ogrady2007-multi-scale characterization of anti-solvent crystallization by

189
Page 1 Multi-Scale Characterization of Anti-solvent Crystallization by Des O‟Grady, B.E. A Thesis presented to the National University of Ireland in fulfilment of the requirements for the degree of Doctor of Philosophy (Ph.D.) Under the supervision of Dr. Brian Glennon December 2007 School of Chemical and Bioprocess Engineering, College of Engineering, Mathematical and Physical Sciences, University College Dublin.

Upload: chaitanya-koduru

Post on 30-Dec-2015

101 views

Category:

Documents


5 download

DESCRIPTION

It's a research paper related to the anti solvent crystallization system.

TRANSCRIPT

Page 1: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

Page 1

Multi-Scale Characterization of Anti-solvent

Crystallization

by

Des O‟Grady, B.E.

A Thesis presented to the National University of Ireland in fulfilment of the

requirements for the degree of Doctor of Philosophy (Ph.D.)

Under the supervision of

Dr. Brian Glennon

December 2007

School of Chemical and Bioprocess Engineering,

College of Engineering, Mathematical and Physical Sciences,

University College Dublin.

Page 2: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

ACKNOWLEDGEMENTS

First and foremost I would like to thank my parents Dorothy and Dermot. Their love

and support over the years has been unstinting and in one way or another they have

taught me everything I know. For that I will always be grateful.

Thanks to my wonderful sisters, Caoilfhionn, Fiona and Elva, for constantly

endeavouring to figure out what it was that I actually did during my PhD. They‟re

inquisitive nature has hopefully rubbed off on me over the years.

I would like to thank Dr. Brian Glennon for his talented and dedicated guidance. He

has an uncanny ability to get to the heart of a problem and convey the solution in a

clear and meaningful way. My PhD studies were made possible and have been a

valuable and extremely enjoyable experience thanks to him.

I would like to thank my extended family of aunts, uncles and cousins. A special

thanks to Joe and Teresa for taking such a keen interest in my studies over the years.

Countless friends have made my PhD studies all the more enjoyable. They offered a

unique brand of humour that provided the perfect distraction from the lab. I have

ninety-two reasons to thank Skellig and Marko along with Bud, Miller and Mac. Not

to concentrate on that small group I have to also thank many others from college and

school including Ed, Hutchy, Dots, Robbie, Ale, Karol, Marsha, Nikki, and Sarah.

Page 3: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

Thanks to my fellow postgrads for making UCD a fun place to be. Special thanks to

Ale, Brian and Paul who have been were there from the start. Thanks to an influx of

new faces the last couple of years have been very enjoyable. Thanks to Eoin, Jessica,

Kate, Aisling, Barry, and John for starting it off. Thanks to Mark for all his help and

the CFD work in this thesis. Mairtin, good luck!

Thanks to all the lecturers in the School of Chemical and Bioprocess Engineering. A

special thanks to Dermot Malone, Patricia Kieran, Eoin Casey and Frank

MacLoughlin for all their help.

Thanks to everyone else in the school who helped me out along the way, including,

Pat, Liam, Oliver, Jim, Brid, Sinead, Pat and Brian. A special thanks to Aoife.

Thanks to everyone at Mettler Toledo for giving me a great opportunity during my

PhD. A special thanks to Paul Barrett for opening that door.

Finally, I would like to thank Anne for her brilliant friendship and unwavering loyalty

and love. Here‟s to NYC Anneski!!!

Page 4: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

TABLE OF CONTENTS

LIST OF PUBLICATIONS 2

ABSTRACT 1

CHAPTER 1: INTRODUCTION 2

CHAPTER 2: LITERATURE REVIEW 8

2.1 ANTI-SOLVENT CRYSTALLIZATION 8

2.1.1 INTRODUCTION 8

2.1.2 VARIABLES TO CONSIDER IN ANTI-SOLVENT CRYSTALLIZATION 9

2.1.2.1 Choice of Anti-solvent 9

2.1.2.2 Anti-solvent Addition Rate 11

2.1.2.3 Agitation Intensity 13

2.1.2.4 Concentration Effects 14

2.1.2.5 Conclusion 17

2.2 FOCUSSED BEAM REFLECTANCE MEASUREMENT (FBRM) 19 2.2.1 INTRODUCTION 19

2.2.2 MODE OF OPERATION 20

2.2.3 VALIDATION OF FBRM TECHNIQUE – INSTRUMENT PARAMETERS 25

2.2.3.1 Probe Position 26

2.2.3.2 Focal Point Position 27

2.2.3.3 Measurement Duration 29

2.2.4 VALIDATION OF FBRM TECHNIQUE - PROCESS PARAMETERS 30

2.2.4.1 Solids Concentration 30

2.2.4.2 Agitation Rate 32

2.2.4.3 Particle Material 33

2.2.4.4 Surrounding Medium 35

2.2.4.5 Temperature 36

2.2.5 CORRELATING FBRM MEASUREMENTS WITH OTHER PARTICLE SIZE ANALYSERS 37

2.2.6 THEORETICAL MODELLING OF FBRM DATA 40

2.2.6.1 Modelling Spherical Particles 41

2.2.6.2 Modelling Non-Spherical Particles 44

2.2.7 FBRM FOR CRYSTALLIZATION CHARACTERIZATION 47

2.2.7.1 Solubility Curve and Metastable Zone Width Determination 49

2.2.7.2 Nucleation Kinetics 50

2.2.7.2 Monitoring Crystal Size 51

2.2.7.3 Temperature Cycling 54

1.2.7.4 Polymorphic Transitions 55

2.2.7.5 Effect of Impurities 56

2.3 PROCESS VIDEO MICROSCOPE (PVM) 57

2.3.1 INTRODUCTION 57

2.3.2 CHARACTERIZATION OF PARTICULATE SYSTEMS USING PVM 61

2.4 ATTENUATED TOTAL REFLECTANCE FOURIER TRANSFORM INFRARED

SPECTROSCOPY (ATR-FTIR) 62 2.4.1 INTRODUCTION 62

2.4.2 INITIAL WORK USING VARIOUS CALIBRATION MODELS 63

2.4.3 THE USE OF ATR-FTIR TO CHARACTERIZE CRYSTALLIZATION PROCESSES 67

2.4.3.1 Seeding 67

Page 5: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

2.4.3.2 Oiling Out 68

CHAPTER 3: SOLUBILITY MEASUREMENT FOR AN ANTI-SOLVENT

SYSTEM USING GRAVIMETRIC ANALSYSIS, ATR-FTIR AND FBRM 69

3.1 ABSTRACT 69

3.2 INTRODUCTION 69

3.3 EXPRESSION OF SOLUBILITY 71

3.4 EXPERIMENTAL WORK AND ANALYSIS 74 3.4.1 GRAVIMETRIC ANALYSIS 74

3.4.2 POLYTHERMAL METHOD USING FBRM 77

3.4.3 SOLUBILITY MEASUREMENT USING ATR-FTIR TECHNIQUE 81

3.4.3.1 Calibration of Probe 81

3.4.3.2 ATR-FTIR Solubility Measurement 87

3.4.4 OVERALL SOLUBILITY MEASUREMENT 90

3.4 DISCUSSION 93

CHAPTER 4: THE EFFECT OF MIXING ON THE METASTABLE ZONE

WIDTH IN ANTI-SOLVENT CRYSTALLIZATION 95

4.1 ABSTRACT 95

4.2 INTRODUCTION 96

4.3. EXPERIMENTAL METHODS 98

4.4 RESULTS AND DISCUSSION 100 4.4.1 COMPARISON OF FBRM AND ATR-FTIR FOR THE DETECTION OF NUCLEATION 100

4.4.2 IMPACT OF PROCESS PARAMETERS ON THE MSZW 104

4.4.3 NUCLEATION KINETICS 106

4.5 DISCUSSION 111

CHAPTER 5: THE USE OF FBRM AND ATR-FTIR TO MONITOR ANTI-

SOLVENT CRYSTALLIZATION AND ESTIMATE GROWTH RATE

KINETICS 113

5.1 ABSTRACT 113

5.2 INTRODUCTION 113

5.3 MATERIALS AND METHODS 115

5.4 RESULTS AND DISCUSSION 116 5.4.1 FBRM RESULTS 116

5.4.2 ATR-FTIR RESULTS 123

5.4.3 GROWTH RATE KINETICS ESTIMATION 125

5.5 CONCLUSIONS 130

CHAPTER 6: SCALE-UP OF ANTI-SOLVENT CRYSTALLIZATION 131

6.1 ABSTRACT 131

6.2 INTRODUCTION 132

6.3 MATERIALS AND METHODS 133

6.3 RESULTS AND DISCUSSION 136 6.3.1 FBRM RESULTS 136

6.4 DISCUSSION 145

6.5 CONCLUSION 147

Page 6: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

7. THESIS CONCLUSIONS 149

8. NOMENCLATURE 153

9: REFERENCES 155

APPENDIX A 171

1. ABSTRACT 171

2. INTRODUCTION 172

3. COMPUTATION FLUID DYNAMICS MODEL 173

4. RESULTS 174 4.1 500ML SCALE 174

4.2 70 L SCALE 177

6. REFERENCES 179

Page 7: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

LIST OF PUBLICATIONS

D. O'Grady & B. Glennon, „Measurement of Nucleation Kinetics in an Anti-solvent

System‟, Conway Institute Conference, Dublin, Ireland, February, 2005

P. Barrett, B. Smith, B O‟Sullivan, J. Worlitschek, V. Bracken, & D. O'Grady, „A Review of

the Use of Process Analytical Technology for the Understanding and optimization of Batch

Crystallization Processes‟, Organic Process Research and Development,9 (3), 348-355, 2005

D. O'Grady & B. Glennon, „Growth and Nucleation Kinetics for an Anti-solvent System

Using In Situ Instrumentation', Industrial Symposium on Industrial Crystallization, Dresden,

Germany, September 2005

D. O'Grady & B. Glennon, 'Characterisation of Anti-solvent Addition Crystallization Using

In-Line Tools', British Association for Crystal Growth Annual Conference, Sheffield, UK,

September 2005

D. O'Grady & B. Glennon, 'Use of In-Situ Instrumentation to Characterise Anti-solvent

Addition Crystallization', AIChE Annual Meeting, Cincinnati, USA, October, 2005

D. O’Grady & B. Glennon, „Comparing a Fast and Slow Crystallization‟ UCD 150

Engineering Exhibition, Dublin, Ireland, January, 2006

Page 8: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

M. Barrett, D. O'Grady, B. Glennon, 'Characterisation of Anti-solvent Addition

Crystallization Using In Situ Tools and Computational Fluid Dynamics', 2006 International

Real Time Analytics Users' Conference, Barcelona, Spain, February 2006

D. O'Grady, M. Barrett, E. Casey & B. Glennon, 'PAT and the Crystallization Toolkit',

Pharmaceutical Manufacturing, 5 (6), 44-47, 2006

D. O' Grady, M. Barrett & B. Glennon, 'Scale-up of Anti-Solvent Crystallization Using in

Situ Tools and Computational Fluid Dynamics', AIChE Annual Meeting, San Francisco, USA,

November, 2006

M. Barrett, D. O'Grady, B. Glennon & E. Casey, 'The Application of CFD to the

Multi-Scale Characterization of Anti-Solvent Addition Crystallization', AIChE Annual

Meeting, San Francisco, USA, November, 2006

D. O’Grady & B. Glennon, „Fundamentals of Anti-solvent Crystallization‟, Mettler

Toledo AutoChem Seminar Series – New Jersey, USA, July 2006

D. O'Grady, M. Barrett, E. Casey & B. Glennon, 'The Effect of Mixing on the Metastable

Zone Width and Nucleation Kinetics In the Anti-solvent Crystallization of Benzoic Acid,

Chemical Engineering Research and Design, Transactions IChemE part A, 85 (7) 945-953,

2007

Page 9: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

D. O'Grady & B. Glennon, „Solubility Measurement for an Anti-solvent Crystallization

System Using Gravimetric Analysis, ATR-FTIR and FBRM', Crystal Growth and Design, in

press

Page 10: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By
Page 11: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

1

ABSTRACT

The anti-solvent crystallization of benzoic acid from ethanol-water solutions using

water as the anti-solvent, at scales ranging from 500 mL to 70 L, is presented. A

thorough review of the literature focusing on anti-solvent crystallization and the use

of in situ tools for crystallization characterization was undertaken prior to

experimental studies. A novel method for measuring solubility in organic systems,

based on the FBRM technique, is presented. This method has been compared to other

solubility measurement techniques and found to be reliable over a wide range of

concentrations. The observed trends were found to be consistent with the UNIQUAC

liquid activity coefficient model. Using the FBRM technique, novel methods for the

determination of nucleation and growth kinetics were developed. By measuring the

metastable zone width under a range of process conditions, including addition rate,

agitation intensity and feed location, a kinetic expression for nucleation in organic

anti-solvent systems that incorporates the influence of agitation, was developed and is

presented for the first time. A kinetic expression for growth was elicited from growth

rate data, gathered using the FBRM technique, and liquid phase concentration data,

gathered using the ATR-FTIR technique. This novel method serves to highlight the

possibility of using in situ tools to measure growth rate kinetics in process and in real

time. The extensive characterisation work performed at the laboratory scale provided

the basis for an investigation of system performance upon scale-up to a geometrically

dissimilar 70 L pilot plant crystallizer. Suitable operating parameters were chosen

based on the small scale characterization and the scale-up was deemed successful as

product of a similar size and yield was produced at the 70 L scale using the same

cycle time.

Page 12: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

2

CHAPTER 1: INTRODUCTION

In recent years the pharmaceutical industry has shown a renewed interest in the field

of crystallization. Crystallization is seen within the industry as a key process

bottleneck during API drug development and manufacture. Through effective

characterization at the laboratory scale, development times can be reduced, effective

and efficient scale-up can be ensured and cycle times can be improved. At the

manufacturing scale, improved crystallization monitoring and optimization can reduce

the number of failed batches, ensure regulatory compliance and enhance downstream

processing operations such as filtration, drying and formulations. Such work

ultimately leads to reduced costs, improved margins and a shorter time to market for

new drugs.

A key element of the process development lifecycle is the scale-up of the

crystallization operation to large-scale, whilst maintaining product and process

compatibility. A major challenge to address is the impact of scale-related effects on

the process. In this work, the characterization of a crystallization process on a variety

of scales is presented. An anti-solvent crystallization process is chosen for study as it

is a technique that has been relatively neglected in the literature, in favour of cooling,

despite its prevalence in API drug development and manufacture. Effective anti-

solvent crystallization scale-up also presents specific mixing-related problems worthy

of investigation.

In situ tools are employed for the measurement of key crystallization parameters such

as crystallization kinetics and supersaturation profile. Focused Beam Reflectance

Page 13: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

3

Measurement (FBRM) is used to study the degree and rate of change of the size, and

number of crystals in suspension and Attenuated Total Reflectance – Fourier

Transform Infrared Spectroscopy (ATR-FTIR) is used to measure solution

concentration. Process Video Microscope (PVM) is used to image crystals as they

exist in process. The use of in situ tools has gained much attention within industrial

and academic crystallization groups for their ease of use, ability to minimize or

remove sampling and real time response to changes in the crystallization system.

Chapter 2 of this thesis is a review of the literature in the field of anti-solvent

crystallization and use of in situ tools for crystallization characterization. Attention is

paid to areas of the field that have been neglected, for example anti-solvent addition

location and the use of organic crystallization systems. A complete review of the

FBRM technique is presented, with the aim being to outline its evolution from novel

technology for particulate systems characterization to the industrial standard for in

situ crystallization characterization.

Chapter 3 will focus on the solubility curve for anti-solvent crystallization systems. A

novel method for the determination of solubility in anti-solvent systems was

developed using the FBRM technique. To assess the reliability of this method two

other techniques, gravimetric analysis and ATR-FTIR, were employed to measure the

solubility. The results for each method were consistent over a full range of

concentrations. The solubility data measured using the FBRM technique also proved

consistent with the UNIQUAC liquid activity coefficient model. Solubility data for

anti-solvent systems is typically presented in terms of grams of solute per gram of

solution. This can make interpretation of key crystallization parameters such as the

Page 14: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

4

metastable zone width and supersaturation difficult. In this case the solubility is

expressed on an anti-solvent free basis (grams of solute per gram of solvent) making

subsequent characterization of the system simple and effective.

A method for the estimation of nucleation kinetics in antisolvent systems is presented

in Chapter 4. An extensive study aimed at characterizing the impact of mixing

conditions on the metastable zone width was undertaken and focused on the impact of

addition rate, agitation intensity and feed location. As expected, a widening of the

metastable zone width at higher addition rates was observed. The feed location proved

to be an extremely important parameter with extreme variation in the metastable zone

width observed for two different feed point locations. The impact of agitation

intensity on the metastable zone width proved to be heavily dependant on the feed

location chosen. By modifying traditional nucleation kinetics equations for the anti-

solvent system, a kinetic expression for nucleation in organic anti-solvent systems that

accounts for the impact of agitation was developed and is presented here for the first

time. The formulation of such an expression was not possible under certain condition,

specifically a feed location close to the wall of the vessel. This was due to a non-

repeatable point of nucleation and in some cases nucleation prior to the attainment of

bulk solubility. In order to better understand the mechanism by which this occurred

computational fluid dynamics (CFD) was employed to model mixing conditions in the

vessel. Poor incorporation of the anti-solvent into the bulk solution, when it was

added close to the wall, proved the important factor and the results of this

investigation can be found in Appendix A.

Page 15: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

5

A quick, efficient and novel method for estimating growth kinetics for organic

crystallization systems is presented in chapter 4. Growth rate kinetics can be

estimated by finding the relationship between the crystal growth rate and its driving

force, supersaturation. ATR-FTIR can be used to monitor supersaturation and FBRM

can be used to measure a relative growth rate, hence the combination of these in situ

tools can, in theory, be used to generate real time, in situ, growth rate information.

For this study a specific set of process conditions is chosen to ensure a low nucleation

rate and repeatable crystallization with consistent yield. The choice of these

conditions is made based on the findings of chapter 3. The experiment is conducted in

triplicate and the results averaged to elicit the desired growth rate information. The

results indicate the ease with which reliable and representative growth rate

information may be estimated, in situ and in real time.

The implementation of such a technique for crystallization control may prove

extremely useful. Much of the research published to date focuses on controlling the

prevailing level of supersaturation in the crystallizer to achieve the desired product

and process performance. While such work is valuable, in order to truly control the

crystallization process the crystal nucleation and growth rates must also be known.

With this information better control can be achieved that allows specific particle size

distributions to be targeted in a specific batch time. The ability to measure the crystal

growth rates in situ may prove extremely useful for tacking this difficult but

potentially rewarding problem.

Page 16: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

6

The scale-up of a successful crystallization process from the laboratory to production

is extremely challenging and is tackled in Chapter 6. Increasing the crystallizer size

impacts the mixing regime, heat and mass transfer properties as well as the surface

area to volume ratio and crystal suspension profile. A combination of these factors

can change the physical and chemical characteristics of the system that may have

been established at smaller scales, resulting in a very different process and product.

In the pharmaceutical industry there are added pressures associated with large scale

crystallization operations due to its time consuming nature and expense, in terms of

operational and product costs. In situ tools can play a vital role in addressing these

many and varied challenges. The direct comparison of continuous and representative

data between scales can be used to identify the source of variation and suggest a

suitable course of action.

For this study scale up from the 500mL laboratory scale to a 70L pilot scale is

performed. The ultimate goal of the scaled up process is to produce a similar yield, in

a similar time and for the crystals to be of a similar size. The extensive

characterization work, outlined in chapters 3-5, is used to determine the set of process

conditions that were most likely to achieve this goal with computational fluid

dynamics employed to identify a suitable feed location. The scale-up ultimately

proved successful with the yield, batch time and crystal size being within acceptable

limits.

Despite this, scale-up success the in situ FBRM data raises concerns over the

mechanism by which the results were achieved. The FBRM data clearly indicates that

while the growth and nucleation rates at both scales were similar, the nucleation point

Page 17: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

7

is different and the pilot scale batch nucleates prior to the attainment of bulk

saturation. The FBRM data also indicates that there is significant segregation in the

vessel.

Page 18: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

8

CHAPTER 2: LITERATURE REVIEW

2.1 ANTI-SOLVENT CRYSTALLIZATION

2.1.1 Introduction

That a solute can have different solubilities in different solvents has long being

recognised. Anti-solvent crystallization exploits this fact to crystallize solid product

from solution. Solution supersaturation is generated by the addition of a second

solvent (anti-solvent) that reduces the solubility of the solute and induces

crystallization. The solvent with the lower solubility is often called an anti-solvent.

The process is often referred to as drowning-out, salting-out, solventing-out or

extractive crystallization.

The industrial use of an anti-solvent for the separation and purification of solid

product was first demonstrated by Gee et al., (1947). Iron impurities were separated

from the desired aluminium sulphate in a 2-ton-per-day continuous pilot plant using

ethyl alcohol as an anti-solvent. This method had apparently first been suggested by

Vittorf (1924). The process relied on the lower solubility of the aluminium sulphate

compared to the iron impurities in alcohol solutions but ultimately proved

uneconomical due to a low cost differential between the crude and finished product.

In the pharmaceutical industry, where the final solid product is extremely valuable

and the product thermal stability is often low, anti-solvent crystallization provides an

important alternative to cooling and evaporative crystallization.

Page 19: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

9

As in a cooling or evaporative crystallization there are a number of issues to consider

when designing an anti-solvent crystallization. The size, number and shape of the

particles produced are important, as these factors can impact on downstream

processing operations such as filtration, drying and formulation. The yield of a batch

is important as the solute in the pharmaceutical industry is extremely valuable and it is

desirable to keep losses at a minimum. It is very important to ensure the correct

polymorphic form is crystallized. In an anti-solvent system supersaturation is

generally high which can lead to a non-stable polymorph. Additionally, because a

second solvent is being added to the crystallization system solvate and hydrate issues

are common. Much research has been carried out to investigate how various process

variables can affect these crystallization characteristics.

