control of Pharmaceutical Crystallization

Systems in the Presence of Impurities Using Online UPLC Technology", 2nd Annual

Process Analytical Technologies Symposium, Indianapolis, IN, Sep. 2016. (Invited talk)

 Yang, Y., Pal, K., Peña, R., Nagy, Z.K. "Application of Ultra-Performance Liquid

Chromatography (UPLC) as an Online Process Analytical Technology (PAT) in

Pharmaceutical Crystallization & Spherical Agglomeration of Active Pharmaceutical

Ingredient", 25th Annual Purdue Graduate Research Symposium, West Lafayette, IN,

Aug. 2016. (Poster)

 Yang, Y., Song, L., Nagy, Z.K. "Application of Automated Direct Nucleation Control

in Continuous Cooling Crystallization and Wet Milling Processes", 8th Annual AIChE

Midwest Regional Conference, Chicago, IL, Mar. 2016. (Oral)

 Yang, Y., Song, L., Nagy, Z.K. "Automated Multi-Loop Direct Nucleation Control

Using FBRM in Multistage Continuous Mixed Suspension Mixed Product Removal

Crystallizer", 2015 AIChE Annual Meeting, Salt Lake City, UT, Nov. 2015. (Oral)

 Yang, Y., Song, L., Nagy, Z.K. "Application of FBRM-Based Feedback Control in

Wet Milling Integrated Continuous Cooling Crystallization", 24th Annual Purdue

Graduate Research Symposium, West Lafayette, IN, Aug. 2015. (Oral)

203

 Yang, Y., Acevedo, D., Nagy, Z.K. "Design and Control of Continuous Crystallization,

Wet Milling and Filtration Units", 24th Annual Purdue Graduate Research Symposium,

West Lafayette, IN, Aug. 2015. (Poster)

 Yang, Y., Nagy, Z.K. "Application of Nonlinear Model Predictive Control in

Continuous Crystallization Systems", 2015 American Control Conference, Chicago, IL,

Jul. 2015. (Oral)

 Yang, Y., Nagy, Z.K. "Advanced Control Approaches for Combined

Cooling/Antisolvent Crystallization in Continuous Mixed Suspension Mixed Product

Removal Cascade Crystallizers", 7th Annual AIChE Midwest Regional Conference,

Chicago, IL, Mar. 2015. (Oral)

 Yang, Y., Nagy, Z.K. "Steady-state and Startup Optimization for Combined

Cooling/Antisolvent Crystallization in Continuous Mixed Suspension Mixed Product

Removal Cascade Crystallizers", 2014 AIChE Annual Meeting, Atlanta, GA, Nov.

2014. (Oral)

 Yang, Y., Nagy, Z.K. "Advanced Control Approaches for Combined Cooling and

Antisolvent Crystallization (CCAC) in Continuous MSMPR Cascade Crystallizers",

2014 AIChE Annual Meeting, Atlanta, GA, Nov. 2014. (Poster)

 Yang, Y., Nagy, Z.K. "Advanced Control Approaches for Combined

Cooling/Antisolvent Crystallization in Continuous Crystallizers", Purdue Industrial

Advisory Council Meeting, West Lafayette, IN, Sep. 2014. (Poster)

204

 Yang, Y., Nagy, Z.K. "Modeling and Control of Combined Cooling/Antisolvent

Crystallization in Mixed Suspension Mixed Product Removal (MSMPR) Cascade

Crystallizers", 23th Annual Purdue Graduate Research Symposium, West Lafayette, IN,

Aug. 2014. (Oral)

 Yang, Y., Nagy, Z.K. "Combined Cooling and Antisolvent Crystallization (CCAC) in

Continuous MSMPR Cascade Crystallizers", Eli Lilly/Purdue Tech Day, West

Lafayette, IN, May 2014. (Poster)

 Yang, Y., Nagy, Z.K. "Startup Optimization of Continuous MSMPR Cascade

Crystallizers Using a Population Balance Model", Computational Science and

Engineering Student Conference (CSESC), West Lafayette, IN, Apr. 2014. (Oral)

 Yang, Y., Nagy, Z.K. "Model-based Systematic Design and Analysis Approach for

Unseeded Combined Cooling and Antisolvent Crystallization (CCAC) Systems",

Purdue Industrial Advisory Council Meeting, West Lafayette, IN, Sep. 2013. (Poster)

 Yang, Y., Nagy, Z.K. "A Novel Experimental Design Methodology Through Model-

Based Simulation for Combined Cooling and Antisolvent Crystallization", 22th Annual

Purdue Graduate Research Symposium, West Lafayette, IN, Aug. 2013. (Poster)

PROFESSIONAL ACTIVITIES

 Reviewer for Ind. Eng. Chem. Res.

 Reviewer for Chem. Eng. Res. Des.

 Reviewer for Org. Process Res. Dev.

 Reviewer for Chem. Eng. Process. Process Intensif.

205

 Reviewer for J. Chem. Thermodyn.

PUBLICATIONS

1. Yang, Y., Nagy, Z.K., 2014. Model-based systematic design and analysis approach for

unseeded combined cooling and antisolvent crystallization (CCAC) systems. Cryst.

Growth Des. 14, 687–698. Chapter 3.

2. Yang, Y., Nagy, Z.K., 2015. Combined cooling and antisolvent crystallization in

continuous mixed suspension, mixed product removal cascade crystallizers: steady-

state and startup optimization. Ind. Eng. Chem. Res. 54, 5673–5682. Chapter 4.

3. Yang, Y., Nagy, Z.K., 2015. Advanced control approaches for combined

cooling/antisolvent crystallization in continuous mixed suspension mixed product

removal cascade crystallizers. Chem. Eng. Sci. 127, 362–373. Chapter 5.

4. Yang, Y., Song, L., Gao, T., Nagy, Z.K., 2015. Integrated upstream and downstream

application of wet milling with continuous mixed suspension mixed product removal

crystallization. Cryst. Growth Des. 15, 5879–5885. Chapter 6.

5. Yang, Y., Song, L., Nagy, Z.K., 2015. Automated direct nucleation control in

continuous mixed suspension mixed product removal cooling crystallization. Cryst.

Growth Des. 15, 5839–5848. Chapter 7.

6. Yang, Y., Song, L., Zhang, Y., Nagy, Z.K., 2016. Application of wet milling-based

automated direct nucleation control in continuous cooling crystallization processes. Ind.

Eng. Chem. Res. 55, 4987–4996. Chapter 8.

206

7. Yang, Y., Zhang, C., Pal, K., Koswara, A., Quon, J., McKeown, R., Goss, C., Nagy,

Z.K. Application of ultra-performance liquid chromatography as an online process

analytical technology tool in pharmaceutical crystallization. Cryst. Growth Des.

Submitted. Chapter 9.

8. Yang, Y., Pal, K., Koswara, A., Sun, Q., Zhang, Y., Quon, J., McKeown, R., Goss, C.,

Nagy, Z.K. Application of feedback control and in situ milling to improve particle size

and shape in the crystallization of a slow growing needle-like active pharmaceutical

ingredient. Int. J. Pharm. Submitted.

9. Yang, Y., Nagy, Z.K., 2015. Application of nonlinear model predictive control in

continuous crystallization systems. Proceedings of the American Control Conference,

4282–4287.

10. Acevedo, D., Peña, R., Yang, Y., Barton, A., Firth, P., Nagy, Z.K., 2016. Evaluation

of mixed suspension mixed product removal crystallization processes coupled with a

continuous filtration system. Chem. Eng. Process. Process Intensif. 108, 212–219.