2.1.2 Variables to Consider in Anti-solvent Crystallization

2.1.2.1 Choice of Anti-solvent

The first step in the design of an anti-solvent crystallization is the identification of a

suitable anti-solvent. Important characteristics include low solute solubility, high

miscibility in the primary solvent, ease of recovery and potential environmental

impact. Apart from the obvious operational constraints associated with choosing an

anti-solvent, outlined above, there are a number of important crystallization

parameters that can be affected.

Page 20: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

10

The choice of anti-solvent can affect particle size and population, through the

different mixing conditions associated with the addition of anti-solvents of different

density and viscosity. Takiyama et al., (1998) found that the number of crystals

produced by mixing a saturated aqueous solution and a saturated ethanol solution was

greater, by two orders of magnitude, than that produced by mixing two saturated

ethanol aqueous solutions having different concentrations. The reason proposed was

that pure water and pure ethanol mixtures took far longer to mix fully leading to areas

of high supersaturation and elevated nucleation rates. Barata and Serrano, (1998b)

showed that the choice of anti-solvent had some impact on crystal size, for the

crystallization of potassium dihydrogen phosphate (KDP) using three different

aqueous alcohol solutions, but did not propose a mechanism for this effect. However,

further work by this group, using the same system, showed that nucleation kinetics

(Barata & Serrano, 1996a) and growth kinetics (Barata & Serrano, 1998a) were

affected by the choice of anti-solvent. It appeared the nucleation kinetics were

influenced by the dielectric constant of the solution and the size of the alcohol

molecules in the anti-solvent. The growth kinetics were influenced by the size of the

alcohol molecules, in that, large molecules increase the physical barrier to solute

deposition. However it was noted that larger alcohol molecules could be less adsorbed

on the crystal surface, thus inducing higher growth rates.

Finally, product yield can be improved by choosing an anti-solvent in which the

solute is most insoluble. This has been shown by Oosterhof et al., (1999), for the

crystallization of sodium carbonate from aqueous solutions. Diethylene glycol was

deemed the most suitable anti-solvent as the as it provided the highest recovery of

sodium carbonate from solution.

Page 21: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

11

Another important factor to consider when choosing a suitable anti-solvent is phase

separation. Due to the high level of supersaturation typically encountered during anti-

solvent addition the precipitation of a second liquid phase prior to nucleation of the

solute is not uncommon (Barata & Serrano, 1996a; Oosterhof et al., 1999). Care

should be taken to ensure phase separation is not an issue when selecting a suitable

anti-solvent.

2.1.2.2 Anti-solvent Addition Rate

The anti-solvent addition rate is an important process variable that must be controlled

in order to produce a suitable crystal product. High addition rates typically result in a

wider metastable zone and consequently high supersaturation generation rates. This

has been shown and characterized by Pina et al., (2001) for the crystallization of

single and double sulphates from aqueous solutions using methanol as an anti-solvent.

Similarly, high supersaturation levels can be encountered in regions close to the

addition location, especially at elevated addition rates (Takiyama et al., 1998). Thus,

high addition rates often lead to elevated nucleation rates and the formation of fine

crystals. This has been shown by Holmbäck & Rasmuson (1999) for the benzoic acid-

ethanol-water system and Charmoloue and Rousseau (1991) for the crystallization of

L-serine using methanol as an anti-solvent. In both of these cases the crystal

morphology was also influenced by the addition rate. In some cases, very high

addition rates may lead to an increase in the final product size due to elevated

agglomeration rates (Granberg et al., 1999; Jones et al., 1987)

Page 22: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

12

Budz et al., (1986) experimentally determined the most suitable anti-solvent addition

rate, for the crystallization of cocarboxylase hydrochloride from its aqueous solution

by the addition of acetone, by correlating the final product size with addition rate. The

same addition rate produced maxima in the average size and bulk density of the

crystals produced via an unseeded process. A similar result was observed for the

seeded process, but the flow rate that produced the result was double that for the

unseeded process. This indicates the beneficial impact seeding may have, by

maintaining low levels of supersaturation even at high addition rates.

The impact of the addition rate on crystallization is not limited to the crystal size and

shape. High supersaturation levels, resulting from elevated addition rates, can lead to

the formation of undesirable solvates, hydrates and solvates. Kitamura & Sugimoto,

(2003) investigated the crystallization of polymorphs of thiazole-derivative (BPT) by

the addition of water to methanol solutions. When a very low addition rate was used,

nucleation of the solvated methanol crystal form (D) occurred preferentially. At high

water addition rates the proportion of the hydrated crystal form (BH) increased.

Beckmann (1999) studied the effect of addition rate on the crystallization of Abecarnil

from isopropyl acetate by the addition of hexane. At slow anti-solvent addition rates

the most stable polymorphic form, form C, was generally crystallized but at very high

addition rates the unstable B form was observed.

In general, high addition rates have a negative impact on the properties of the final

crystal product, either through poor crystal size distribution, incorrect form or both.

However, at the production scale a balance must be struck between implementing a

Page 23: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

13

slow enough addition rate to ensure a suitable crystal product and ensuring the batch

time is not excessive.

2.1.2.3 Agitation Intensity

The agitation intensity is an extremely important variable in any crystallization. It has

increased importance for anti-solvent crystallization where two liquid phases must be

mixed to create supersaturation. The influence of agitation intensity on crystallization

processes is often conflicting. Charmoloue & Rousseau (1991) showed that increasing

agitation intensity reduced agglomeration levels and improved the impurity profile,

but this in turn led to elevated breakage rates. Similarly, Budz et al., (1986) showed

that a high agitation intensity prevented scaling on the walls of the crystallizer but this

was at the expense of excessive secondary nucleation.

Crystallization kinetics are also influenced in conflicting ways by the agitation

intensity. Barata & Serrano (1998a) showed that increasing agitation intensity

improved growth rates because more crystals were formed initially, and the surface

area available for growth increased. However, the growth rate was limited at very

high agitation intensities. This may have been due to crystal breakage but was not

proved. It has also been reported that agitation intensity has no impact on growth

kinetics. Jones & Mylardz (1989) studied the crystallization of potassium sulphate

from aqueous and aqueous acetone solutions using acetone and acetone-water

mixtures as an anti-solvent in an MSMPR. Crystal growth rates were largely

independent of agitation intensity. This may indicate that the crystallization was

Page 24: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

14

surface diffusion controlled rather than by the bulk transport of the solute molecules

to the crystal.

Agitation intensity has also been shown to affect nucleation kinetics. Barata &

Serrano, (1996b) studying the KDP-alcohol-water system, showed that, between 50

and 400 rpm, induction periods did not depend on the agitation intensity but above

400 rpm the induction periods became smaller. Takiyama et al. (1998) studied the

number of sodium chloride crystals formed by mixing saturated aqueous ethanol

solutions, under different agitation conditions. The number of crystals formed

exhibited a dependence on the agitation condition leading to the conclusion that

nucleation rates were strongly influenced by local mixing.

Further examples of the conflicting influences of agitation intensity, and mixing in

general, have been presented by Paul et al., (2005). For example, increasing the

agitation can improve heat and mass transfer, minimise settling and improve the

purity profile. However the same increase in agitation may cause attrition, secondary

nucleation and entrain gas from the headspace.

2.1.2.4 Concentration Effects

The concentration of the solution to be crystallized is another important variable that

must be considered in an anti-solvent crystallization. Typically the solution

concentration is chosen to maximise the yield per batch; however, there may be

advantages to crystallizing from a solution of different concentration. Crystallizing

from a dilute solution can avoid regions of high supersaturation and improve product

Page 25: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

15

quality, as shown by Budz et al., (1986), for the crystallization of cocarboxylase

hydrochloride by the addition of acetone to dilute and concentrated aqueous solutions.

Crystallizing from concentrated solutions typically results in higher supersaturation

and a smaller final crystal product, due to higher nucleation rates. This has been

shown by a number of researchers including, Holmbäck & Rasmuson (1999),

Beckmann (1999) and Kaneko et al., (2002). Kaneko et al., (2002) studied the

crystallization of sodium chloride from aqueous ethanol solutions using ethanol as the

anti-solvent. At high ethanol solution concentrations the supersaturation was reduced

and monodispersed unagglomerated crystals were formed. As the concentration of

solute in the starting solution increased the number of crystals formed increased.

The influence of solution concentration may also be independent of supersaturation.

Jones & Mylardz (1989), studying the crystallization of potassium sulphate from

aqueous and aqueous acetone solutions in an MSMPR, showed that the growth rate

was strongly affected by the acetone concentration in the solution. At high acetone

concentration the growth rate was significantly reduced for the same level of

supersaturation. Similarly, Granberg et al., (2001) showed that for similar

thermodynamic driving forces the induction time, for the nucleation of paracetamol

from acetone-water mixtures using water as the anti-solvent, increased with

increasing water content in the solution.

The solution concentration may also impact on the polymorphic form of the crystal

product and the rate of transformation from a less stable form to a more stable form.

Kitamura & Sugimoto (2003) studied the crystallization and subsequent

transformation of polymorphs of thiazole-derivative (BPT) by the addition of water to

Page 26: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

16

methanol solutions of the solute. At low initial concentrations the hydrated crystal

form (BH) crystallized independent of the addition rate. At higher initial

concentrations the solvated methanol form crystal (D) nucleated. This may have been

due to the hindrance of the hydration of the BPT molecule. At low initial

concentration, transformation of the BH form to the other polymorphic form (A) was

observed and the transformation rate appeared to decrease with initial concentration.

At higher initial concentrations both the BH and D forms transformed to the A form.

At yet higher initial concentrations the D form transformed to the BH form.

Anti-solvent addition can sometimes result in the production of a large number of

very fine crystals. This is typically due to elevated nucleation rates, encountered at the

point of primary contact between solution and anti-solvent. In cases where the

formation of these fine crystals is undesirable, dilution of the anti-solvent with the

system solvent can be beneficial. The usefulness of this technique has been shown for

the crystallization of potassium sulphate from aqueous solution using aqueous acetone

as an anti-solvent (Mullin et al., 1989) and for the crystallization of benzoic acid from

ethanol solutions using aqueous ethanol as an anti-solvent (Holmbäck & Rasmuson,

1999). In the latter case the anti-solvent feed concentration also influenced the crystal

morphology. Diluting the anti-solvent with the system solvent has shown its use in

other crystallization systems, at two different scales (Jones et al., 1987), and in a

mixed-suspension, mixed product removal (MSMPR) crystallizer (Jones and Mylardz,

1989). Barata and Serrano (1998b) also showed that anti-solvent concentration

influences the yield, due to a dilution effect, and the growth rate, due to the formation

of smoother crystals and a reduced mass deposition rate.

Page 27: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

17

2.1.2.5 Conclusion

It is clear that a wealth of research has been conducted on the factors affecting anti-

solvent crystallization. However there are some areas that warrant further

investigation especially in respect of their application to pharmaceutical

crystallization. Nucleation kinetics for a large number of anti-solvent systems have

been calculated using induction time experiments. However, pharmaceutical

crystallizations are typically performed by continuously generating supersaturation

until nucleation occurs, rather than generating a given level of supersaturation and

then waiting for nucleation to occur. For this reason there is value in measuring the

metastable zone width under various process conditions and applying this data to

calculate nucleation kinetics.

The effect of process conditions, on various crystallization parameters have been

investigated and outlined above. Much of this research has surmised that

supersaturation is a key variable in defining the final crystal product. However, little

research has been done to actually measure the supersaturation and provide an

understanding of the underlying mechanisms involved in the crystallization. The

ATR-FTIR probe provides the opportunity to monitor supersaturation in situ.

One important process variable that has not been characterized adequately is the feed

point location. Since mixing is a vital aspect of the anti-solvent system the position of

the anti-solvent feed warrants investigation.

Page 28: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

18

Much of the work characterising the particle size has been performed using off line

techniques such as microscopy. While useful, these methods rely on sampling which

can be time consuming and unrepresentative. With in line techniques such as FBRM

and PVM now available it is possible to characterize anti-solvent crystallization

without the need to sample.

Finally much of the initial work on anti-solvent crystallization used inorganic

compounds as model systems. Pharmaceutical compounds are in the main organic

compounds and it is best to characterize the anti-solvent system using organic

compounds.

Thus the scope for this study is to characterize anti-solvent addition crystallization

using an organic system with a view to applying the data gathered to pharmaceutical

crystallizations. The solubility of the model system will be measured using a number

of techniques. Metastable zone width data will be measured at various process

conditions including varying feed location. Nucleation kinetics will be calculated

from this data. The crystallization will be examined using in-line tools. FBRM and

PVM will be used to monitor the particle size and shape and ATR-FTIR will monitor

supersaturation. Attempts will be made to scale up the process.

Page 29: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

19

2.2 FOCUSSED BEAM REFLECTANCE MEASUREMENT (FBRM)

2.2.1 Introduction

In recent years, Focused Beam Reflectance Measurement (FBRM) has emerged as a

powerful technique for the characterization of particulate systems. FBRM is a probe

based instrument that measures a function of the size, shape and population of a

particulate system in-process and in real time. It can measure the degree of change to

these properties as well as the rate of change making it a useful tool for the study of

dynamic particulate systems. Its advantage over other off-line techniques is that it

measures the particles in situ, eliminating the need to sample and ensuring that the

particles are measured under process conditions. Off line techniques rely on a

potentially unrepresentative sampling procedure and for some it is necessary to

prepare the sample before analysis through for example sonication or dilution. Even

in the case of off-line microscopy, preparing a slide may alter the particles

significantly through breakage. In the case of crystallization, nucleation, growth or

dissolution may occur after sampling thus providing unrepresentative data.

There is a growing body of literature outlining the use of FBRM for the

characterization of a diverse array of particulate applications. These include

granulation (Sistare et al., (2005)), polymerisation (Hukkanen, (2003)), flocculation

(Blanco et al., (2002)), biofilm (Choi (2003)), gas hydrate systems (Clarke & Bishnoi,

(2004)), wastewater treatment (de Clercq (2004)) plant cell suspensions (Jeffers et al.,

(2003)), and filamentous organism fermentations (Pearson, (2003)). It is in the field of

crystallization however, that FBRM has received most attention.

Page 30: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

20

2.2.2 Mode of Operation

FBRM operates by shining a monochromatic laser beam, of wavelength 790 nm,

generated by a class 1 laser source, via a fibre optic conduit, to an optical assembly

housed within a probe shaft. This optical assembly (Figure 1.1) consists of a lens

mounted eccentrically, and this entire assembly rotates in a circular motion at high

speeds. The circular motion is generated mechanically in the models used for this

work (M400L and S400A) however for use in explosive environments the assembly

can be rotated pneumatically. The user may change the speed of rotation (scan speed)

in the software to values of 2 m s-1

, 4 m s-1

, 6 m s-1

and 8 m s-1

. For the duration of

this work the scan speed was held constant at 2 ms-1

at which the measurement range

is between 0.5 and 1000 µm. At higher scan speeds the measurement range can be

extended at the upper level but in doing so the sensitivity at the lower end of the range

is compromised. As the monochromatic beam passes through the lens it is focused to

a fine point (d ~ 4μm) just inside the probe window. The circular motion of the

assembly traces the fine point of the laser over the circumference of a circle (Figure

1.1).

When the beam comes in contact with a particle, light is reflected in all directions.

Some of the light is reflected back up the probe to a detector. The detector measures

the time duration for which the light is reflected (Figure 1.2). This time duration can

be used to measure a chord length across the particle according to equation 2.1.

l = t × v Eq. 2.1

Page 31: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

21

Where l is the chord length (µm), t is the time duration of the backscatter (s) and v is

the speed of rotation of the laser beam (µm s-1

). For a chord to be accepted the

intensity of backscattered light must exceed a certain threshold. Additionally the rise

time of the electronic pulse generated by the backscattered light must be short. The

theory is that only particles that pass directly through the focal point will have a short

rise time. Thus particles that are too far from the focal point and are out of focus will

not be considered (Sparks & Dobbs 1993).

Since the laser rotates at high speed and a chord length is an entirely random

measurement there is the possibility to measure thousands of chord lengths per

second.

Page 32: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

22

Figure 2.2: Chord Length Measurement (Mettler Toledo Users Site)

These individual measurements can be combined to produce a chord length

distribution (CLD) that provides a robust fingerprint of the particulate system. This

distribution is sensitive to changes in the size, number and shape of particles in the

system. Importantly it is sensitive not only to the degree of change but also the rate of

change making the FBRM technique extremely powerful for studying particulate

systems.

The measurement duration is the length of time for which the instrument scans before

the data are logged and can be altered using the software, to between 1 s and 1 hr. A

long measurement duration ensures a large number of counts and therefore

statistically robust data. However, if the measurement duration is too long, key

process changes that occur quicker than the measurement duration will be missed.

Mettler Toledo recommends that a measurement duration of 10 s is used in the

Page 33: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

23

laboratory and 1 min is used in the plant. This gives a good balance between

sensitivity to process changes and statistically robust data.

A typical FBRM installation consists of a probe, which is connected via a fibre optic

cable to an electronics box known as the field unit. The field unit is connected to a PC

for instrument operation, data acquisition and data review (Figure 1.3). A more

detailed description of how FBRM operates can be found in the literature (Sparks &

Dobbs, 1993; Tadayyon &Rohani, 1998; Barrett & Glennon, 2002).

An important attribute of the FBRM software is its ability to trend important statistics

of the chord length distribution over time (Figure 1.4). For example the number of

particles in a small size range, e.g. 1 – 10 µm, can be trended over time to assess the

change in the fine end of the distribution. Additionally the same can be done for a

large size range, e.g. 100 – 250 µm. In fact it is possible to focus on any size range or

statistic (i.e. mean, median, or mode) of the distribution and trend it over time to

improve understanding of the process at hand. It is also possible to weight the

statistics to provide the most relevant information. This is especially useful for

monitoring a dynamic process such as attrition where there is a decrease in the size of

the coarse particles and an increase in the number of fine particles (Kougoulos

2005c). Suitable statistics for measuring such a process, apart from counts in

individual size ranges, may be the median for fine particles and a square weighted

mean for the coarse particles. By trending suitable statistics over time it is possible to

not only confirm that agglomeration is taking place but also assess the rate at which it

is occurring and when it has stopped.

Page 34: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

24

Figure 2.3: Typical FBRM Setup with probe, field unit and computer (Mettler Toledo User‟s Site)

Figure 2.4: Chord Length Distribution Trended over time (Mettler Toledo User‟s Site)

Page 35: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

25

The chord length distribution itself can be manipulated to analyse key particle

parameters. A variety of weightings may be applied to the chord length distribution in

order to highlight the fine and coarse end of the distribution. A 1/length weighting is

applied by dividing the number of counts in each size range by the midpoint of that

size range. This emphasises small particles. A square weighting is applied by

multiplying the number of counts in each size range by the square of the midpoint. In

this way the large particles are emphasised. Heath et al., (2002) provides an excellent

overview of the difference in these distributions and how they may be used and

Fujiwara et al., (2002) compares how different weightings can be used to identify the

onset of nucleation in the aqueous batch cooling crystallization of paracetamol.

2.2.3 Validation of FBRM Technique – Instrument Parameters

Much work has been carried out in order to validate the FBRM technique and assess

how process and instrumental parameters affect the chord length measurement.

Instrumental parameters that have been investigated include probe position, focal

point position, scan speed and measurement duration. Process parameters investigated

include solids concentration, agitation intensity, surrounding medium and particle

material.

It should be noted that the FBRM technique existed under a different name in the

1990s. The Par Tec 100 and Par Tec 200 were precursors to the current FBRM probe.

The Par Tec probes operated on the same principle of backscattered light as the

current FBRM probe. It was possible to display the raw chord data, a spherical

equivalent diameter or a volume distribution. A lot of early validation work was

Page 36: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

26

performed on the Par Tec probe that still applies to the current FBRM models. For the

following sections all examples are for the current FBRM probe unless otherwise

stated.

2.2.3.1 Probe Position

Probe positioning plays a vital role in ensuring that a representative sample is

presented to the window. Poor probe position can lead to unrepresentative data and

poor process understanding. It is imperative to position the probe so the particles are

flowing into the window. Tadayyon & Rohani (1998) identified this need and also

suggested the probe window should be sufficiently far away from any light reflective

object such as the stirrer blades to avoid the generation of signal noise, especially at

low concentrations. Barrett and Glennon (1999) assessed the effect of probe position,

on the CLD of alkaline frit in water, at two different scales (1.5 L and 70 L stirred

tanks). The most statistically robust data were achieved when the count data were

highest. This was achieved, at both scales, with the probe mounted from the side at an

angle to the dominant flow direction. The manufacturer provides advice on

positioning the probe in a reactor and a pipeline. In both cases the probe must be

placed at an angle to the dominant flow direction (Figure 1.5).

Page 37: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

27

Figure 2.5: Suitable probe position (circled) Mettler Toledo Lasentec Users Site

2.2.3.2 Focal Point Position

The choice of focal point is extremely important to ensure representative data is

gathered. If the focal point is set far from the probe window there will be a reduced

sensitivity to fine particles. This is due to the deterioration of the light signal as it

leaves and re-enters the probe. This fact has been identified by a number of

researchers including Worlitschek and Mazzotti (2003), for the study of a single

particle in water and in a dispersion of fine particles and Heath et al. (2002), for the

study of the CLD of aqueous calcite suspensions. Heath et al., (2002) also showed

that setting the focal point far from the window increased sensitivity to coarse

particles. The reason given for this was that large particles are less likely to enter the

Page 38: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

28

measurement zone when the focal point is set close the window. Monnier et al.,

(1996) studied the effect of the focal point position on the measured mean size of

latex and orgasol particles in various solvents using the Par Tec device. The mean size

varied depending on the focal point chosen. For small particles the most accurate size

data was achieved when the focal point was set close to window. For large particles

the opposite was true with the optimal focal point being between 1.5 and 2 mm from

the window. Ruf et al., (2000) monitored how the distance of a particle from the focal

point affected the measured chord length of a single particle. They found that as the

particle moved farther and farther from the window the size of the measured chord

length decreased. This was due to the weakening and broadening on the laser beam

that makes it harder for the FBRM signal-processing unit to detect precisely the

boundary of the particle. For the same reason, a number of chord lengths much

smaller that the particle itself were measured. Worlitschek and Mazzotti (2003) found

similar results for the single particle systems and their findings suggested that when

the whole particle size distribution is of interest, the best focal point to detect large

and small particles was at the probe window. Furthermore, they suggest that if a

specific known size range of particles is of interest then the focal point should be

positioned half the diameter of the smallest particle away from the window. This

eliminates the risk of not detecting the smallest particles.

In each of these studies the optimum focal point position for fine material was

determined to be close to the window. Since fine crystals typically are of most interest

in crystallization applications the focal point is usually set at the window. While some

sensitivity to larger crystals is lost, this set-up ensures maximum sensitivity to the

more important fine crystals. Some of the newer laboratory-based FBRM systems

Page 39: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

29

(S400 range) have a fixed focal point at the window. In the case where larger particles

are of interest, for example in granulation, a probe with a variable focal point may be

preferable.

2.2.3.3 Measurement Duration

The choice of measurement duration is a balance between ensuring the collection of

statistically robust data and ensuring key process events are not missed. Additionally,

applying a short measurement duration results in a large amount of data that can be

difficult to handle. Pearson (2004) assessed the effect of measurement duration on the

CLD of filamentous bacteria measurement measured offline. Values between 1 and 60

s were investigated and a measurement duration of 30 s was chosen as it showed only

minor variation in all the statistics of the CLD examined and allowed relatively rapid

collection of data. Monnier et al., (1996) investigated the effect of measurement

duration on the size distribution measured using the Par Tec device for glass beads

between 125 and 250 µm (measured by sieving) in water. The cycle time was varied

between 0.2 s and 3.2 s for the same suspension of glass beads. For a measurement

duration of 0.2 s the total count data was too small to allow correct statistical

processing of the data. When the measurement duration was changed to 3.2 s the

reproducibility of the data increased significantly and the standard deviation halved

indicating a more robust measurement.

Page 40: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

30

2.2.4 Validation of FBRM Technique - Process Parameters

2.2.4.1 Solids Concentration

An increase in solids concentration leads to an increase in total counts measured by

the FBRM. This is simply due to the presence of more particles in the measurement

zone at higher concentrations. A linear relationship between solids concentration and

counts has been observed by (Pearson et al., 2004), for a filamentous bacterial system,

and, by Monnier (1996), for aqueous suspensions of glass beads. Other studies have

noted a similar linear relationship but at very high concentration count data has been

observed to stop increasing with increasing solids concentration. Barrett and Glennon

(1999) investigated the relationship between the CLD and solids concentration

aqueous alkaline fit suspensions at scales ranging from 500 mL to 70 L. At all scales

the relationship between counts and concentration was linear at low concentration but

tapered off at high concentrations. Similar results were found by Heath et al., (2002)

investigating the change in total counts of aluminium and calcite particles in water

due to an increase in solids concentration.

The reason for this non-linearity is due to an increase in the instrument dead time at

high concentrations. At low concentration only a small fraction of the distance

traversed by the laser is used to measure a particle. However at high concentration a

significant percentage of the distance is used reducing the time available to detect

other particles. This dead time is slightly larger than the time spent traversing the

particle because the reflected light must return to zero for a short period between

particles. Additionally, the dead time is increased at high concentrations because

Page 41: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

31

particles outside the viewing zone are measured and then rejected if the rate at which

the signal increases is not fast enough. Furthermore, overlapping particles may be

counted as one large particle exacerbating the problem. It is possible to calculate this

dead time but it is system specific. Heath et al., (2002) measured the value for

aluminium and calcite particles.

Another possible reason for the non-linear relationship between counts and

concentration is that at high concentration, the intensity of light at the focal point can

be reduced by blocking of the laser path by other particles. Therefore, the intensity of

backscattered light may not be sufficient to generate pulses with sufficient amplitude

to be detected by the instrument (Tadayyon & Rohani 1998).

The impact of solids concentration on the size of the particles measured has also being

investigated by a number of researchers. In some cases a strong influence has been

observed but in others the impact has been negligible. Dowding et al., (2001) showed

that increase in solids concentration, of poly(vinyl chloride) beads in water, resulted

in a decrease in length weight mean chord length and an increase in the cube weighted

mean chord length. It was proposed that at high solids concentration the mean chord

length increased due to clustering of the PVC beads and the cube weighted chord

length decreased due to the reduced probability of seeing the largest particles. Law et

al., (1997) analysed the size distribution measured by the Par Tec 100 of various

particles (pollen, sand, algal cells etc) in water at solids concentrations ranging from

10 to 50,000 mg L-1

. As concentration increased the mean and median particle size

remained relatively constant. The mean standard deviation was about 12% and the

median standard deviation was about 5%. A slight trend to a larger mean particle size

Page 42: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

32

at higher concentrations was attributed to overlapping of particles leading to a larger

particle size.

Monnier et al., (1996) made similar investigations for glass beads in water using the

Par Tec instrument but found the mean size to be independent of solids concentration

in the range 0 and 450 g L-1

2.2.4.2 Agitation Rate

Effective agitation is critical to ensure that a representative sample of the particulate

system is presented to the window. Heath et al., (2002) noted a reduction in the

average measured size as agitation intensity was reduced. Settling of the particles was

visibly noted confirming that the large particles were falling out of the measurement

zone impacting on the measured size.

Statistical robustness may also be improved by agitating at a sufficient level. Pearson

et al., (2004) examined the effect of agitation rate on the measured CLD of a

filamentous bacteria system. Agitation was varied between 100 and 500 rpm at a

constant solids concentration. As agitation increased the total number of counts

increased as more of the mycelial aggregates were presented to the probe window.

The mean and median remained constant indicating the new particles presented to the

window were no different in size. However by increasing the count data a more robust

measurement was achieved.

Care must also be taken not to entrain air into the system at elevated agitation rates.

Dowding et al. (2001) investigated the influence of agitation intensity on the CLD of

Page 43: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

33

PVC in water at a fixed concentration (2.3 w/w%). The average chord length

increased with increasing agitation and further investigation showed that at higher

agitation levels air bubbles were been entrained in the suspension. These bubbles

were measured by the FBRM as particles in the system and the average chord length

increased. Monnier et al., (1996) investigated the effect of agitation intensity on the

size distribution of glass beads measured by the Par Tec 100 instrument in toluene and

water. In water the total counts increased 10-fold when the agitation intensity was

increased from 250rpm to 750 rpm. However, at higher agitation levels (up to

2000rpm) the total counts actually dropped. This was not the case for the toluene

system where total counts increased with increasing agitation from 500 rpm to 1500

rpm. At the highest values the total counts increased significantly. This was put down

to bubbles being entrained at this high level of agitation. The mean size of particles

measured increased with increasing agitation. This may have been due to a more

representative sample being presented to the window.

An increase in agitation intensity does not always result in changes to the chord length

distribution. Law et al., (1997) found no effect on the size distribution of various

particles (pollen, sand, algal cells etc) generated by the Par Tec 100 instrument for a

change in the agitation intensity. The agitation intensity was varied between 500 and

1000 rpm and no effect on the mean or median size was observed.

2.2.4.3 Particle Material

Since the FBRM technique is based on reflection the optical properties of the particles

under investigation are extremely important. Sparks & Dobbs (1993) investigated a

Page 44: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

34

number of different materials using the Par Tec instrument. They surmised that

particle shape, or surface characteristics (spherical, flat, porous), absorbing properties

(colour), refractive index (relative to suspension fluid) and high or low reflectivity

may influence the measurement. They found that large transparent particles in general

displayed very low count rates while opaque highly reflective particles displayed the

highest count rates. It has been noted that in some cases FBRM is a poor technique for

measuring extremely reflective particles such as TiO2. This is due to specular

reflection in which light is reflected in every direction except back to the probe. For

most particulate systems, however, this is not a problem

Ruf et al., (2000) showed the importance of the optical properties of the particle

material by measuring a single particle of ceramic material and a single particle of

quartz in toluene. For the ceramic material a single chord length was returned when

the particle was measured at a specific position. However for the quartz particle a

number of chord lengths were measured. This was due to a splitting of the signal due

to different backscattering properties of different faces of the crystal and a weakening

of the signal. This effect was not observed when the particle was measured in air. This

was due to the more favourable refractive index difference between the fluid and the

particle. Heath et al., (2002) compared the CLD generated by various size fractions of

aluminium and calcite particles. These particles were chosen because they had similar

densities and therefore a similar volume fraction when made up by weight leading to

similar flow properties to maintain suspension. The calcite particles showed a larger

tail in the fine end of the unweighted chord length distribution and this was attributed

to abrasion of fine particles from the particles and the finer edges of the particles

producing more fine counts. The aluminium particles were not subject to this abrasion

Page 45: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

35

and had more round edges leading to a more symmetrical unweighted distribution.

The square weighted distributions proved remarkably similar despite dissimilar

morphologies and a presumed difference in reflectivities. This was not expected to be

applicable to all materials and was supposed to be more a consequence of the

relatively low aspect ratio of the two materials.

2.2.4.4 Surrounding Medium

The medium in which particles are suspended may also influence the FBRM

measurement. This is typically down to differences in refractive index associated with

different solvents. Monnier et al., (1996) investigated the influence of solvent on the

size distribution of glass beads in water, THF, toluene and isoctane using the Par Tec

instrument. For the organic liquids the total counts increased at all concentrations as

refractive index decreases. This was not the case for water however where counts

were higher compared to those in isoctane despite having a higher refractive index.

This indicates that there is a parameter other than refractive index that influences the

total counts. The size distribution of the measured glass beads was very similar in all

solvents. Ruf et al., (2000) also showed the impact of refractive index by measuring

the chord length of a single 200μm ceramic bead in water, toluene and air. They

showed that when measuring the ceramic bead in solvent the actual focal point is

different to the nominal focal point. This is due to the difference in refractive index of

air, found inside the probe and the solvent, found outside the probe. It was suggested

that the focal point should be kept at or inside the window to minimise the number of

times the light passes through a medium of differing refractive index.

Page 46: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

36

2.2.4.5 Temperature

In many particulate systems, specifically crystallization, the temperature of the system

will change. Monnier et al., (1996) investigated the effect of temperature on the

particle size distribution measured by the Par Tec 100 instrument by measuring the

evolution of mean size and total counts for glass beads in water as the suspension was

subjected to a series of heating and cooling ramps. Total counts decreased form 15000

s-1

to 12000 s-1

for an increase in temperature of 60 C. The counts returned to their

original value on cooling back to the original temperature. Mechanical effects on the

probe, i.e. expansion of the probe, were discounted due to the small variation in

temperature. The refractive index of the suspension changes with increasing

temperature and this may have an effect but the refractive index of the solvent

decreases with an increase in temperature and this should lead to an increase in counts

at elevated temperatures. No parameter was identified for this phenomenon. The mean

size was unaffected by any change in temperature.

Page 47: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

37

2.2.5 Correlating FBRM measurements with other particle size analysers

In order to further validate the FBRM measurement a number of researchers have

sought to compare the chord length distribution measured by FBRM and the particle

size distribution measured by other size measurement techniques. Correlation between

the FBRM data and other particle size analysis techniques will depend on a number of

factors, including, technique investigated, particle system, and sampling mechanism.

FBRM data has been compared to a number of different techniques for various

systems and Heath et al., (2002) provides a comprehensive review of much of the

research carried out in this area.

Laser diffraction is an industry standard particle size analysis tool and is used

extensively in the pharmaceutical industry for product release. For this reason much

work has been devoted to comparing it to the FBRM measurement. Alfano et al.,

(2002) produced a correlation between the mean chord length measured by FBRM

and the median particle size measured by laser diffraction for microcrystalline

cellulose. However, it was noted that when assessing the validity of the correlation

two important issues come into play. Firstly, laser diffraction yields a true spherical

equivalent particle size distribution whereas FBRM provides a chord length

distribution that is related in a complex way to the true particle size and shape.

Secondly, to measure a particle size distribution using laser diffraction it is necessary

to pull a sample of the particulate system from the reaction vessel whereas the FBRM

probe measures the system in-situ. They noted that the FBRM probed the

microcrystalline cellulose slurry in the high-shear region of the impeller tip region,

whereas the laser diffraction measures sample aliquots taken from the low-shear bulk

Page 48: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

38

region, transferred to a low-shear sample cell. Additionally, the laser diffraction

method required a 30-fold dilution of the sample. These two points clearly highlight

the difficulty associated with comparing these techniques. However, despite these

difficulties a number of researchers have achieved useful correlations between FBRM

and laser diffraction for spherical or near spherical particulate systems.

Heath et al., (2002) found a linear relationship between the d50 measured using laser

diffraction and the median chord length measured by the FBRM for aqueous

aluminium suspensions. For large particles the correlation was improved by applying

a length weighting to the FBRM data. A similar linear relationship between the

techniques was observed by Abbas et al., (2002), for aluminium oxide particles. In

this case, the volume weighted mean measured by FBRM and the volume based

diameter measured using laser diffraction were used and the study only concentrated

on particles less than 25μm. It was noted that a more critical evaluation of the any

correlation would only be possible if the size range studied was increased to 500μm or

more.

Another common particle size analysis technique that has been compared to the

FBRM measurement is sieving. McDonald et al., (2001) found a correlation between

the mean cube weight of the chord length distribution and the particle size measured

by sieving for plant cell suspension cultures. The correlation was dependent on the

system studied with wild Chinese cucumber cells producing a linear correlation and

rice cells producing a non-linear fit. Tadayyon & Rohani (1998) compared the size

distribution of ion exchange resin generated by the Par Tec 100 and sieving. For

spherical particles there was a good correlation between the volume weighted chord

Page 49: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

39

length distribution and the unweighted particle size distribution generated by the

sieve. However the difference between the actual particle size distribution and the

chord length distribution increased as the particle shape deviated from a sphere and

the size distribution became non-normal..

Comparisons between the size distribution measured using the Par Tec 100 and

various measurement techniques, including laser diffraction, image analysis and

coulter counter have also been made. In general it has been observed that the Par Tec

device overestimated the size of the smallest particles and underestimated the size of

the largest particles. This was the case for a number of systems including fluvial

suspended sediments, Philips & Walling (1998), sand particles, Law et al., (1997),

and a variety of spherical particles (coulter microspheres, Bayer resin, orgasol and

micronized pharmaceutical compound), Monnier (1996). In the case of Philips &

Walling (1998), and Law et al., (1997), a calibration model was derived that

accounted for this deviation.

Sparks & Dobbs (1993) compared the spherical equivalent diameter and volume

diameter generated by the Par Tec device with a Microtrac laser forward scatter based

technique. For opaque lightly coloured material such as aluminium hydrate results

between the two methods correlated well. For large transparent particles the

correlation was poor.

Page 50: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

40

Comparing size distributions using different techniques is difficult. First and foremost

the various techniques produce a measure of a different particle size property. Abbas

et al., (2002) noted that when comparing the size distribution of aluminium oxide

using laser diffraction Turbiscan and FBRM the Malvern returned a volume based

measure D[4,3], the Turbiscan measured a mean D[3,2] based on the volume fraction

and the FBRM reported a weighted mean based on a chord length distribution. Also,

as noted above, Alfano et al., (2002) saw the difficulties associated with correlating

between two techniques with different measurement principles as well as the inherent

sampling issues associated with off-line techniques. Despite these issues it is possible

to form useful correlations between the FBRM measurement and other size analysis

techniques. These correlations will vary greatly depending on the measurement

technique used and the system under investigation Heath et al., (2002).

.

It is clear that a large volume of research has been carried out on the comparison of

the FBRM method with other particle size analysis techniques. A number of empirical

correlations have been found that typically work well for spherical particles with

suitable optical properties. In addition to empirical methods theoretical studies have

been carried out to investigate whether the chord length distribution measured by

FBRM is a true representation of the particulate system in question.

2.2.6 Theoretical Modelling of FBRM Data

The use of theoretical methods to convert chord length data into particle size data has

gained much attention in recent years. Empirical methods have been used to compare

the chord length measurement with other particle size analysers and the degree of

Page 51: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

41

success in these correlations depends greatly on the shape of the particle. A suitable

model that can convert the measured chord length of a variety of particle sizes and

shapes into a usable particle size distribution would prove useful in further validating

the FBRM measurement.

In a more practical sense a correlation between the FBRM measurement and another

particle size measurement device may prove useful industrially. The standard for

particle size measurement across many industries, most notably the pharmaceutical, is

laser diffraction. A calibration between the chord length data measured using FBRM

and the mean size measured by laser diffraction would prove extremely useful in

predicting whether a crystal slurry, for example, was within specification.

In addition to this practical application the conversion of chord length data into

particle size data is an extremely difficult problem and is mathematically complex.

The difficulty of the problem and the challenges associated with it have attracted

many researchers. Initial study in this area concentrated on modelling the FBRM

measurement of spherical particles. The projected area of a spherical particle is

always a circle no matter what the orientation of the particle. This makes conversion

of the chord length data into particle size data less challenging than for other

morphologies.

2.2.6.1 Modelling Spherical Particles

For spherical particulate systems, e.g. bubbles and droplets, a number of researchers

have had success converting chord length distributions into particle size distributions.

Page 52: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

42

Liu et al., (1998) analysed the relationship between bubble sizes and chord lengths in

a heterogeneous bubbling system. They developed a forward transform to calculate a

chord length distribution from a known bubble size distribution and three backward

transforms to infer bubble sizes from a chord length distribution. The techniques were

tested and compared using Monte-Carlo simulations. Simmons et al., (1999)

presented two methods for estimating droplet size distributions from chord length

measurements. The first is a probability apportioning method (PAM), that assumes

particles are randomly cut and calculates the diameter probability distribution from

each chord size detected. The distribution is accumulated over all chord

measurements. The second is a finite element method (FEM) that solves

simultaneously the equations relating the chord data and the diameter distribution.

Both methods were tested on sets of chord data developed from ideal probability

functions, a distinct element simulation and actual experimental data. It was found

that the PAM method was fairly robust when the particle diameters were known, but

the FEM method was generally more applicable when there was a wide range of

unknown particle diameters. Both models predicted the chord length distribution of

droplets in a 63mm pipeline measured by FBRM with a high degree of accuracy. The

data returned by the FEM model proved slightly better. It was concluded that the

FEM model was better for engineering applications where actual particle sizes are

unknown. Langston et al., (2001) improved the PAM method proposed by Simmons

et al., (1999) to take into account particles of unknown size distribution by

incorporating Bayes theorem and an iterative procedure (PAM 2). The improvement

comes from the implementation of the chord length data in a collective manner rather

than in isolation. A range of particle size distributions were modelled accurately

where the size of the particles was known and unknown. Experimental studies were

Page 53: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

43

carried out on droplets of oil in water and the model predicted the measured data

accurately.

Barrett & Glennon (1999), moved away from perfectly spherical systems and

successfully modelled FBRM chord length data of dilute aqueous alkaline frit

suspensions of known size distribution, using a method similar to that of Simmons et

al., (1999). They assumed that the suspension was dilute enough to ensure particle

interactions were negligible and a good agreement between experimental results and

the model was achieved. It was concluded that the good agreement was only expected

if the particles were spherical, or at least if the dominant projected area was circular in

profile.

The research outlined above indicates that there has been some success in modelling

the chord length distribution of spherical particles. When modelling spherical

particles a 2-dimensional approach is sufficient since the orientation of the particle

does not affect the projected area. However for non-spherical particles, i.e. ellipses,

the orientation of the particles plays a very important role since it affects the projected

area on which a chord length is measured.

Tadayyon and Rohani (1998) developed a model to predict the response of the Par

Tec instrument in measuring the CLD of a suspension of spherical and ellipsoidal

particles and to infer the actual PSD using the measured CLD output. The model

showed that the measured CLD was reasonably accurate for spherical particles

however the measurement progressively deteriorated as particles became more

Page 54: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

44

ellipsoidal. For spherical particles the error in the mean sizes reported by both

methods was improved by employing a volume weighted mean on the Par Tec data.

A further aspect of this research used the Par Tec instrument to examine a suspension

of ion exchange resin of normal size distribution in water. The chord length data was

transformed into particle size data using a model and compared to the actual particle

size distribution measured by sieving. For a normal size distribution the model

predicted the particle size distribution well. However for a non-normal size

distribution the model performed poorly.

Tadayyon and Rohani (1999) showed this when the model they proposed was

unsuitable for ellipsoidal particles and became more and more unsuitable as the

ellipse became more elongated and less spherical. For real particulate systems,

especially crystal systems, the particle morphology is typically highly non-spherical.

In order to model the chord length distribution of a real particulate it is absolutely

necessary to analyse non-spherical particles.

2.2.6.2 Modelling Non-Spherical Particles

Modelling non-spherical particles is difficult due to the difference in the projected

area depending on the orientation of the particle. A number of researchers have

developed methods to account for this that are usually computationally intensive.

Langston et al., (2001) attempted to analyse non-spherical particles using a 2-

dimensional approach by developing the PAM 3 model, an improvement on the PAM

2 method, Langston et al., (2001). This model was used to predict particle size

distributions from chord length measurements in a 2-dimensional assessment of non-

Page 55: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

45

circular shapes, which included circles, ellipses and “blocky” shapes. They supposed

that the model was applicable to 3-dimensional assessments but conceded that the

random orientation of the particles would have to be accounted for. Wynn (2003)

derived a relationship between the underlying particle size distribution and the

measured chord length distribution measured using FBRM. A number of assumptions

regarding the measurement technique were taken into account. The principle of the

method is similar to earlier “peeling” methods, but an improvement is made here in

that the method is shown to be stable in some cases, due to finer spacing of the size

intervals.

In order to account for random particle orientation a 3-dimensional approach is

needed. Ruf et al., (2000) proposed a 3-dimensional model to transform particle size

distributions into chord length distributions. The shapes of different particles were

defined mathematically and the 2-dimensional projection of each particle was

calculated for every possible orientation in 3-dimensional space. The chord length

distribution of each 2-dimensional projection was calculated and by summing these

individual distributions and applying suitable weightings the overall chord length

distribution was calculated. Theoretical chord length distributions were calculated for

a variety of particle morphologies including ellipsoids and cuboids. This model

represents an improvement on the 2-dimensional model proposed by Tadayyon &

Rohani (1999). It can account not only for particles of non-spherical morphology but

it can also account for particle populations of the same shape but different size.

Li & Wilkinson (2005) and Li et al. (2005) presented a theoretical and experimental

study of the conversion CLD to PSD and PSD to CLD for non-spherical particles.

Analytical solutions were used to calculate the PSD-CLD models for spherical and

Page 56: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

46

ellipsoidal particles and numerical solutions were used to non-spherical particles. To

convert CLD to PSD a non-negative least squares (NNLS) method that is insensitive

to the measurement noise and the particle shape was employed. For the model to work

it is necessary to give it some information on the shape of the particles in question.

The effectiveness of the proposed methods was validated by extensive simulations. To

test the validity of these models on a real experimental system the particle size

distributions of ceramic beads, plasma aluminium and zinc dust were measured using

image analysis and compared to the restored particle distribution calculated from

experimental chord length data. Image analysis was used to define the shape of the

particles and values of sphericity, circularity and chunkiness were used in the model.

The results compared well for each of the test particles. The accuracy of the

translation depended heavily on particle shape, which determines the optimal aspect

ratio that is used in the model.

Worlitschek et al., (2005) developed a method to restore the PSD of a particulate

system from the measured CLD. The restoration constitutes a mathematically ill-

posed problem and its solution is based on a two-step procedure. Firstly, the

computation of a matrix that converts the PSD of a population of particles with given

shape into the corresponding CLD using a 3-dimensional geometric model and

secondly the solution of the resulting linear matrix equation for the PSD. The

restoration of the PSD of spherical particles represented the easiest case whereas for

non-spherical shapes such as octahedral and needles lead to highly ill-posed inversion

problems. However, experimental investigations were carried out on populations of

paracetamol crystals (octahedral shape) and the model predicted measured data

accurately.

Page 57: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

47

It is clear that the goal of producing a particle size distribution from a measured chord

length distribution is becoming a reality. The restoration of the PSD of a spherical

population of particles from a measured CLD is a relatively simple problem and

recent work has shown that this is now possible to restore the PSD of populations of

ellipses and octahedra. Further work is needed to model highly non-spherical

morphologies such as needles.

2.2.7 FBRM for Crystallization Characterization

While the possibility of inferring actual particle size distributions from chord length

distributions may be possible in the future it should be noted that all of the current

models require some information on the shape of the particles. For a well-behaved

system of spherical particles, such as oil in water, this is possible, but for a

crystallization system this is extremely difficult and may be impossible. In a

pharmaceutical crystallization system, for example, the particle morphology is

typically needle-like and the aspect ratio is usually non-uniform. Attrition can lead to

numerous small crystals with a low aspect ratio and agglomeration can lead to large

disordered particles with no recognisable shape. For a polymorphic system the shape

of the crystals can change over the course of the crystallization. These factors make it

extremely difficult to decide on the shape of the crystals under investigation.

FBRM avoids this problem by assuming no shape. The chord length distribution is a

function of the size, shape and number of particles that enter the measurement zone.

In a crystallization system, the chord length distribution will change if the size, shape

Page 58: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

48

or number of crystals entering the measurement zone changes. Consequently key

crystallization phenomenon such as nucleation, growth, attrition, agglomeration,

dissolution and polymorphic transition can be identified and monitored. To facilitate

identification and monitoring of these dynamic crystallization processes, PVM may

be implemented as c complementary technique. When one considers that this can be

done in situ in real time the power of FBRM for monitoring crystallization systems

become apparent. When changes to the chord length distribution can be related to

process parameters, such as cooling rate, anti-solvent addition rate, agitation or seed

loading, it is possible to design a crystallization that will produce a suitable particle

size distribution that behaves appropriately during downstream processing operations

such as filtration, drying and formulation.

The vast majority of the research carried out on crystallization using FBRM is

performed in the pharmaceutical industry. Due to the secretive nature of much of the

research it is rare for work to be published. In fact the number of published articles

documenting the use of FBRM for the characterization of crystallization is

surprisingly few. Some of the research undertaken industrially is presented at

conferences, the proceedings of which can be useful. Of particular use is the Mettler

Toledo Users Site www.mt.com/lasentec where papers presented at the annual Mettler

Toledo Real Time Analytics Users Forum (formerly the Lasentec User‟s Forum) are

posted. While one must be wary of studying in too much detail the research presented

by the vendor itself, it nonetheless provides a wealth of information on how FBRM

has been used to design, develop and optimise crystallization processes in many of the

major pharmaceutical companies.

Page 59: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

49

2.2.7.1 Solubility Curve and Metastable Zone Width Determination

The starting point for the characterization of any crystallization system is the

solubility curve. Barrett and Glennon (2002) implemented the polythermal method,

Nyvlt (1968), to generate solubility data for aluminium potassium sulphate in water

and used FBRM to monitor the point of dissolution. They chose counts between 50μm

and 250μm to monitor the dissolution as large particles have a large surface area to

volume ratio, and hence dissolve at a slower rate, compared to smaller crystals. The

results obtained were in excellent agreement with other published data, Mullin et al.,

(1965). In addition to solubility data, metastable zone width information was

generated for the same system. FBRM was used to detect the onset of nucleation

when solutions of aluminium potassium sulphate were cooled at various rates. The

number of counts between 0 and 20 µm was used to detect the onset of nucleation.

Since the instrument was set to measure every 10 s this size range was chosen since

published data reported that aluminium potassium sulphate may grow up to 15 µm in

10 s. Fujiwara et al., (2002) performed similar work, on an aqueous paracetamol

system, but compared the three techniques for the measurement of the MSZW,

FBRM, ATR-FTIR and visual observation. FBRM proved to be the most effective

technique as it reliably identified the onset of nucleation earlier than the other

methods. Various weightings on the chord length distributions were examined to see

which provided the best data. A 1/length weighting was tested to concentrate on the

small particles but increased noise and drift of the background signal was observed. A

square weighting was also tested but resulted in a slightly delayed detection. The most

reliable response was from the unweighted chord length distribution and this

parameter was chosen to identify the onset of nucleation. Liotta and Sabesan (2004)

Page 60: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

50

showed that this technique could be automated, reducing the time and manpower

associated with measuring solubility and MSZW information. To do this required a

link that enabled the crystallizer to respond in real time to the FBRM data. A solution

of high concentration was held at a temperature above the solubility point and then

cooled until crystals nucleated. The FBRM detected the onset of nucleation (point on

the metastable zone) and when the total counts exceeded a certain threshold the

cooling ramp was terminated and a heating ramp was initiated. This heating ramp

continues until the total counts fall to the baseline (solubility point). At this point the

solution is held at an elevated temperature and a solvent is added to dilute the system.

The process is then repeated to gather further solubility and metastable zone width

data. An automated technique such as this significantly reduces the time needed to

generate solubility and metastable zone width data.

2.2.7.2 Nucleation Kinetics

FBRM is very useful for tracking nucleation as it can focus on specific size ranges.

By focussing on the smallest chords and trending them over time it is possible to

characterize nucleation behaviour. This is also facilitated by the sensitivity FBRM

displays to small particles. Shi et al., (2003) used FBRM to monitor the evaporative

crystallization of burkeite. Using FBRM trends the crystallization process was divided

into three regimes each dependant on the dominant mechanism of nucleation

involved. Unusual nucleation behaviour was observed and the inhibitory effect of

specific impurities on the nucleation of burkeite was identified. Monnier et al., (1997)

used the Par Tec instrument to calculate nucleation rates for the batch crystallization

of adipic acid in water. Global nucleation rates were calculated using Par Tec and by

Page 61: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

51

subtracting the primary nucleation rates from this it was possible to calculate

secondary nucleation parameters. This information was then used in a model that

predicted the final size distribution of the adipic acid crystals. The PSD calculated

using the model compared well image analysis.

Nucleation kinetics can also be inferred from MSZW information gathered using

FBRM. Barrett & Glennon (2002) used such information to calculate nucleation

kinetics for the aluminium potassium sulphate. The results agreed with literature data

2.2.7.2 Monitoring Crystal Size

Even though FBRM does not report the true particle size it is sensitive to the size,

shape and population of the crystal slurry. FBRM is useful for comparing these

parameters under different process conditions. Loan et al., (2002) showed this by

comparing ferrihydrite precipitation at two different pH values. They found that at

low pH smaller crystals were formed and that poor filtration characteristics for the

low pH process were as a result of the unsuitable particle size distribution. FBRM was

used for a similar purpose by Fujiwara et al., (2002) for the study of the aqueous

batch cooling crystallization of paracetamol. FBRM indicated an increase in the

number of crystals during a run where supersaturation was high due to nucleation. For

a second batch where supersaturation was kept at a constant low level there was no

increase in the number of crystals indicating the absence of nucleation.

FBRM has an advantage over off-line techniques die the fact that sampling is not

necessary. Loan et al. (2002) noted that the FBRM chord length measurement was a

superior indicator of the particle characteristics, compared with scanning electron

Page 62: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

52

microscopy, since there was no sample removal, dilution or drying needed. A similar

advantage was noted by Abbas et al., (2002) in that FBRM was the only technique

suitable for reactive crystallization, since sampling was not needed. (Off-line

sampling of the nickel hydroxide particle resulted in the alteration of the

physiochemical structure).

Another advantage FBRM has over other techniques is that it can measure a lot of

data in real time. This allows dynamic crystallization processes such as growth,

breakage and agglomeration to be studied in real time. By focussing on size ranges of

interest it is possible to identify these crystallization mechanisms. For example,

breakage is identified by a decrease in the number of chords counted in large size

ranges, and an increase in the number of chords counted in small size ranges, as well

as a decrease in the mean chord length. Such analysis can be facilitated through the

use of PVM to identify particle morphology as well as differentiate between

mechanisms that are sometimes confused, such as growth and agglomeration. Abbas

et al., (2002) used FBRM to monitor the reaction crystallization of nickel hydroxide

and the cooling crystallization of ammonium sulphate. In each case FBRM monitored

the size of the crystals over the course of the batch. For the reactive crystallization an

initial increase in particle size was identified as agglomeration and the subsequent

reduction in particle size was identified as the breakage of these agglomerates due to

the action of the agitator. For the cooling crystallization the size of the particles

increased steadily over the course of the batch. Kougoulous et al., (2005b) used

FBRM to monitor the cooling crystallization of an organic fine chemical in a MSMPR

crystallizer. Upon heating a decrease in particle counts was observed indicating

dissolution of the organic fine chemical. Upon cooling counts increased indicating

Page 63: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

53

nucleation. After a certain period of time the trends reached steady state and a re-

circulation loop was initiated. The FBRM trends remained steady indicating breakage

due to the action of the peristaltic pump was not an issue (Kougoulos 2005a). After a

certain period a secondary loop was opened and the MSMPR cooling crystallization

process commenced. Crystals were removed from the system and FBRM successfully

identified the size range of crystals removed. There was a significant decrease in

coarse crystals and a significant increase in fine crystals. This indicated that large

particles were being removed while the nucleation and growth of finer particles

became more evident. Complementary video images provided by process video

imaging (PVI) confirmed that this was the case. Steady state was identified after about

6-9 residence times by identifying the point at which the FBRM trends flat-lined. The

same researchers used FBRM to monitor particulate attrition and breakage of an

organic fine chemical in a turbulently agitated system Kougoulos et al., (2005c). A

dilute solution of the organic fine chemical in isopropyl alcohol and water was

agitated for 150 min at 300rpm. The impeller frequency was then increased to 400

rpm for 150 min and then to 500 rpm for a further 150 min. An FBRM probe placed

in the vessel monitored the chord length distribution over the course of the

experiment. The FBRM data indicated that that the number of fine particles (0-10 µm)

increased and the number of coarse particles (100-300 µm) decreased. The number of

intermediate particles (50-100 µm) initially increased and then decreased. FBRM was

successful in monitoring the death and birth of crystals by disruption for different size

classes and provided evidence of micro-attrition effects. The advantage of FBRM,

over other size measurement techniques for monitoring such a process was noted.

This was because FBRM can monitor, in situ, the crystal size distribution for different

particle size classes (fine, intermediate, coarse). Information on particle attrition and

Page 64: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

54

disruption proved extremely useful for the prediction of particle behaviour upon scale

up.

The previous examples have shown some of the utility FBRM has for the design,

development and optimization of crystallization processes. In each case the chord

length distribution was sufficient to characterize the system and there was no need to

assume a shape or convert to a crystal size distribution. However, this can be achieved

using some of the modelling work outlined in the previous section. Worlitschek and

Mazzotti (2004) used FBRM, coupled with a model-based method to calculate

particle size distribution from the raw chord length data (Worlitschek et al., 2005), to

monitor particle size in the batch cooling crystallization of paracetamol in ethanol. An

optimal temperature trajectory that minimised the difference between the final PSD

(measured using FBRM) and a given optimal monomodal PSD was calculated using a

non-linear constrained minimization algorithm. The size distributions of the crystals

produced using optimal cooling profile and a simple linear cooling profile were

compared. The FBRM data combined with the CLD-PSD model showed that the

linear cooling profile produced a significant amount of small crystals while the

optimal cooling profile provided a PSD close to the desired monomodal distribution.

2.2.7.3 Temperature Cycling

Temperature cycling is a useful technique used to increase crystal size and reduce the

number of fines at the end of a batch. FBRM is useful here because the cycling

regime (number of cycles, cycle duration temperature change) may be optimised by

focussing on the removal of small chords and the increase in the mean chord length.

Page 65: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

55

Doki et al., (2004) applied this technique from the start of a crystallization to ensure a

large final product crystal size and minimise the number of fine crystals. FBRM was

used to monitor the number of crystals formed in the aqueous batch cooling

crystallization of glycine. The goal of the research was to produce crystals of a

suitable particle size distribution and of the correct polymorphic form by alternating

the temperature profile and the termination temperature. Initial work showed that

there was a linear correlation between the number of crystals of glycine suspended in

a saturated solution and the number of counts measured by the FBRM. This

correlation was applicable for a number of size ranges. During the batch experiments

seed crystals were added at the saturation temperature and FBRM monitored the

number of crystals in solution as cooling progressed. When the number of crystals in

solution exceeded a threshold value (due to nucleation) cooling was terminated and a

heating step was initiated to dissolve the fine crystals. When the number of crystals

measured returned to the original value the cooling step was restarted. In this way

large particles of suitable size distribution were formed.

1.2.7.4 Polymorphic Transitions

In certain cases different polymorphs have different morphologies. Since FBRM is

sensitive to changes in shape for these certain instances it can be used to track

polymorphic transitions. O‟ Sullivan et al., (2003) used FBRM to monitor the

polymorphic transition of metastable δ-form mannitol to the stable β-form. A small

amount of δ-form mannitol was added to a saturated solution of β-form mannitol in

water. FBRM trends were used to track the appearance of the δ -form as it was

charged, the dissolution of the δ -form and the subsequent growth and nucleation of

Page 66: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

56

the β –form. This δ-β transition was corroborated using off-line X-Ray Powder

Diffraction (XRPD) and differential scanning calorimetry (DSC) as well as in process

imagery provided by PVM.

2.2.7.5 Effect of Impurities

FBRM is useful for measuring the rate of crystallization. Impurities often effect the

crystallization and there impact can be assessed using FBRM. Scott & Black (2005)

used FBRM to monitor the effect of impurities on the crystallization of urea and an

in-house system. FBRM successfully identified a difference in the crystal growth rate

when impurities were absent and present. Crystallization in the presence of the

impurity was 7-times slower for the urea system and ~20-times slower for the in-

house system. The work identified the potential use of FBRM in the plant for the

monitoring of impurities. In commercial manufacture impurity levels often change

during process development and on scale-up. This can affect morphology and thus

downstream processing operations. Additionally if impurity levels increase but the

crystallization time remains the same the yield can be affected due to the retarded

growth rate. Alternatively, if the impurity level decreases and the crystallization time

remains the same, crystallization may be complete long before the end of the batch

resulting in a decrease in productivity. This can be avoided by in-process tests or by

the use of in-line technology.

Page 67: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

57

2.3 PROCESS VIDEO MICROSCOPE (PVM)

2.3.1 Introduction

Imaging can provide useful information on the size and morphology of particles.

These variables are important in any particulate system. In a crystallization system the

size and shape of the crystals can influence downstream processing operations such as

filtration, drying and formulation. For example, fine particles have a large surface

area and can take a long time to dry. Large flat crystals may pack on top of each other

leading to long filtration times. In polymorphic systems imaging can be extremely

useful as it is often possible to distinguish between two polymorphic forms according

to their shape.

The traditional imaging method used to gather information on size, and more

specifically shape, has been off-line microscopy. Microscopy is a useful and relatively

cheap method to gather data but there are some problems associated with it. To gather

statistically robust data on the size of the particles hundreds of images must be taken

and processed. In addition obtaining a representative sample can be difficult and is

time consuming and labour intensive. Once a sample is isolated the slurry must be

dried and prepared for analysis. This can alter the particles and lead to data that is not

representative of what is actually going on in the process. For example brittle particles

can break upon isolation, generating a large umber of fine particles. This may lead

one to conclude that the process is producing the fine particles rather than the

sampling method.

Page 68: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

58

For a crystallization system there are some additional problems. Sampling is

undesirable because once a sample is taken changes to the crystals can occur due to

further crystallization phenomena e.g. nucleation, growth, breakage agglomeration

dissolution or polymorphic transformation. In some situations it may not be possible

to isolate a sample at all due to process or safety constraints.

The problems associated with off line imaging can be avoided by imaging within the

process environment. The Particle Vision and Measurement (PVM) (Figure 1.6)

system is a probe-based, high-resolution video microscope that can perform such a

task. PVM uses six independent lasers of wavelength 905 µm to illuminate the

particulate system. The lasers‟ source is located at the back of the probe and the laser

light is sent to the probe tip via a fibre optic connection. At the tip there are six lenses

arranged hexagonally, each of which focus the light from one laser onto an area with a

fixed area of approximately 2 mm2. A micrometer positioned at the back of the probe

can focus this illuminated area into or out of the system under investigation. As

particles intercept the illuminated area light is reflected in all directions and some of it

is reflected back up the probe. The backscattered light passes through an objective

lens, focussing it onto a light sensitive charge coupled device (CCD) that produces the

image. A recording of this image is made using a CCD camera at the back of the

probe.

The PVM software allows 10 images per second to be recorded. Settings such as the

gain and the offset of the CCD can be changed using the software to optimise the

image. If the material under investigation backscatters too much light the laser

intensity can be reduced. A useful feature of the software is that the laser intensity,

Page 69: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

59

gain and offset can be altered automatically at specified time intervals to optimise the

images obtained in a dynamic process, such as crystallization. This means the probe

can be left unattended while images of sufficient quality are recorded.

A typical PVM unit consists of a probe, a field unit (which contains the electronics)

and a computer for image storage and review. An FBRM 700L was used for this

research and is 25 mm in diameter, has a wetted length of 300 mm and is made from

stainless steel. The window of the probe is made from sapphire and a kalrez o-ring is

used to seal the window to the probe tip. A more detailed description of the mode of

operation of PVM is available elsewhere (Barrett (2002)).

PVM is an extremely useful tool in for the study of particulate systems as it can

provide qualitative information on the size and shape of the particles in situ and in

real time (Figure 1.7). Furthermore analysis of a sufficient number of PVM images

can provide qualitative information on the size of the particles being studied.

Additionally when PVM is used in conjunction with the quantitive information that

the FBRM can provide, it is possible to achieve an excellent understanding of the

particulate system under investigation.

Page 70: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

60

Figure 2.6: PVM internals (Mettler Toledo User‟s Site)

Figure 2.7: PVM images

Page 71: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

61

2.3.2 Characterization of Particulate Systems Using PVM

O‟ Rourke & MacLoughlin (2005) PVM to analyse the evolving droplet size

distributions in lean silicone oil-water dispersions in an agitated system of standard

geometry. A satisfactory level of agreement between results obtained using PVM and

the more traditional sampling method was achieved. PVM held an advantage over the

sampling method in that it was less labour intensive and was suitable when there was

a large density difference between the two phases. However a minimum of three

minutes was required to acquire the large number of images needed for a

representative sample. This made PVM unsuitable form monitoring very rapid

changes in the droplet size distribution.

Barrett and Glennon (2002) used PVM to monitor the crystallization and dissolution

of potassium aluminium sulphate. PVM was used to validate the rapid increase in the

number of coarse crystals identified by the FBRM as the solution was cooled. PVM

images highlighted the presence of fine crystals, but also the presence of large single

crystals and agglomerates thus validating the FBRM observation. PVM images taken

during the heating period showed the complex nature of the dissolution process. A

clear transition between the distinctive octahedral shape of the crystals to a more

rounded morphology was observed.

McDonald et al., (2001) used PVM to study rice, tobacco and wild Chinese cucumber

cell suspensions. The PVM clearly identifies the different morphologies of the

different suspensions without the need to sample.

Page 72: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

62

2.4 ATTENUATED TOTAL REFLECTANCE FOURIER TRANSFORM

INFRARED SPECTROSCOPY (ATR-FTIR)

2.4.1 Introduction

Supersaturation is a non-equilibrium state where there is more solute dissolved in

solution than in the equilibrium state at a given temperature. A solute will remain in

solution until there is a sufficient level of supersaturation to induce crystal formation.

Supersaturation is the driving force for subsequent crystal nucleation, growth and

agglomeration. As such it impacts on the final crystal size distribution. By measuring

supersaturation the underlying mechanism that governs the final crystal size

distribution can be controlled. In doing so the opportunity to produce crystals of a

suitable size, number and morphology is afforded. A method for the direct

measurement of supersaturation has been proposed by Löfflemann and Mersmann,

(2002), however the most common technique for measuring supersaturation is to

measure the solution concentration and infer the prevailing supersaturation from

solubility data. In this case it is more accurate to say that supersaturation is being

monitored, rather than measured.

Attenuated Total Reflection Fourier Transform Infrared (ATR-FTIR) spectroscopy

has emerged as a useful technique for the measurement of solution concentration in

crystallizations. It has an advantage over other concentration measurement

techniques, as sampling is not required. This eliminates the problems associated with

temperature-controlled external sampling loops and removes the need for phase

separation devices. Additionally, multiple solute and multiple solvent crystallizations

Page 73: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

63

can be characterised. This can be useful for anti-solvent and reactive crystallization as

well as crystallizations where impurities are present. With accurate solubility data the

measured solution concentration can be used to infer the prevailing level of

supersaturation.

Briefly, ATR-FTIR spectroscopy measures solution concentration by irradiating the

solution with infrared light to produce an infrared spectrum. This spectrum is

characteristic of the vibrational structure of the substance in immediate contact with

the ATR probe and can be described as a unique “fingerprint” of the liquid phase. An

ATR crystal is chosen so the depth of penetration of the infrared energy is smaller

than the liquid phase barrier between the probe and the solid crystal particles. Hence

when the probe is inserted into a crystal slurry the probe should be in immediate

contact with the liquid phase and interference from the solid phase should be

negligible. Despite this assumption some researchers have noted interference of the

spectrum by the solid phase in crystallization systems (O‟ Sullivan, 2005). The liquid

phase concentration is a function of the infrared spectrum generated and a calibration

model is used to inter-relate the data. . A more detailed description of spectroscopy in

general and the mode of operation of the ATR-FTIR probe is available elsewhere,

(Dunuwilla et al., 1994; Togkalidou et al., 2001; Lewiner et al., 2001a; O‟ Sullivan,

2005).

2.4.2 Initial Work Using Various Calibration Models

Initial studies using ATR-FTIR to measure solution concentration provide a useful

insight into the evolution of the technique, the problems associated with using it and

Page 74: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

64

perhaps most usefully the various calibration models that relate the infrared spectrum

to the solution concentration.

Dunuwilla et al., (1994) were one of the first researchers to examine the use of ATR-

FTIR for the measurement of solution concentration. The solubility of citric acid in

aqueous solution was measured by taking a sample from a crystallization vessel and

placing it in a micro circle open boat cell equipped with an ATR rod. Relatively good

agreement between the measured data and literature values was achieved. However at

high temperatures the experimental solubility value was overestimated. Subsequent

research by Grön & Roberts (1999) repeated this work in situ and showed results with

increased accuracy at all temperatures. This indicates that the sampling method used

by Dunuwilla et al., (1994) may have been poor.

Dunuwilla & Berglund (1997) were the first researchers to use the ATR-FTIR

technique to monitor supersaturation in situ. The crystallization of maleic acid from

aqueous solutions was examined using a parabolic, intermediate parabolic and linear

cooling profile. The supersaturation was monitored for each of the cooling profiles

and the final crystal size distribution was measured at the end of each batch by

sieving. The ATR-FTIR successfully monitored the supersaturation in each case.

Supersaturation for the parabolic cooling profile remained low and constant

throughout the batch and large crystals were formed. For the linear cooling profile

supersaturation was higher over the course of the crystallization resulting in an

average crystal size that was half that produced using the parabolic cooling profile.

Page 75: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

65

In this work supersaturation, solubility and the metastable zone width were

represented in terms of the ratio of the transmittance of two maleic acid peaks,

however a calibration between the transmittance ratio, concentration and temperature

was supplied and the solubility data generated was within 3% of literature values.

Lewiner et al., (2001a) used ATR-FTIR to study the crystallization of three fine

chemical products. For this study a different calibration model was used that

calibrated the height of one relevant solute peak with the concentration of the solute

and the temperature. The crystallization of Bifenox (a weed killer) from methanol

solutions by cooling was examined. Combining supersaturation data with information

on the size of the product crystals (measured by SEM and laser diffraction) it was

concluded that the final product size was mainly controlled by agglomeration and

secondary nucleation. The second system studied was the cooling crystallization of

another weed killer, isoproturon (IPU), from ethanol solutions. The supersaturation

measurements combined with information on the final crystal size distribution

gathered using image analysis indicated a different crystallization behaviour. In this

case there were large variations in the metastable zone width and low levels of

supersaturation due to very fast crystal growth. Finally the cooling crystallization of

an active pharmaceutical ingredient, compound F was studied. Compound F exhibited

four polymorphic forms and the ATR-FTIR successfully monitored the transition

from one form to the other over the course of a cooling crystallization.

Togkalidou et al., (2001) combined ATR-FRIR spectroscopy with robust

chemometrics to produce a very accurate calibration model for the estimation of the

solubility of potassium dihydrogen phosphate (KDP) in water. A number of

Page 76: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

66

chemometric models were applied and the one that gave the most accurate predictions

was selected. In this case the most accurate model arose from choosing a whole

region of the spectrum rather than a single peak or a number of peaks. This improved

accuracy may have been due to a significant shift in one of the peaks meaning more

accurate predictions were made when all the data in the region was used or, it may

have been due to an increase in the apparent signal-to-noise ratio resulting from the

random noise in the absorbance being averaged over many more frequencies. The

solubility of KDP in water was measured using a number of models and the results

were compared. The solubility data gathered using a single peak chemometric model

proved inaccurate, while the data produced using the whole spectral region proved

this most accurate. This indicates that multiple absorbances and chemometrics

produce the most accurate calibration models. The most accurate model produced

concentration estimates with and accuracy of ± 0.12 wt%.

The research outlined above outline three different methods for the calibration of the

infrared spectrum with concentration. Lewiner et al., (2001b) commented on the

accuracy of the data presented by Togkalidou et al., (2001), describing it as

“remarkable”. However they noted that in industry there is a reluctance to use, so

called chemometric methods since a calibration using well-defined absorption bands

has a clearer chemical and physical meaning. However many of the ATR-FTIR

probes available for commercial use include useful chemometric packages that can be

used simply and effectively to generate an accurate calibration model. For the purpose

of this study the chemometric software package QuantIR, supplied with the Mettler

Toledo ReactIR 4000 ATR-FTIR probe will be used to generate calibration models.

Page 77: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

67

2.4.3 The Use of ATR-FTIR to Characterize Crystallization Processes

2.4.3.1 Seeding

Lewiner et al., (2001b) continued previous studies of the cooling crystallization of

isoproturon (IPU) from ethanol solutions using a calibration model based on

absorption at a single peak. Unseeded crystallization resulted in large variations in the

metastable zone width, to the extent that no identifiable metastable zone width could

be identified at a constant cooling rate. This variability resulted in a highly variable

final crystal size. Supersaturation measurements indicated that the point of nucleation

was followed by a dramatic decrease in the supersaturation regardless of the cooling

rate used, indicating that supersaturation was minimal during the growth period. By

introducing a seeded crystallization the variability in the metastable zone width was

reduced and the final mean crystal size was increased and the CSD narrowed. A

suitable seeding temperature was identified by monitoring the supersaturation for runs

where the seed was introduced at different temperatures. When the seed was

introduced close to the solubility curve supersaturation continued to increase until

there was a burst of secondary nucleation. When the seed was introduced close to the

metastable zone width a rapid decrease in the supersaturation was observed similar to

the unseeded case indicating the introduction of the seeds was not effective. An

intermediate seeding temperature proved the most effective as the final crystal size

increased and the coefficient of variation decreased. With a suitable seeding

temperature identified it was possible to assess the effect of the seed mass and the

cooling rate after seeding. Supersaturation was measured for different cooling rates

after seeding. Higher cooling rates led to increased supersaturation and hence a higher

nucleation rate that was confirmed by a decreased final product size. With a suitable

Page 78: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

68

cooling rate identified the effect of the seed mass was assessed. An increase in the

mass of seeds added resulted in a decrease in the prevailing level of supersaturation

due to the increased surface area available for growth.

2.4.3.2 Oiling Out

Groen & Roberts, (2001) combined ATR-FTIR supersaturation measurements with

optical turbidity measurements to identify liquid-liquid phase separation (also known

as oiling out) in the cooling crystallization of citric acid. Batch cooling experiments

were conducted and a reduction in optical transmittance was observed a significant

length of time prior to the onset of nucleation measured by a reduction in the

measured supersaturation. This change in the optical properties of the system without

crystallization indicated the formation of non-crystalline phase. Oiling out was

suggested as a possible explanation for the phenomenon as the supersaturation level

was extremely high and previous researchers had noted similar results for the citric

acid water system. The actual reduction in the solute concentration due to the phase

separation was estimated to be about 1 % (w/w), enough to facilitate nucleation. This

value was similar to the peak-to-peak noise of the ATR-FTIR signal. Hence no

change in the supersaturation was observed.

Page 79: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

69

CHAPTER 3: SOLUBILITY MEASUREMENT FOR AN ANTI-

SOLVENT SYSTEM USING GRAVIMETRIC ANALSYSIS, ATR-

FTIR AND FBRM

3.1 ABSTRACT

The solubility of benzoic acid in ethanol-water mixtures was measured using

gravimetric analysis (solid analysis and liquid analysis), Focused Beam Reflectance

Measurement (FBRM) and Attenuated Total Reflectance – Fourier Transform Infra

Red Spectroscopy (ATR-FTIR). A suitable method that describes the solubility curve

for an anti-solvent addition system is presented. By expressing the solubility on an

anti-solvent free basis, the dilution effect may be eliminated and expression of the

metastable zone width (MSZW) and supersaturation is simplified. The FBRM, ATR-

FTIR and solid analysis methods produced similar solubility data validating the

techniques and highlighting the accuracy of the data. The liquid analysis

underestimated the solubility. This may have been due to the esterification of the

benzoic acid whilst drying. A correlation between the measured data and the

predictions of the UNIQUAC liquid activity coefficient model is also presented.

3.2 INTRODUCTION

Key to the successful design and optimisation of a batch crystallization process is

knowledge of the solubility over the range of operating conditions likely to be

encountered during the batch. In processes involving the crystallization of a solute

through the addition of an anti-solvent to a saturated, or near saturated, solution the

Page 80: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

70

solute solubility should be known over a wide range of solvent mixture

concentrations. Gravimetric analysis is the traditional standard method for the

determination of solubility (Mullin, 1993) and its use has been widely reported for a

range of anti-solvent systems (e.g. Hoijati & Rohani, 2006a; Granberg and Rasmuson,

2000). Perhaps its principle drawback is the time-consuming nature of the method.

The use of in ATR-FTIR has become an increasingly popular technique for

investigation of solubility characteristics in crystallization systems including anti-

solvent processes (Hoijati & Rohani, 2006a). ATR-FTIR requires a substantial effort

to initially calibrate the system, but having developed the calibration data set,

subsequent application is relatively quick. Investigations based on the Focussed Beam

Reflectance Measurement (FBRM) probe have generally involved cooling

crystallization processes through the application of an automated polythermal method

for solubility determination (Barrett and Glennon, 2002). FBRM-based methods have

the advantage of requiring no calibration and of rapid measurement duration.

However, measurement of solubility in anti-solvent systems is not immediately

amenable to automation.

A priori prediction of solubility has been less successfully demonstrated to date,

although the use of liquid activity coefficient models has facilitated good correlation

of available experimental data (Gracin et al, 2002; Hoijati & Rohani, 2006b).

In this paper, a comparison of these three techniques for the measurement of solubility

in anti-solvent systems is presented, using benzoic acid-ethanol-water as the model

system. The correlation between the measured data and the predictions of the

UNIQUAC liquid activity coefficient model is also presented.

Page 81: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

71

3.3 EXPRESSION OF SOLUBILITY

In an anti-solvent system, supersaturation is generated by a combination of reduced

solubility due to the action of the anti-solvent, and dilution. Typically, solubility is

expressed in terms of the mass of solute per total mass (or volume) of solvent and

anti-solvent. The combination of these factors makes expression of key crystallization

variables, specifically supersaturation and the metastable zone width, difficult. To

overcome this problem researchers have generally chosen to depict the path of anti-

solvent addition on the solubility curve as a diagonal line that accounts for the dilution

and reduced solubility (Granberg et al., 2001; Pina et al., 2001; Takiyama et al., 1998;

Kaneko et al., 2002). Sometimes dubbed a “mixing line” this diagonal line represents

the apparent concentration of the solution once the solution and anti-solvent have

been mixed, in the absence of crystallization. In this case expression of the

supersaturation and metastable zone width becomes potentially complex.

Figure 3.1 is a traditional solubility curve with the solubility expressed in terms of the

total mass of solvent and anti-solvent. A diagonal line on the solubility curve

represents the change in concentration as anti-solvent is added. Expression of the

supersaturation generation rate is, for such systems, cumbersome (Pina et al., 2001).

Page 82: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

72

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 20 40 60 80

anti-solvent %

g s

olu

te /

(g

so

lven

t + a

nti

-so

lven

t)

Solubility Data

Solute Concentration

Figure 3.1: Change in concentration, as anti-solvent is added.

Solubility is generally given as a function of anti-solvent concentration, i.e. cs = cs(m).

Therefore,

s sdc dc dm

dt dm dt Eq. 3.1

When concentrations are reported in terms of g/g solution, and a constant anti-solvent

addition rate is used:

a Rt

m ts w a Rt

Eq. 3.2

2

R s wdm

dt s w a Rt Eq. 3.3

s

c ts w a Rt

Eq. 3.4

2

dc sR

dt s w a Rt Eq. 3.5

Page 83: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

73

Therefore, the supersaturation generation rate is given by

s

s

2

dcsR R s w

d c dc dc dm

dt dt dt s w a Rt Eq. 3.6

When concentrations are reported in terms of g/g solvent (i.e. on an anti-solvent-free

basis):

a Rt

m tw

Eq. 3.7

dm R

dt w Eq. 3.8

s

c tw

Eq. 3.9

dc

0dt

Eq. 3.10

Therefore, the supersaturation generation rate is given by

sd c dc R

dt dm w Eq. 3.11

By defining all concentrations on an anti-solvent free basis, the solute concentration

follows a horizontal path as anti-solvent is added and is analogous to cooling on a

temperature-concentration curve (3.2). In this way the metastable zone width is

clearly defined as the difference in anti-solvent concentration at the point of

nucleation and at the point of saturation. The supersaturation is the difference (or

ratio) between the actual concentration and the equilibrium concentration.

Page 84: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

74

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 1 2 3 4 5

g anti-solvent / g solvent

g s

olu

te /

g s

olv

en

t

Solubility Data

Solute Concentration

Figure 3.2: Change in concentration (anti-solvent free basis), as anti-solvent is added.

3.4 EXPERIMENTAL WORK AND ANALYSIS

3.4.1 Gravimetric Analysis

An ethanol-water mixture of known composition is charged to a 500ml glass jacketed

vessel and the temperature allowed to equilibrate at 25 C. A known excess of benzoic

acid is added making the total solution mass about 150 g. The solution is then stirred

at 300rpm for 2 hours. Ideally the solution should be left for 24 hours, however

experiments conducted using the ATR-FTIR indicate that equilibrium is reached after

as little as 10 minutes. At this point the slurry is filtered and the filtrate and filter cake

placed in separate conical flasks and covered with filter paper. The filter paper allows

MSZW (g/g)

ΔC (g/g)

Page 85: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

75

the evaporation of the solvent but prevents the apparent sublimation of the solute, a

phenomenon observed during initial runs. The conical flasks is placed in an oven at

80 C and left for a number of hours. The mass of the filtrate and filter cake are

measured at regular intervals to identify when drying is complete. When three

consecutive measurements yield three similar masses the sample is deemed to be dry.

This typically occurred after 24 hrs for the filter cake and 60 hours for the filtrate

(Figure 3.3).

Two methods were used to determine the solubility. Firstly, the mass of solute

remaining once all of the solvent from the filtrate had evaporated provided „liquid

analysis‟ solubility. Secondly, the mass of undissolved solute remaining in the filter

cake once it had been dried was subtracted from the initial charge of solute to the

vessel to provide „solid analysis‟ solubility (3.4).

It is clear that there is a systematic difference between the solubility values generated

from the solid analysis and the liquid analysis. The liquid analysis method appears to

underestimate the solubility over all the water concentrations. There are a number of

possible reasons for this. It was difficult to transfer all of the slurry from the

crystallizer into the filter. Every effort was made to ensure all the saturated solution

and solute was transferred but inevitably some was left in the crystallizer. It is also

possible that there may have been an esterification reaction between the benzoic acid

and the alcohol at the elevated temperatures in the oven. It is possible that an ester

formed and evaporated with the rest of the solvent mixture. By drying samples at a

low temperature in a vacuum oven it may be possible to avoid this problem.

Page 86: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

76

0

40

80

120

160

0 20 40 60 80 100 120 140 160

Time (hrs)

Fil

trate

Mass (

g)

0

2

4

6

8

10

12

14

16

Filte

r C

ak

e M

as

s (

g)

Filtrate

Filter Cake

3.3: Typical drying curve for gravimetric analysis run

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

(g water/ g ethanol)

(g b

en

zoic

acid

/ g e

thano

l)

Solid Analysis

Liquid Analysis

Figure 3.4: Solubility of benzoic acid in ethanol water at 25 C – Gravimetric Analysis

Page 87: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

77

3.4.2 Polythermal Method Using FBRM

An ethanol-water mixture of known composition, and a mass of about 1200g, is added

to a 2-litre stainless steel jacketed vessel fitted with a Julabo chiller, to provide

temperature control, and a condenser to prevent evaporation. A known mass of

benzoic acid is charged to the vessel and the solution heated slowly. The slurry is

monitored using FBRM and the dissolution temperature noted. At this point the

solution is cooled to below the saturation temperature and another known mass of

benzoic acid is charged. The slurry is reheated and the point of dissolution noted

giving a new saturation temperature. In this way the solubility of benzoic acid in a

given mixture of ethanol and water is calculated for a range of temperatures close to

25 C and the actual solubility at 25 C can be interpolated. Once a sufficient number

of data points are gathered to allow interpolation more ethanol is added and the

process repeated so the solubility for another concentration can be calculated. In this

way the solubility for a full range of ethanol-water compositions at 25 C can be

gathered.

Figure 3.5 is a plot of a typical run for the dissolution of the benzoic acid. Counts

between 100 and 1000 microns are used to identify the point of dissolution. This is

because large particles are the last to dissolve due to their high surface area to volume

ratio. The point at with the counts dropped to zero was clear for all the runs making

identification of the saturation temperature simple.

It is important to assess the effect of heating rate on the dissolution temperature

measured by FBRM. If the heating rate is too high, dissolution may be observed at a

Page 88: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

78

temperature above that of the actual saturation temperature. A number of runs were

carried out to assess the saturation temperature measured at different heating rates

(Figure 3.6). A heating rate above 0.3 C/min is too fast and the saturation temperature

is overestimated. A heating rate of 0.2 C/min was chosen for all the runs since this

rate garnered the same saturation temperature as for a heating rate of 0.1 /min but

allows a shorter batch time.

Figure 3.7 shows the solubility of benzoic acid in four ethanol-water mixtures close to

25 C. As expected the solubility increases with increasing temperature and decreasing

water concentration. By interpolating these curves the solubility at 25 C in various

ethanol-water mixtures can be evaluated (Figure 3.8).

Page 89: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

79

15

20

25

30

180 205 230 255 280

Time (min)

Tem

pera

ture

(°C

)

0

50

100

150

200

250

300

#/s

(100-1

000 m

icro

ns)

Temperature

#/s (100-1000 microns)

Figure 3.5: Nucleation and dissolution of benzoic acid monitored using FBRM

23.2

23.3

23.4

23.5

23.6

0 0.1 0.2 0.3 0.4 0.5

Heating Rate (°C/min)

Sat

ura

tio

n T

em

pera

ture

(°C

)

Figure 3.6: Heating rate vs. saturation temperature

Page 90: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

80

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

15 20 25 30

Temperature (°C)

g b

enzo

ic a

cid/g

eth

an

ol

0.90 (g w ater/g ethanol)

1.37 (g w ater/g ethanol)

2.08 (g w ater/g ethanol)

3.37 (g w ater/g ethanol)

Figure 3.7: Temperature vs. Solubility in various water concentrations

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Water Concentration (g water/ g ethanol)

So

lub

ilit

y (

g b

en

zo

ic a

cid

/ g

eth

an

ol)

Figure 3.8: Solubility of benzoic acid in ethanol water mixtures measured using FBRM

Page 91: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

81

3.4.3 Solubility Measurement Using ATR-FTIR Technique

3.4.3.1 Calibration of Probe

ATR-FTIR spectroscopy can be used to measure the concentration of a dissolved

solute. The IR absorbance measured by the probe (Mettler Toledo, React IR 4000) is

a function of concentration of the dissolved solute and a calibration is needed to

convert the absorbance data into concentration. Operation at a constant temperature of

25 C eliminated any temperature effect on the absorbance.

The method of partial least squares (PLS) has been commonly used to perform such

calibrations. Although the use of PLS does not produce a physical understanding of

the system, it is well suited for the primary goal of prediction and estimation (Liotta &

Sabesan 2004). A PLS model developed by Mettler-Toledo (Quant- IR) was used to

perform the calibration. A region to two-point baseline model was used as this

produced the lowest error. The spectral region used was between wave numbers 1003

and 1814 cm-1

(Figure 3.9).

40 standards were used to calibrate the ATR-FTIR probe and 12 standards were used

to validate the model (Figure 3.10). Spectra for all the standards were taken in a

500ml glass jacketed vessel at a temperature of 25 C. A concerted effort was made to

get standards inside the metastable zone. However the system exhibits a very narrow

metastable zone width and often nucleation would occur before a measurement could

be taken. This was especially the case at low water concentration where the solubility

curve is steep and high supersaturation is generated for a small increase in water

Page 92: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

82

concentration. It was relatively easy to gather standards inside the metastable zone

width at higher water concentrations where the solubility curve is less steep.

0

0.2

0.4

0.6

0.8

1

1.2

100011001200130014001500160017001800

Wavenumber (cm -1)

Abso

rbance

Figure 3.9 Spectral region chosen for quant

Figure 3.10: Standards used to calibrate and validate the ATR-FTIR probe. Weak validation points

arrowed.

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

(g water/ g ethanol)

(g b

en

zoic

acid

/ g

eth

an

ol)

Calibration Points

Validation Points

Page 93: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

83

Figure 3.11 shows the comparison between the known and predicted water

concentration calculated using the PLS model. Despite the excellent correlation

coefficient, there is a slight offset of about 4% between the known and predicted

concentration. Closer investigation of the validation set indicates that this offset is due

to an unusually high error for two of the low water concentration experiments

(indicated in Figure 3.10). For these validation points the error is 5.2% and 5.0%,

while for the other 10 validation points the error is always less than 3%. If these

values are removed from the validation set the offset is reduced to below 0.1%. It is

possible that the calibration model is less accurate at these low water concentration

values. Inspection of Figure 3.10 indicates that the validation solutions that produce

the largest error are relatively isolated compared to the calibration set. The

relationship between known and predicted benzoic acid concentration is excellent

with no offset (Figure 3.12). This indicates that if there is a weakness in the

calibration model at low water concentrations, the prediction of the benzoic

acid concentration is not affected. The average error was 1.8% of the measured water

concentration and 1.4% of the measured benzoic acid concentration.

Page 94: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

84

y = 0.9642x + 0.0422

R2 = 0.9979

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0

Water Concentration Known (g/g)

Wate

r C

on

centr

atio

n C

alc

ula

ted

(g/g

)

Figure 3.11: Predicted vs. Known water concentration

y = 1.0003x - 0.001

R2 = 0.9993

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.1 0.2 0.3 0.4 0.5

Benzoic Acid Concentration Known (g/g)

Ben

zo

ic A

cid

Co

ncen

trati

on C

alc

ula

ted (

g/g

)

Figure 3.12: Predicted vs. Known benzoic acid concentration

Page 95: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

85

In order to further validate the model a number of experiments were performed under

different conditions and the predicted water concentration and benzoic acid

concentration was compared to the measured values. Table 3.2 compares known and

calculated water and benzoic acid concentrations across the three separate validation

standards. In each case the predicted concentration is within an acceptable margin of

error. Liotta & Sabesan (2004) noted the level of agitation can affect the measurement

of spectra. To assess the effect of agitation intensity, a standard solution was

measured at various levels of agitation. Table 3.2 shows the effect an increase in the

agitation rate has on the water and benzoic acid concentration calculated by the

calibration model. In each case the predicted concentration was acceptable. Since the

crystallization under investigation is an anti-solvent addition and the total volume in

the vessel will increase over time, the effect of total volume on the calibration model

was assessed. Standards of the same composition but of different total volumes were

made up and measured. Table 3.3 shows the difference between the known and

calculated values. Once again the concentrations were within acceptable limits.

Page 96: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

86

Table 3.1: Error across three standards for (a) water concentration and (b) benzoic acid concentration

(a) (b)

Water Concentration (g/g) Benzoic Acid Concentration (g/g)

Standard # Known Calculated Error % Standard # Known Calculated Error %

1 1.292 1.255 -2.90 1 0.188 0.193 +2.55

2 1.292 1.292 0.00 2 0.188 0.189 +0.34

3 1.292 1.263 -2.2 3 0.188 0.186 -0.96

Table 3.2: Effect of agitation on quant: (a) water concentration and (b) benzoic acid concentration

(a) (b)

Water Concentration (g/g) Benzoic Acid Concentration (g/g)

rpm Known Calculated Error % rpm Known Calculated Error %

200 1.292 1.292 0.00 200 0.188 0.189 +0.34

300 1.292 1.277 -1.20 300 0.188 0.187 -0.55

400 1.292 1.271 -1.46 400 0.188 0.187 -0.44

Table 3.3: Effect on volume on quant: (a) water concentration and (b) benzoic acid concentration

(a) (b)

Water Concentration (g/g) Benzoic Acid Concentration (g/g)

Volume (ml) Known Calculated Error % Volume (ml) Known Calculated Error %

110 1.292 1.277 -1.20 110 0.188 0.186 -0.55

220 1.292 1.259 -2.55 220 0.188 0.188 -1.72

330 1.292 1.287 -0.40 330 0.188 0.187 -0.80

Finally, the effect of a hold period on the solution was investigated. The purpose of

this was two-fold. Firstly it gives an indication of the error associated between two

different measurements. Secondly it indicates if evaporation plays a role. A standard

solution was made up and spectra were taken every three minutes for one hour. 3.13

shows the plot of water and benzoic acid concentration against time.

Page 97: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

87

0.0

0.5

1.0

1.5

2.0

2.5

0 10 20 30 40 50 60

Time (min)

wate

r co

ncen

trati

on (

g/g

)

0.0

0.1

0.2

0.3

0.4

0.5

benzo

ic a

cid

con

centr

ati

on

(g/g

)

w ater concentration - calculated

w ater concentration - know n

benzoic acid concentration - calculated

benzoic acid concentration - know n

Figure 3.13: Effect of a hold period on the quant model

The validation experiments combined with the low average error across 12 validation

samples of different concentration indicate that the calibration model is suitable for

the range of concentrations and operating conditions encountered for this study.

3.4.3.2 ATR-FTIR Solubility Measurement

The method applied in this case was similar to that outlined for the gravimetric

analysis but instead of weighing the mass of undissolved solute at the end of the hold

period, a direct measure of the concentration was taken using the ATR-FTIR.

Benzoic acid was ground in a pestle and mortar for a number of minutes to ensure as

small a particle size as possible. An excess of the milled benzoic acid was added to an

ethanol-water mixture, of known composition and mass roughly equal to 150g, that

had been stirring for 30 minutes in a 500ml glass lined, temperature controlled vessel.

Page 98: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

88

The temperature was allowed to equilibrate at 25 C and the slurry was then stirred at

300rpm for 30mins. Spectra were taken every 30s and the calibration model was

applied so the benzoic acid concentration and water concentration could be calculated.

At the end of the 30-minute hold period a known mass of ethanol was added to the

slurry to alter the solubility and the procedure was repeated. This allowed a number of

solubility measurements to be gathered over time. Figure 3.14 shows the results from

one of the solubility determination experiments using ATR-FTIR. Six ethanol

additions were made, allowing seven points on the solubility curve to be measured. A

hold period of 30 mins was deemed sufficient since the concentration of benzoic acid

in solution increased to an equilibrium level, and remained constant at that level, after

as little as 10 mins (Figure 3.14). In viscous solutions and at low temperatures it can

take a number of hours or even days for a system to reach equilibrium. However in

this case it appears that 30mins is sufficient to allow the system to equilibrate. Unlike

increasing temperature, which takes time to increase solubility, the addition of the

solvent (in this case ethanol) is instantaneous which leads to a rapid increase in

solubility. Additionally, the benzoic acid had been milled to a small size prior to

addition. It is possible that this improved its dissolution properties. The water

concentration and benzoic acid concentration at the end of each 30-minute hold period

were used to generate the solubility data. In all 18 points on the solubility curve were

calculated (Figure 3.15).

Page 99: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

89

0

0.5

1

1.5

2

2.5

3

0 2000 4000 6000 8000 10000 12000

Time (s)

Wate

r C

on

centr

atio

n (

g/g

)

0

0.1

0.2

0.3

0.4

0.5

0.6

Ben

zo

ic A

cid

Co

ncen

trati

on (

g/g

)

Water Concentration

Benzoic Acid Concentration

Figure 3.14: Time vs. water concentration and benzoic acid concentration

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Water Concentration (g water/ g ethanol)

So

lub

ilit

y (

g b

en

zoic

acid

/ g

eth

an

ol)

Figure 3.15: Solubility curve of benzoic acid in ethanol-water using ATR-FTIR

Page 100: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

90

3.4.4 Overall Solubility Measurement

The solubility data gathered using the four different methods compares well (Figure

3.16). For the FBRM, ATR-FTIR and solid analysis techniques the solubilities

measured similar data for the solubility. Interestingly, the solubility increases for a

small increase in water concentration at high ethanol concentration. This is not

unusual and solute systems have been shown to exhibit increased solubility in mixed

solvents even if one of them is an anti-solvent (Granberg & Rasmuson, 2000; Hojjati

& Rohani, 2006a).

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00

Water Concentration (g water / g ethanol)

g b

en

zo

ic a

cid

/ g

eth

an

ol

ATR-FTIR Solubility

FBRM Solubiliy

Solid Analysis

Liquid Analysis

Figure 3.16. Solubility of benzoic acid in ethanol-water using four different methods

To establish the thermodynamic consistency of the measured data, the recorded

solubility can be compared with the theoretical predictions of the UNIQUAC model

for liquid activity coefficients (Hoijati & Rohani, 2006b).

At equilibrium,

Page 101: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

91

S L

i i i i if T,P x T,P,x f T,P Eq. 3.12

where

L fus

i m

S

i m

f T, P h T , P 1 1ln

f T, P R T T Eq. 3.13

The UNIQUAC model for prediction of the liquid activity coefficients is given by

c r

i i iln ln ln Eq. 3.14

where

c i i ii i i j j

ji i i

zln ln q ln l x l

x 2 x Eq. 3.15

and

j ijr

i i j ji

j j k kj

k

ln q 1 ln Eq. 3.16

with

i ii

j j

j

x q

x q Eq. 3.17

i ii

j j

j

x r

x r Eq. 3.18

and

i i i i

zl r q r 1

2 Eq. 3.19

The biggest challenge in the use of the UNIQUAC model is the determination of the

binary interaction parameters, ij. To establish the consistency of the measured data

reported here with the model, the binary interaction parameters are estimated by non-

linear regression of the measured data. The sum of squares of the difference between

the experimental and model data was minimized by changing the binary interaction

parameters using SOLVER in Microsoft Excel. The goodness of fit can then be used

Page 102: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

92

as an indication of the consistency. A comparison of the model predictions with the

experimental data is shown in Figure 3.17, using the pure component data given in

Table 3.4 and the regressed binary interaction parameters reported in Table 3.5. As

shown, the UNIQUAC model follows the observed trends, with the measured increase

in the solubility of benzoic acid in ethanol for low water concentrations well

described by the model.

Table 3.4.: Pure component data for UNIQUAC model (Yaws, 2003, Reid et al., 1987)

hfus

(J/mol) Tm (K)

ri

qi

benzoic acid 16,230 395 4.323 3.344

ethanol - - 2.1055 1.972

water - - 0.920 1.400

Table 3.5. Estimated binary interaction parameters for benzoic acid (1) - ethanol (2) - water system.

12 13 21 23 31 32

0.359 8.774 0.000 9.214 0.000 2.465

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 0.5 1 1.5 2 2.5 3 3.5 4

water conc (g/g ethanol)

so

lute

co

nc (

g/g

eth

an

ol)

Figure 3.17. Comparison of measured FBRM solubility data with predictions of fitted UNIQUAC

model.

Page 103: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

93

3.4 DISCUSSION

Three methods for the determination of the solubility of an anti-solvent system have

been presented. Gravimetric analysis provides a cheap and simple method for

estimating the solubility. However this method can be time consuming with the

recommended hold time being up to 24 hours. While the solid analysis provided

accurate data the liquid analysis underestimated the solubility either through the

saturated solution remaining in the crystallizer or an esterification reaction occurring

whilst drying. FBRM provided accurate data and lends itself to automation. While a

turbidity probe would measure the solubility in a similar way the FBRM will provide

further information after the nucleation event and is more sensitive. ATR-FTIR

provided much information about the dissolution process. It was clear that a 30-

minute hold time was sufficient. To gather quantitive data it is necessary to build a

calibration model. While this is time consuming, once it is complete, it is possible to

monitor supersaturation during crystallization by measuring the liquid phase

concentration in situ.

A suitable method to express solubility for an anti-solvent system has been presented

that allows for clear depiction of the metastable zone width and supersaturation in

terms of simple units. The solubility of benzoic acid at 25 C has been measured for a

full range of ethanol-water mixtures. The ATR-FTIR, FBRM and solid analysis

methods all produced similar data while the liquid analysis systematically

underestimated the solubility. The trends observed are in agreement with theoretical

Page 104: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

94

considerations and are similar to trends reported for other ternary anti-solvent

systems.

Page 105: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

95

CHAPTER 4: THE EFFECT OF MIXING ON THE

METASTABLE ZONE WIDTH IN ANTI-SOLVENT

CRYSTALLIZATION

4.1 ABSTRACT

The effects of anti-solvent addition rate and location, and agitation speed on the

metastable zone width (MSZW) of an anti-solvent system were investigated using

Focused Beam Reflectance Measurement (FBRM) and Attenuated Total Reflectance

Fourier Transform Infra-Red Spectroscopy (ATR-FTIR). Benzoic acid in ethanol-

water mixtures, with water acting as the anti-solvent, was chosen as the model system

and was studied at a 500 mL scale.

FBRM proved to be the more sensitive method for the detection of nucleation. In

general, the MSZW widened with increasing addition rate, with the effect most

pronounced when the anti-solvent was added close to the impeller. At this location, an

increase in agitation intensity resulted in a narrower MSZW for all addition rates. For

an addition location close to the vessel wall, the MSZW was narrower and the impact

of addition rate and agitation were less pronounced. Substantial variation in the

MSZW was also observed, with nucleation occasionally occurring at bulk

concentrations less than the saturation level. It is proposed that the MSZW is

influenced by the differing degrees of anti-solvent incorporation at each addition

location. Close to the impeller anti-solvent is rapidly incorporated leading to

consistent results, but, close to the vessel wall, incorporation is hindered by

Page 106: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

96

unfavourable mixing conditions leading to premature nucleation and more

variability*.

Using the MSZW information nucleation kinetics at two different agitation intensities

were estimated. Using this data, an agitation dependent expression for the nucleation

rate was generated.

4.2 INTRODUCTION

The addition of an anti-solvent to a concentrated solution to induce crystallization of

the solute through a reduction in its solubility in the combined solvent system is a

power isolation and purification technique particularly where the temperature

coefficient of solubility is low or where the solute is unstable at elevated

temperatures. Its use in the pharmaceutical industry is, therefore common. Particular

problems associated with anti-solvent crystallization include wide batch-to-batch

variability, fine and irregularly shaped product crystals and solvate/hydrate formation.

The effects of various process parameters, including type of anti-solvent (Takiyama et

al., 1998; Oosterhof et al., 1999), feed concentration (Holmbäck & Rasmuson, 1999;

Barata & Serrano 1998(b)) solution concentration (Kaneko et al., 2002; Kitamura &

Sugimoto, 2003), anti-solvent addition rate (Beckmann, 1999; Holmbäck &

Rasmuson, 1999) and agitation intensity (Takiyama et al. 1998; Yu et al., 2005), on

the size, number, shape, degree of agglomeration and polymorphic form of product

crystals have been reported. However, less attention has been paid to the metastable

zone width and its relationship with the process conditions (Guo et al., 2005; Pina et

al., 2001).

* This proposal has been confirmed by the application of computational fluid dynamics (CFD) to characterise mixing conditions

at both locations. A full treatment of the application of CFD to this system can be found in Appendix A.

Page 107: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

97

The metastable zone width is an extremely important parameter in the design and

optimisation of crystallization processes. A solute will remain in solution until a

sufficiently high level of supersaturation is generated to induce spontaneous

nucleation. Typically, it is desirable to operate away from this metastable limit so as

to ensure reliable process performance. Metastable zone information can be used to

optimise crystallization processes (Ulrich & Strege, 2002) and calculate nucleation

kinetics (Nyvlt, 1968). The MSZW may be affected by various process parameters,

such as supersaturation generation rate (Barrett & Glennon, 2002), agitation speed

(O‟Sullivan, 2005) and the presence of impurities (Myerson & Jang, 1995; Sayan &

Ulrich, 2001).

Recent research aimed at modeling the impact of mixing on crystallization has

focused on reactive crystallization (Tavare, 1995; Phillips et al., 1999; David, 2001;

Torbacke & Rasmusson, 2001 & d2004). Little work has been conducted studying the

impact of mixing on the MSZW for anti-solvent systems. Mixing is critical for all

crystallization systems but in the case of anti-solvent addition adequate mixing is

needed to incorporate the anti-solvent into the bulk solution and maintain a constant

level of supersaturation throughout the crystallizer, over a potentially wide volume

range. This must be achieved while considering many other mixing sensitive

parameters such as solids suspension, attrition, impurity profile, agglomerate

formation/break-up, of entrainment of gas/vapor from the headspace and rate of heat

transfer (Paul, 2005). In this work, the effect of addition rate, agitation speed and feed

point location on MSZW are investigated at a 500 mL scale. FBRM and ATR-FTIR

are implemented to detect nucleation and the results are compared. Appendix A

Page 108: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

98

highlights the use of Computational Fluid Dynamics (CFD) software is used to model

the mixing behaviour of the systems under investigation and to provide an explanation

of the observed results.

With accurate MSZW information it is possible to estimate nucleation kinetics. This

is achieved my modifying classical nucleation theory (Nyvlt, 1968) for an anti-solvent

system. Much of the literature focuses on static induction time experiments to

estimate nucleation kinetics for an anti-solvent system (Barata & Serrano, 1996(b)).

However useful nucleation kinetics can be gathered using dynamic experiments and

this is the method will be used here.

4.3. EXPERIMENTAL METHODS

The solute in this work, benzoic acid, is soluble in ethanol (58.36 g/100g @ 25 C)

and essentially insoluble in water (0.25g/100g @ 25 C) (Perry). All metastable zone

width measurements were carried out at 25 C.

An initially undersaturated solution, containing 75g water, 75g ethanol and 21g

benzoic acid, was held at 25 C in a 500 mL glass jacketed vessel with a Julabo chiller

fitted for temperature control. A motor driven, pitched blade impeller provided

agitation. FBRM (S400 A; Mettler Toledo), ATR-FTIR (ReactIR 4000; Mettler

Toledo) and temperature probes were placed in the vessel to monitor the

crystallization and as a consequence provided baffling (Figure 4.1). Water at 25 C

was fed to the vessel at a variety of rates (0.05 gs-1

, 0.14 gs-1

, 0.24 gs-1

, 0.34 gs-1

and

Page 109: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

99

0.48 gs-1

). Two addition locations (Figure 4.1) and two agitation intensities (325 rpm

and 475 rpm) were investigated. All experiments were performed in triplicate and the

reported values are the mean of the three experiments. The error bars reported for the

standard deviation of the mean. FBRM and ATR-FTIR were used to detect nucleation

and the MSZW is reported in terms of concentration (g anti-solvent / g solvent) as

outlined in Chapter 2. Solubility data for the system can also be found in Chapter 2.

Figure 4.1: Crystallizer configuration

Page 110: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

100

4.4 RESULTS AND DISCUSSION

4.4.1 Comparison of FBRM and ATR-FTIR for the detection of nucleation

FBRM and ATR-FTIR detect the formation of crystals in different ways. FBRM

detects the point at which crystals are present in the solution, whereas ATR-FTIR

detects the point at which the solution concentration decreases, indicating solute is

being forced out of solution and crystals are being formed. To measure the actual

solution concentration a calibration model is needed, but for the purpose of

identifying the point of nucleation such a model is not necessary, since it is sufficient

to examine peaks in the spectrum associated with benzoic acid.

Figure 4.2: ATR-FTIR waterfall plot for standard MSZW experiment

Figure 4.2 shows a waterfall plot for a typical crystallization. The peak at 1275 cm-1

represents the C-O bond present in the benzoic acid molecule. The peak at 1480 cm-1

Benzoic Acid Peak

Page 111: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

101

is representative of a C-C bond indicating an aromatic group. After a 10-minute hold

period to equilibrate the temperature, water is added, in this case, at 0.045 g/s (and

agitation speed of 475 rpm). Examining the peak at 1275 cm-1

the concentration of

benzoic acid in solution decreases at a steady rate due to dilution. After about 5

minutes there is a sharp decrease in this peak, indicating that the concentration of

benzoic acid in solution is decreasing at a faster rate, due to the formation of benzoic

acid crystals. In this way the onset of nucleation can be identified using ATR-FTIR

with no need for a calibration model.

The height of a specific peak can be trended against time so the onset of nucleation

can be identified easily. Figure 4.3 (a) shows the comparison between FBRM counts

(10-100 μm), and the height of the benzoic acid peak at 1275 cm-1

. It was noted that

the counts per second data in the region 0-10 μm exhibited more noise than the counts

data in the 10-100 μm region. Therefore, the count rate in the latter region was used

for nucleation detection. The chosen range also ensured that agglomerates of newly

formed crystals were also detected. In general, however, it is advisable to choose the

most sensitive detection range by inspection. The FBRM indicates clearly the point at

which the first crystals are observed while the decrease in intensity of the benzoic acid

peak indicates the point at which solute comes out of solution. Figure 4.3(b) focuses

on the region where nucleation occurs and shows that FBRM detects the onset of

nucleation before the ATR-FTIR. For all addition rates the FBRM detected the point

of nucleation before the ATR-FTIR. (Figure 3.4) with the difference most pronounced

at higher addition rates. FBRM has an advantage over ATR-FTIR in detecting the

point of nucleation due to its much shorter sampling time of 2 seconds rather than 30

seconds. It is possible to scan faster with the ATR-FTIR probe, but this increases the

Page 112: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

102

likelihood of errors in the measurement. For this reason FBRM was chosen to indicate

the point of nucleation.

0

500

1000

1500

2000

2500

3000

3500

0 5 10 15 20 25

TIme (min)

#/s

10-1

00 m

icro

ns

0

0.02

0.04

0.06

0.08

0.1

0.12

Rela

tive A

bs

FBRM

ATR-FTIR

Figure 4.3 (a): Time vs. FBRM counts and peak height

5

10

15

20

12 13 14 15 16 17 18

TIme (min)

#/s

10-1

00 m

icro

ns

0.02

0.04

0.06

0.08

0.1R

ela

tiv

e A

bs

FBRM

ATR-FTIR

Figure 4.3 (b): Time vs. FBRM counts and peak height: Point of nucleation

FBRM Nucleation

ATR-FTIR Nucleation

Page 113: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

103

0

0.1

0.2

0.3

0.4

0 0.1 0.2 0.3

Addition Rate (g/s)

MS

ZW

(g

/g)

ATR-FTIR

FBRM

Figure 4.4: Comparison of MSZW measured by FBRM and ATR-FTIR

Page 114: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

104

4.4.2 Impact of Process Parameters on the MSZW

Figure 4.5 shows the impact of addition rate and agitation intensity on the MSZW

when anti-solvent is added close to the impeller (addition point 1 in Figure 3.1). At a

given agitation intensity the MSZW becomes wider as addition rate increases. The

increase in MSZW with increasing supersaturation generation rate is typical of most

crystallization systems.

For each addition rate studied an increase in the agitation intensity resulted in a

natrower MSZW. Agitation can have two effects on the metastable zone width.

Increased agitation can narrow the MSZW by increasing the probability of solute

molecules contacting to form the critical sized nuclei necessary for nucleation.

However this increase in agitation can also break up these clusters widening the

MSZW. For an anti-solvent system there is the added factor of mixing between the

solvent and anti-solvent.

Page 115: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

105

0.0

0.1

0.2

0.3

0.4

0 0.1 0.2 0.3 0.4 0.5

Addition Rate (g water/s )

MS

ZW

(g

wate

r/g

eth

an

ol)

475 rpm - impeller addition location

325 rpm impeller addition location

Figure 4.5: Addition Rate vs. MSZW – addition location close to the impeller

Similar experiments were performed with an addition point close to the wall of the

vessel (addition point 2 in Figure 4.1). In this case the addition rate had substantially

less effect on the MSZW (Figure 4.7). In general the MSZW is significantly narrower

when the anti-solvent is added close to the wall. This can be explained by considering

the unfavourable mixing conditions at the wall. Addition close to the impeller ensures

rapid incorporation of the anti-solvent into the bulk solution, whereas addition close

to the wall leads to areas of locally high supersaturation close to the addition location.

There was also a far greater variability in the MSZW, especially at 325 rpm. At high

addition rates it becomes harder to incorporate the anti-solvent into the bulk solution.

This leads to an area of local supersaturation close to the feed point. This can result in

premature nucleation often at bulk concentrations below the saturation level.

Increasing the agitation facilitates dissipation of the local supersaturation. This

Page 116: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

106

explains why increasing the agitation intensity results in a wider MSZW as well as

reduced variability.

-0.1

0.0

0.1

0.2

0.3

0.4

0 0.1 0.2 0.3 0.4 0.5

Addition Rate (g water/s )

MS

ZW

(g

wate

r/g

eth

an

ol)

475 rpm - wall addition location

325 rpm - wall addition location

Figure 4.7: Addition Rate vs. MSZW – addition location close to the vessel wall

4.4.3 Nucleation Kinetics

With sufficient metastable zone width data it is possible to calculate nucleation

kinetics for the anti-solvent system. It is important to identify an operating regime

where kinetics may be reliably calculated. In the situation where anti-solvent was

added at the wall location, and at 325 rpm, it is probable that true nucleation kinetics

cannot be estimated. This is because the observed nucleation occurs due to excessive

local supersaturation near the addition point rather than true supersaturation in the

bulk of the liquid. When the addition location is close to the impeller, the

nucleationdoes not appear to be so dominated by hydrodynamic conditions. In this

case useful kinetic information may be estimated.

Page 117: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

107

Classical nucleation theory (Nyvlt, 1968) can be applied to the MSZW data as used

with modifications for anti-solvent addition rather than cooling.

no Ck

dt

dNJ 4.1

During anti-solvent addition, the rate of supersaturation generation can be expressed

as a function of the specific anti-solvent addition rate,

dA

dcR

dt

cd s. 4.2

At the point of nucleation, the supersaturation is related to the metastable zone width

by

dA

dcAC s.max

max 4.3

where the MSZW is given by

nucs AAAmax 4.4

Using the solubility data reported in Chapter 3 dcs/dA is equal to 0.2231 g solute g-1

solvent g-1

anti-solvent for the benzoic acid-ethanol-water system. At the point of

Page 118: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

108

nucleation, it may be assumed that the rate of supersaturation generation equals the

rate of formation of new crystals which is given by

nn

nnuco CkCrk

dt

dM 3 4.5

Combining equations 4.2, 4.3 and 4.4 yields,

maxlnlnln1ln AnkdA

dcnR n

s 3.7

Using equation 4.7 a plot of ln (R) verses ln (ΔAmax), will yield a line of slope n.

(Figure 4.8). Fitting a simple trend line to the data allows n and kn to be calculated.

Figure 4.8: ln (MSZW) vs. ln (addition rate) with trend-lines

Page 119: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

109

Figure 4.8 clearly indicates that agitation has an impact on the nucleation kinetics. In

order to study the effect of agitation on the kinetics it is proposed that the nucleation

rate expression can be modified to include an expression for the agitation intensity

(β),

dA

dcrA

dA

dckCNkJ s

n

sn

n

n max 4.8

Taking logarithms again,

)ln(lnlnln1ln max NAnkdA

dcnR n

s 4.9

In this case a plot of ln(R) vs. ln( maxA ) will once again yield a straight line. In order

to solve for n, kn and β a non-linear regression was applied to the data using the

known values of dA

dc*

, N and maxA . The regression analysis yielded the following

expression for the nucleation rate (Equation 4.10).

5.21.178.2 CNJ 4.10

From equation 4.9, the metastable zone width can be obtained thus

Page 120: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

110

111

'max

ns

n

ndA

dc

Nk

RA 4.11

Good agreement is obtained between the predictions of equation 4.11 and the

experimental data (see Figure 4.9), supporting the contention that impeller speed

plays a critical role in determining the level of nucleation in the vessel, and thus

providing a correlation to facilitate quantification of the role of impeller speed in the

determination of nucleation rates. This is consistent with the observations made based

on the measured trends in metastable zone width.

Figure 4.9: Metastable zone widths predicted form Equation 4.11

Page 121: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

111

4.5 DISCUSSION

The effect of addition rate and location, along with agitation intensity, on the

metastable zone width has been investigated for an anti-solvent crystallization. An

addition location close to the impeller results in a repeatable crystallization and a

positive correlation between the addition rate and the MSZW. An addition location

close to the wall of the vessel results in a significantly narrower MSZW, and more

random nucleation behaviour.

The effect of agitation intensity depends on the addition location. When anti-solvent

is added close to the impeller, an increase in agitation intensity results in a narrower

MSZW, possibly due to the increased probability of contact between solute

molecules. When anti-solvent is added close to the wall, an increase in agitation

results in a wider MSZW and a significant improvement in the batch-to-batch

repeatability. These results can be explained in terms of mixing conditions at each of

the addition locations. Close to the impeller, mixing conditions allow for the rapid

incorporation of the anti-solvent and a homogenous mixture of solution and anti-

solvent. However, close to the wall, mixing conditions are less suitable and areas of

supersaturation build up leading to narrower MSZWs and a reduction in the batch-to-

batch repeatability. In this situation, when the agitation is increased, the local areas of

supersaturation can be dissipated, to some degree, and the MSZW is wider and the

batch-to-batch repeatability improves.

By modifying classical nucleation theory for an anti-solvent system, it was possible to

estimate the nucleation order and the nucleation constant at both agitation intensities,

Page 122: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

112

and for the addition location close to the impeller. Only this location was studied as

addition close to the wall resulted in unrepeatable experiments and nucleation was

judged to be a result of local areas of supersaturation close to the addition location

rather than as a result of supersaturation in the bulk solution. Agitation clearly

impacted on the nucleation kinetics and to account for this the nucleation rate

expression was modified, to account for agitation, and solved using a non-linear

regression. The modified equation fitted the experimental data adequately.

Page 123: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

113

CHAPTER 5: THE USE OF FBRM AND ATR-FTIR TO

MONITOR ANTI-SOLVENT CRYSTALLIZATION AND

ESTIMATE GROWTH RATE KINETICS

5.1 ABSTRACT

The unseeded, anti-solvent crystallization of benzoic acid, from ethanol-water

mixtures using water as an anti-solvent, is presented. Focussed Beam Reflectance

Measurement (FBRM) and Attenuated Total Reflectance Fourier Transform Infrared

spectroscopy (ATR-FTIR) were used to monitor the solid and liquid phase

respectively. FBRM was used to track crystal growth and ATR-FTIR was used to

measure the water concentration and the benzoic acid concentration. In combination

with previously gathered solubility data (Chapter 3) the measured concentration data

was used to monitor the supersaturation. Using previously gathered MSZW

information

(Chapter 4), optimal mixing conditions were chosen to ensure a

repeatable crystallization with a low nucleation rate. Growth rate kinetics were

estimated by combining suitable statistics, from the FBRM chord length distribution,

with the supersaturation data.

5.2 INTRODUCTION

In recent years the use of in situ tools for the study of crystallization systems has

gained much attention. This has been driven by the need to rapidly gather

representative, real time data, without sampling. Sampling can be time and labour

intensive, unrepresentative, and may require the sample crystals to be prepared in

Page 124: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

114

some way, typically through sonication or dilution. Also, the FDA‟s Process

Analytical Technology (PAT) initiative (FDA website) encourages the use of in situ

tools to monitor processes such as crystallization.

FBRM and ATR-FTIR have emerged as excellent tools for the study of crystallization

with much work in the literature devoted to their application. FBRM is a probe-based

instrument that measures a chord length distribution of the crystal slurry, which is a

function of the dimension, population and shape of the crystals in the solution. It is

extremely sensitive to fine crystals and can measures tens of thousands of crystals per

second allowing a very robust real-time measurement to be gathered. FBRM has,

amongst other things, been used to measure solubility and the MSZW (Barrett &

Glennon 2002), study crystal size (Kougoulos et al., 2005), and assess the effect of

impurities on batch time (Scott and Black, 2005).

ATR-FTIR is another probe-based tool that allows the liquid phase concentration to

be measured once a calibration model has been created. Importantly, for the anti-

solvent system the ATR-FTIR probe offers the opportunity to measure both the solute

and anti-solvent concentration. In combination with accurate solubility data it is then

possible to monitor supersaturation. ATR-FTIR has been used to optimise

crystallization processes (Lewiner et al., 2001) and monitor and control

supersaturation (Liotta & Sabesan 2004).

The mode of operation of both of these instruments is outlined in detail elsewhere

(Sparks & Dobbs 1994; Dunwilla & Berglund 1997) and a detailed review of each of

these instruments can be found in Chapter 2.

Page 125: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

115

By combining supersaturation data, gathered using ATR-FTIR, with growth rate data,

gathered using FBRM, it is possible to estimate growth rate kinetics according to

Equation 5.1 (Mullin 1993).

G = kgΔCg 5.1

A plot of supersaturation against growth rate allows the growth coefficient and the

growth order to be estimated. This information is useful for the estimation of batch

time and calculation of suitable anti-solvent addition rates. By assessing the growth

rate and supersaturation under different mixing conditions in the laboratory the impact

of scale on the growth kinetics can be identified.

5.3 MATERIALS AND METHODS

A solution consisting of 75 g of water 75 g of ethanol and 21 g of benzoic acid was

made up and placed in the 500 mL glass reactor. FBRM, ATR-FTIR and temperature

probes provided baffling and a pitched blade impeller provided agitation. The

temperature was allowed to equilibrate at 25 C and the solution was held for 20

minutes to ensure all the benzoic acid had dissolved. Previous work identified that a

slow addition rate close to the impeller, with high agitation produced the most robust

crystallization in terms of consistent metastable zone width and low nucleation rate

(Chapter 4). With this in mind, water was added close to the impeller at a rate of

0.065 g/s for 45 mins, and the agitation was set at 475 rpm. The contents of the vessel

were held for a further 15mins after the end of the addition period to ensure the

supersaturation was consumed and the crystallization complete. Three similar

Page 126: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

116

experiments were performed (Run 1, Run 2, Run 3). In the case of Run 3 the addition

period was 40 mins and the hold period was 5 mins.

5.4 RESULTS AND DISCUSSION

5.4.1 FBRM Results

A comparison of the three runs was made in order to assess the repeatability of the

crystallization under the chosen process parameters. One useful aspect of the FBRM

measurement technique is its ability not only to measure the degree of change of the

size, number and shape of the crystals, but to also measure their rate of change. Figure

5.1 shows three FBRM statistics plotted against time, for each run.

Figure 5.1 focuses on #/s (1-10 microns) and is useful for tracking nucleation, as it

highlights the shortest chord lengths. It is clear that the nucleation rate for each run is

similar and that there is no secondary nucleation event apparent, a phenomenon

typically characterised by a sudden increase in fine chords after the primary

nucleation event.

Figure 5.2 focuses on the larger crystals and tracks #/s (100-1000 microns) and

highlights the longer chord lengths, making it useful for tracking growth. The initial

increase in this statistic occurs after the increase in #/s (1-10 microns) indicating that

this statistic is less sensitive to the primary nucleation event, and that it takes some

time for the crystals to grow into larger size ranges.

Page 127: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

117

The initial rate of increase in #/s (100-1000 microns) is rapid, when supersaturation is

high, and gradually plateaus as supersaturation is consumed. A similar result is

observed for the mean square weighted chord length (MSQW) (Figure 5.3). A square

weighting emphasises the long chords and de-emphasises the fine chords. This makes

the MSQW a very useful statistic for tracking crystal growth. Prior to the nucleation

event there is significant noise in the MSQW measurement. Even though count data

are low the MSQW is influenced by small changes in the chord length distribution.

0

200

400

600

800

1000

1200

1400

0 600 1200 1800 2400 3000 3600

Time (s)

#/s

(1-1

0 m

icro

ns)

run 1

run 2

run 3

Figure 5.1: FBRM data for three similar experiments: (a) #/s 1-10 microns

a

0

100

200

300

400

500

600

0 600 1200 1800 2400 3000 3600

#/s (100-1000 m

icro

ns)

Run 1

Run 2

Run 3

a

Page 128: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

118

0

100

200

300

400

500

600

0 600 1200 1800 2400 3000 3600

Time (s)

#/s

(100-1

000 m

icro

ns)

run 1

run 2

run 3

Figure 5.2: FBRM data for three similar experiments: #/s 100-1000 microns

0

50

100

150

200

250

300

350

400

0 600 1200 1800 2400 3000 3600

Time (s)

Sq

uare

Weig

hte

d M

ean

Ch

ord

Len

gth

(m

icro

ns)

run 1

run 2

run 3

Figure 5.3: FBRM data for three similar experiments: square weighted mean chord length (microns)

b

0

100

200

300

400

500

600

0 600 1200 1800 2400 3000 3600

#/s (100-1000 m

icro

ns)

Run 1

Run 2

Run 3

c

0

100

200

300

400

500

600

0 600 1200 1800 2400 3000 3600

#/s (100-1000 m

icro

ns)

Run 1

Run 2

Run 3

a

Page 129: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

119

The chord length distribution itself can also be used to assess repeatability and is very

useful for identifying crystallization mechanisms. Figure 5.4 shows the (a)

unweighted and (b) square weighted chord length distribution for each run at three

different points in time. Immediately it is clear that the chord length distribution is

similar for each run at each time point.

Figure 5.4 also shows that the crystallization is growth dominated. After 30 minutes

there is little change in the unweighted distribution, but a large increase in the square

weighted distribution. This indicates that after the primary nucleation event, growth

dominates the crystallization. Slight differences in the trends and chord length

distributions between the first two batches and the third batch may be due to the small

difference in operating conditions.

It is sometimes difficult, based on FBRM data alone, to discern between crystal

growth and agglomeration. Typically FBRM measures agglomeration as a reduction

in fine chords and a corresponding increase in large chords as small crystals come

together to form large agglomerates. However, if growth and agglomeration are

occurring simultaneously this may not be the case and fine counts may actually

increase as the likelihood of measuring a short chord increases as the crystals get

larger. PVM images taken during a similar crystallization* show that the benzoic acid

forms large well formed elongated platelets. It also appears that the crystals come

together to form loosely bound agglomerates or flocs (Figure 5.5) PVM is

advantageous for this analysis as taking samples and analyzing the crystals under a

* It was not possible to insert the FBRM, ATR-FTIR and PVM probes into the crystallizer at the same

time. PVM images were taken from an experiment performed under the same conditions but without

the FBRM and ATR-FTIR probes present. It is assumed that the different mixing conditions

encountered do not affect the crystallization appreciably.

Page 130: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

120

microscope can lead to misleading information as a result of agglomeration during

filtration or drying, or simply because crystals lie on top of each other and resemble

agglomerates under the microscope.

Page 131: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

121

0

20

40

60

80

100

120

140

1 10 100 1000 10000

Chord Length (microns)

#/s

run1 t = 20 mins

run2 t = 20mins

run3 t = 20 mins

run1 t = 30 mins

run2 t = 30 mins

run3 t = 30 mins

run1 t = 50 mins

run2 t = 50 mins

run 3 t = 50 mins

0

2

4

6

8

10

12

14

1 10 100 1000

Chord Length (microns)

#/s

run1 t = 20 mins

run2 t = 20mins

run3 t = 20 mins

run1 t = 30 mins

run2 t = 30 mins

run3 t = 30 mins

run1 t = 50 mins

run2 t = 50 mins

run 3 t = 50 mins

Figure 5.4: Chord length distributions at three time stamps for each run (a) unweighted; (b) square

weighted

Page 132: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

122

Figure 5.5: PVM images taken after 30 minutes of crystallization

Page 133: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

123

5.4.2 ATR-FTIR Results

Figure 5.6(a) shows water concentration and benzoic acid concentration, for the three

runs, gathered using ATR-FTIR. The linear increase in the mass ratio of water is

evident, as is the end of the addition period at 2700s for runs 1 and 2 and 2400s for

run 3. Prior to nucleation the benzoic acid concentration remains essentially constant,

with a slight decrease apparent. This subtle change is most likely due to a weakness in

the calibration model in the region close to the starting point of the crystallization. It

may also be a result of variations in temperature due to a heat of mixing. The point at

which the concentration decreases dramatically represents a point on the MSZW and

coincides with the increase in #/s (1-10 microns) (Figure 5.1). In general FBRM is

more sensitive to the nucleation event than ATR-FTIR (Chapter 4).

Supersaturation is calculated by combining accurate solubility data (Chapter 3) with

the measured concentrations (Figure 5.46(a)) and is shown in Figure 5.4 (b). The

point of nucleation is observed when the supersaturation drops and it is clear that all

of the supersaturation is consumed soon after the end of the addition period (2700s).

This indicates that a short hold period is sufficient to ensure the crystallization is

complete and the yield is maximised.

Page 134: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

124

0.5

1.0

1.5

2.0

2.5

0 600 1200 1800 2400 3000 3600

Time(s)

Wate

r C

on

cen

trati

on

(g

/g)

0.1

0.2

0.3

0.4

Ben

zo

ic A

cid

Co

ncen

trati

on

(g

/g)

Water Run 1 Water Run 2 Water Run 3

Benzoic Run 1 Benzoic Run 2 Benzoic Run 3

-0.12

-0.09

-0.06

-0.03

0.00

0.03

0.06

0 600 1200 1800 2400 3000 3600

Time (s)

Su

pers

atu

rati

on

(g

/g)

run 1

run 2

run 3

Figure 5.6: ATR-FTIR data for three similar runs: (a) Water concentration and benzoic acid

concentration; (b) Supersaturation

a

b

Page 135: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

125

5.4.3 Growth Rate Kinetics Estimation

Section 5.4.1 shows that nucleation dominates only in the early stages of the

crystallization and that growth with some agglomeration is dominant for the

remainder. Growth rate kinetics are estimated once the primary nucleation event is

complete. This exact point is open to interpretation, however, close examination of

Figure 5.1 shows that the initial rapid increase in #/s (1-10 microns) is probably over

after about 1300 s. With the region of interest identified, it is necessary to choose a

statistic that will suitably track the growth. As outlined in section 4.4.1 #/s (100-1000

microns) and the MSQW are track the growth phase of the crystallization in a similar

fashion. For the purpose of this study the MSQW will be used to estimate the growth

rate kinetics as it is measured in units of dimension rather than population-per-second

facilitating interpretation.

The measurement of accurate growth rate kinetics in this case is difficult due to the

initial uncontrolled nucleation event and the subsequent loose agglomeration outlined

in the previous section. While pure crystal growth is not apparent in this system the

method described will serve to highlight the possibility of rapidly measuring growth

rates in-process and in real-time. For a seeded crystallization with no agglomeration

the accuracy of this method for estimating the growth kinetics maybe improved.

The growth rate is measured by applying the first derivative to the MSQW in the

FBRM software according to Eq. 5.2 and is plotted in Figure 5.7.

1

1'

kk

kkk

tt

vvv Eq. 5.2

Page 136: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

126

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

0 500 1000 1500 2000 2500 3000 3500 4000

Time (s)

MS

QW

(m

icro

ns)

run 1

run 2

run 3

Figure 5.7: Growth rate data gathered using FBRM data after 1300s: square weighted mean chord

length

The FBRM and ATR-FTIR data were combined, according to Equation 5.1, to

measure growth rate kinetics. By plotting supersaturation (Figure 5.4(b)) against

growth rate (Figure 5.7) the kinetic parameters g and kg can be estimated. Figure 5.8

shows that, as expected, high growth rates occur at high supersaturation. A non-linear

least squares fit was applied to the data and is also shown in Figure 5.8. In this case

the growth order is close to unity.

Page 137: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

127

0

0.05

0.1

0.15

0.2

0.25

0.3

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035

Supersaturation (g./g)

Gro

wth

Rate

(m

icro

ns

/s)

Model

Run 1

Run 2

Run 3

Figure 5.7: Growth rate kinetics using three experiments

To eliminate scatter from the data a second approach was taken where the growth rate

and supersaturation data was averaged over the three experiments (Figure 5.7) and

then combined. Figure 5.9 shows a plot of the averaged supersaturation against the

averaged growth rate. Clearly this approach results in less scatter to the data. Since the

experiments were repeatable this approach was deemed suitable for calculating the

average growth rate over a number of crystallizations. The calculated value of 1.1 for

the growth order is in the acceptable range for organic crystallization processes.

G = 3.61ΔC0.91

Page 138: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

128

-0.12

-0.10

-0.08

-0.06

-0.04

-0.02

0.00

0.02

0.04

0 600 1200 1800 2400 3000 3600

Time (s)

Su

pe

rsa

tura

tio

n (

g/g

)

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0 600 1200 1800 2400 3000 3600

Time (s)

Averg

aed

Gro

wth

Rate

- S

QW

M (

mic

ron

s/s

)

Figure 5.8: (a) Supersaturation average; (b) FBRM growth rate average

a

b

Page 139: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

129

0

0.05

0.1

0.15

0.2

0.25

0 0.005 0.01 0.015 0.02 0.025 0.03

Supersaturation (g/g)

Gro

wth

Rate

(M

SQ

W -

mic

ron

s/s

)

Model

Experimental

Figure 5.9: Growth rate kinetics estimated using averaged values

G = 63.24ΔC1.27

G = 7.8ΔC1.1

Page 140: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

130

5.5 CONCLUSIONS

FBRM and ATR-FTIR have been successfully used to monitor the anti-solvent

crystallization of benzoic acid from ethanol water mixtures using water as the anti-

solvent. FBRM showed that the nucleation rate for each experiment was similar and

that under the chosen process conditions the crystallization was robust and repeatable.

ATR-FTIR successfully tracked the water concentration and the benzoic acid

concentration, and in conjunction with previously gathered solubility data the

supersaturation was monitored.

Growth rate data were calculated by combining the supersaturation data and growth

rate data, calculated using the square weighted mean from the FBRM data. A growth

rate order of 1.1 was calculated, a suitable number for an organic crystallization

system.

The accuracy of the measured kinetics is questionable due to the uncontrolled

nucleation events prior to growth and the agglomeration during the growth phase;

however this technique highlights the ease with which growth rate kinetics may be

estimated in-process. This also opens the opportunity to measure the growth rate

kinetics in real time by combining the FBRM and ATR-FTIR data as the

crystallization is being performed. By choosing a system where agglomeration is

minimal and nucleation can be minimised, and by using seed, the accuracy of the

growth rate kinetics may be improved.

Page 141: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

131

CHAPTER 6: SCALE-UP OF ANTI-SOLVENT

CRYSTALLIZATION

6.1 ABSTRACT

The pilot-scale crystallization of benzoic acid, from ethanol-water mixtures, using

water as the anti-solvent is presented. With extensive small-scale characterisation

performed it was possible to increase the scale of the crystallization from the

laboratory scale (500 mL) to the pilot plant scale (70 L) with a view to assessing

crystallization behaviour on scale-up. Prior to the experimental work the pilot scale

vessel was modelled using computational fluid dynamics (CFD) in order to assess the

mixing regime.

A scale-up strategy based on the laboratory scale experiments was implemented and

three batches were run at the pilot-scale under various operating conditions to assess

which parameters most closely met the requirements of producing crystals of a similar

size. CFD was used to choose a suitable anti-solvent addition point. Each run was

monitored in-line using Focussed Beam Reflectance Measurement (FBRM) and

Particle Vision and Measurement (PVM).

Similar chord length distributions were achieved at the laboratory-scale and the pilot-

scale indicating that the operating conditions chosen were suitable in terms of

conserving the particle size on scale-up.

Page 142: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

132

6.2 INTRODUCTION

While extensive characterisation work is possible at the laboratory scale there is

almost always difficulty associated with reproducing similar results at a large-scale.

This is most often due to differences in mixing conditions between scales. An

intermediate, pilot-plant scale is often used to try and predict how a crystallization

may behave upon scale-up without the expense, time and labour associated with

scaling directly to the plant. With adequate characterisation at the small scale a small

number of experiments can be performed at the pilot scale giving valuable

information regarding potential plant scale issues. In the pharmaceutical industry this

is often performed in tandem with the production of material for use in clinical trials.

In order to assess the impact of scale on the crystallization, key parameters must be

measured. In situ tools provide the opportunity to do this without the need for

sampling and can continuously monitor in real time the progression of the

crystallization. Critically they can be implemented at a variety of scales allowing a

direct comparison to be made between experiments conducted in the laboratory, pilot

plant and in production.

In this study, data gathered at the small scale along with CFD models were used to

choose suitable process conditions at the pilot scale to produce a robust crystallization

with a particle size similar to that produced at the laboratory scale. The laboratory

studies indicated that agitation intensity, addition location and addition rate were the

critical parameters that needed to be controlled. At this scale a robust and repeatable

crystallization was achieved when the agitation intensity was high, the addition rate

Page 143: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

133

was low and the addition location was optimized to ensure rapid incorporation of the

anti-solvent into the bulk solution. With this in mind a scale-up protocol was devised.

The addition rate was chosen to ensure a similar batch time at the lab and pilot scales.

The agitation intensity was chosen as the maximum that would not result in

entrainment of bubbles into the system. The addition location was chosen based on

CFD results for the large scale vessel .these indicated that the optimal addition

location may change during the batch so experiments were designed to take advantage

of that.

6.3 MATERIALS AND METHODS

For the laboratory-scale experiment 75g of ethanol, 75g of water and 21g of benzoic

acid were added to a 500 mL glass vessel, fitted with a pitched blade impeller set at

475 rpm, and held at 25°C. Water was added at 0.065 g/s above the liquid surface

close the impeller for 45 mins. This was followed by a 15 minute hold period. These

experimental conditions resulted in a repeatable crystallization based on the FBRM

trends (Chapter 5).

The same solution concentration was employed at the pilot-scale where 12.5 kg of

ethanol, 12.5 kg of water and 3.5 kg of benzoic acid were added to a 70 L glass stirred

tank reactor fitted with a Rushton turbine and four baffles. Figure 6.1 shows the vessel

dimensions and the position of the probes and internals. The contents were stirred at

130 rpm for 30 mins to ensure complete dissolution of the benzoic acid. This

agitation intensity was the highest possible without entraining air into the system.

Water was then added at 10 gs-1

for 45 mins in order to achieve the same batch time

Page 144: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

134

as the experiments conducted at the laboratory-scale. At the end of the addition period

the solution was held for a further 15 minutes to ensure complete crystallization.

Three pilot-scale batches were run under different mixing conditions and addition

locations. For batch A the agitation was set at 130 rpm, the highest agitation intensity

that did not result in a vortex, and the anti-solvent was added at point 2 (Figure 6.1).

This addition location was chosen based on CFD simulations showing the highest

downward velocity at this point (Appendix A). For batch B the same addition location

was chosen (point 2) but in this case the agitation was increased from 157 rpm to 250

rpm so as to maintain the same power per unit volume for the duration of the batch

(Figure 6.2). The starting agitation of 157rpm was calculated based on a constant

power per unit scale up from the laboratory vessel. At the start of the batch vortexing

and air entrainment were noticeable but dissipated as the volume of the batch

increased. For batch C the same agitation profile as that used for batch B was

implemented but the addition point was changed five minutes into the experiment

from point 1 to point 2. This decision was based on CFD simulations indicating the

optimal addition location changed as the volume in the vessel increased (Appendix

A). Table 6.1 summarizes the operating conditions used for the three pilot-scale

batches.

Page 145: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

135

Figure 6.1: Pilot-scale vessel and internals; (1) Addition Location 1; (2) Addition

Location 2; (3) PVM probe; (4) FBRM probe; (5) Overflow

0

100

200

300

400

500

600

0 10 20 30 40 50 60

Time (mins)

P/V

(W

/m3)

0

50

100

150

200

250

300

350

400

Ag

ita

tio

n In

ten

sit

y (

rpm

)

P/V

Rpm

Figure 6.2: Time vs. agitation intensity and power-per-unit-volume for batches B and C

Page 146: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

136

Table 6.1: Summary of Pilot-Scale Batches

6.3 RESULTS AND DISCUSSION

6.3.1 FBRM Results

Figure 6.3 shows the chord length distributions measured at the end of batches A, B

and C. Clearly batches B and C, with constant power-per-unit-volume, are very

similar at the fine (unweighted) and coarse (square weighted) ends of the distribution.

The change in addition location implemented in Batch C appears to have had no

effect.

For Batch A, a slight increase in fine and coarse particles compared to the other two

batches is observed. In this case it is difficult to ascertain whether this is a function of

the crystallization mechanism or a function of the presentation of the crystals to the

FBRM window. At 130 rpm the mixing was visibly inadequate to fully suspend the

crystal slurry and there were a number of dead zones at the top, bottom and sides of

the reactor. Evidence of this segregation is shown in Figure 6.4 where samples taken

from the top and bottom of the pilot-scale vessel are compared off-line using FBRM.

The chord length distributions show significantly more fine material for the sample

Addition Rate Agitation Intensity Addition Location

Batch A 10gs-1

130 rpm

2

Batch B

10gs-1

(157 rpm – 200rpm) 2

Batch C 10gs-1

(157 rpm – 200rpm) Points 2 & 1

Page 147: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

137

taken from the top and significantly more coarse material for the sample taken at the

bottom. Smaller crystals floated to the top of the vessel and were not re-incorporated

into the bulk of the slurry while larger crystals sank to the bottom and were not re-

suspended. Thus, for batch A, the chord length distribution measured by the FBRM in

a turbulent zone close to the impeller may not be representative of the crystallization

as a whole.

0

20

40

60

80

100

120

1 10 100 1000

Chord Length (microns)

#/s

Batch A

Batch B

Batch C

Page 148: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

138

0

2

4

6

8

10

12

14

16

18

1 10 100 1000

Chord Length (microns)

#/s

Batch A

Batch B

Batch C

Figure 6.3: Chord Length Distributions for Batches A, B and C; (a) unweighted, (b) square weighted

0

500

1000

1500

2000

2500

3000

1 10 100 1000

Chord Length (microns)

#/s

Bottom Sample

Top Sample

Page 149: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

139

0

1

2

3

4

5

6

1 10 100 1000

Chord Length (microns)

#/s

Bottom Sample

Top Sample

Figure 6.4: Offline FBRM measurements taken from samples from the top and bottom of the pilot plant

vessel at the end of Batch A

The large degree of segregation observed during batch A indicates that agitating at

130 rpm for the duration of the batch is not adequate to ensure a robust and repeatable

crystallization. Batches B and C during which the agitation was increased to maintain

a constant power per unit volume resulted in much better mixing and no segregation.

Additionally these batches produced an almost identical chord length distribution

indicating repeatability, assuming the impact of changing the addition location (batch

C) was negligible.

In order to assess the effectiveness of the scale-up in terms of the crystal size

predicted, chord length distributions at the end of batch B are compared with the

Page 150: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

140

standard laboratory experiment. In order to effectively compare the chord length

distributions the FBRM data are normalized to allow for the differences in scale.

Figure 6.5 indicates that in general the chord length distribution measured at each

scale is very similar. There are small differences at the fine end of the distribution

where there is slightly more fine material produced at the pilot scale. At the coarse

end of the distribution the distributions are essentially identical with the lab laboratory

scale producing slightly larger material. There is an increase in the number of fine

crystals for the pilot-scale crystallization compared to the laboratory-scale

experiment. Additionally the squared weighted chord length distribution indicates the

size of the largest crystals is essentially identical with the laboratory-scale

crystallization producing slightly larger crystals. Statistics from each distribution are

compared in Table 6.2 and indicate the similarity between the crystallization

conducted at both scales.

Page 151: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

141

0

0.5

1

1.5

2

2.5

3

1 10 100 1000

Chord Length (microns)

#/s

Standard Laboratory

Batch B

0

1

2

3

4

5

1 10 100 1000

Chord Length (microns)

#/s

Standard Laboratory

Batch B

Figure 6.5: Comparison of standard laboratory experiment with batch B (a) unweighted and (b) square

weighted

Page 152: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

142

Table 6.3: Summary of Chord Length Distribution Statistics

The data shown in Table 6.3 is complemented by the PVM images taken for each

crystallization, show in Figure 6.6. The images indicate the similarity in crystal size

for each batch. They also help validate the FBRM measurement and indicate that the

mean square weight is probably a good indicator of the length of the crystals while the

unweighted mean is probably a good indicator of the width of the crystals.

Clearly the scale-up has been effective in terms of producing crystals of a similar size

at both scales at the end of the batch. However the in-situ tools offer the opportunity

monitor the crystallization in its entirety. Figure 6.7 compares FBRM trends in two

size ranges for the laboratory scale experiment and Batch B.

Comparison of the trends show that even though the chord length distributions are

similar at the end of the batch the mechanism through which this is achieved is

different depending on the scale. By studying #/s (1-10 microns) (Figure 6.7) it is

clear that nucleation occurs earlier at the pilot scale. It is difficult to see if the rate of

nucleation was similar at both scales as some probe coating was observed at the pilot

scale just after nucleation. The probe was manually cleaned soon after and the counts

returned to normal. The rate of increase of the large counts (100-1000 microns) is

Median

(μm)

Mean

(μm)

MSQW

(μm)

% counts

(1-5 μm)

% counts

(5-100 microns)

% counts

(100-1000 microns)

Standard

Laboratory

20.53 43.13 222.52 11.9 76.2 11..9

Batch B 19.54 41.12 208.9 15.5 73.4 11.1

Page 153: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

143

very similar for each batch. Undoubtedly agglomeration is playing some part in the

crystallization as shown in the PVM images (Figure 6.7 (b)). The fact that the FBRM

trends are similar at both scales indicates suitable process parameters have been

chosen and the change in mixing conditions between the scales does not impact the

kinetics of the crystallization.

Figure 6.6: PVM images of (a) Pilot Scale (b) Laboratory Scale.

a

b

Page 154: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

144

0

500

1000

1500

2000

2500

3000

0 500 1000 1500 2000 2500 3000 3500 4000

Time (s)

#/s

(1-1

0)

Pilot-Scale

Lab-Scale

Figure 6.7: Comparison of #/s (1-10 microns) between batch B and lab scale

0

100

200

300

400

500

600

0 500 1000 1500 2000 2500 3000 3500 4000

Time (s)

#/s

(100-1

000)

Pilot-Scale

Lab-Scale

Figure 6.8: Comparison of #/s (100-1000 microns) between batch B and lab scale

Page 155: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

145

6.4 DISCUSSION

Crystallization scale-up requires knowledge of every impact the change in scale has

on the crystallization. There are liquid-liquid interactions to be taken into account i.e.

how the antisolvent is incorporated into the bulk solution. This is influenced by scale,

geometry, type of impeller, agitation intensity, addition location, addition rate, feed

pipe diameter and solution and antisolvent concentration (i.e. liquid viscosities and

densities). All of these factors can influence the micro-, meso- and macromixing

properties of the system and hence the supersaturation profiles within the system. In

turn the supersaturation profile impacts the nucleation, dissolution, growth and

agglomeration rates. These in turn impact on the yield, purity and size of the crystal

product. With this in mind it becomes clear that to try and scale an antisolvent

crystallization it is impossible to do so based on a single traditional scale up parameter

such as blend time or power per unit volume as no single parameter can take into

account all of the factors.

Equally important is the fact that a change in scale impacts on the physical aspects of

the crystallization. Thus, segregation is more likely to occur at larger scales where the

mixing is not sufficient to fully suspend the crystal slurry (as was the case in Batch A

of this study). Similarly, attrition and secondary nucleation rates will be affected by a

change in scale. In the lab where the surface area to volume ratio is high, attrition and

secondary nucleation rates are more sometime significantly elevated compared to

larger scales where this ratio is low.

Page 156: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

146

With these points in mind care should taken when scaling crystallization processes,

specifically antisolvent crystallization, not to make assumptions based on a single

mixing parameter. In many cases by combining a sound understanding of the

crystallization system with a degree of common sense, effective scale-up can be

achieved, as observed in this study. For example, by choosing a pitched blade

impeller better crystal suspension may be achieved but this may be at the cost of

entraining air into the system for a prolonged period. By studying the impact of

mixing conditions on attrition levels in the laboratory the choice of impeller may

become easier.

In order to scale-up effectively, intimate knowledge of every impact the change in

scale will have on the crystallization is necessary. Current research in this area

focuses on the use of CFD coupled with the population balance equations. This

approach has the ability to take into account almost every parameter that can affect

the liquid-liquid interactions. However it remains some way off in terms of predicting

crystallization mechanisms such as growth rates, nucleation rates and agglomeration

rates that are influenced in competing directions by the mixing effects. Despite this

the approach appears to be the only one available that can fully take into account all

of the variables that impact crystallization scale-up.

Before this becomes a reality there are still some common sense approaches that can

be taken to enhance the likelihood of scale-up success. For antisolvent crystallization

homogeneity is vital. By eliminating areas of local supersaturation problems such as

elevated nucleation rates, incorrect polymorph nucleation and phase separation can be

avoided. This can be achieved be adding slowly, at high agitation in a good addition

Page 157: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

147

location. However, care must be taken to ensure that these parameters are suitable.

For example the addition rate must not be so slow that the batch-time is compromised

and throughput reduced. The agitation rate must not be so high as to introduce

significant attrition and/or secondary nucleation. If this is a problem there are other

techniques that can be used to improve dispersion such as adding anti-solvent through

a narrow pipe in the form of a jet facilitating dispersion. Finally the choice of addition

point can be aided by employing CFD to model how the anti-solvent will be

dispersed.

6.5 CONCLUSION

The pilot-scale crystallization of benzoic acid from ethanol-water mixtures using

water as an anti-solvent has been studied. Adequate scale-up in terms of particle size

was achieved. CFD was used to assess the ideal addition location and it was observed

that this point changed with the increase in the liquid volume. Changing the addition

location during a batch however had no effect on the final crystal product. Effective

scale-up in terms of particle size was achieved by using extensive data gathered at the

laboratory scale to design a scale-up strategy. Results at the lab scale showed that a

slow addition rate, intense agitation and a precise addition location (based on CFD)

allowed for a robust and repeatable crystallization. These conditions were mimicked

at the pilot scale and results showed that the final product particle size, in terms of

chord length distribution and in process images, was almost identical. This was

achieved for a vessel that was geometrically dissimilar to laboratory scale vessel

where the initial characterization studies were conducted. This study shows that scale-

up is possible without relying on traditional scale-up parameters such as power per

Page 158: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

148

unit volume. And it is proposed that for anti-solvent crystallization using such

parameters may not be the most useful approach for crystallization given the

numerous mixing interactions that impact on crystallization. Current research aimed

at modelling the complete crystallization process using computational fluid dynamics

coupled with the population balance will in the future be useful but is still limited.

Currently, common sense allied with sound knowledge of the crystallization

mechanisms for a given system are the most useful tools for effective crystallization

scale-up.

Page 159: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

149

7. THESIS CONCLUSIONS

The anti-solvent crystallization of benzoic acid from ethanol-water solutions using

water as the anti-solvent at multiple scales has been presented. A thorough review of

the literature indicated that anti-solvent crystallization has been neglected, in favour

of the study of cooling crystallization, and, in the cases where anti-solvent

crystallization is studied the model system used is often inorganic. While some

crystallization process parameters have been studied in detail (addition rate,

concentration and agitation intensity) many have been neglected, including, anti-

solvent addition location and supersaturation.

The starting point for crystallization characterization is the solubility curve and

metastable zone width. For this study, the solubility was measured using three

separate techniques – gravimetric analysis, FBRM, and ATR-FTIR – and the results

were in good agreement. The gravimetric method proved a simple and cost effective

technique, however, it can be time consuming and there was some disagreement

between the results of the “liquid” and “solid” analysis. The ATR-FTIR technique

required the development of a calibration model before the solubility could be

measured. However, once the model was developed and validated the solubility was

measured rapidly. The ATR-FTIR method can also be used in other ways to study

crystallization experiments – most notably to monitor supersaturation (Chapter 5) –

and in this respect can offset the significant investment required for implementation.

The FBRM technique can be used to rapidly measure solubility by quickly identifying

the point of crystal dissolution. It does not require any calibration and its application

as the standard in process tool for crystallization characterization makes it an obvious

Page 160: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

150

choice for measuring solubility. In order to validate the experimental results garnered

using the three techniques the UNIQAC method was used to model the system. The

model results correlated well with the experimental values even in the region at low

anti-solvent concentration where solubility increased with anti-solvent concentration.

This solubility measurement work highlighted the varied methods available to

investigators wishing to gather accurate solubility information and also provided the

platform necessary for further characterization of this system including the metastable

zone width and nucleation kinetics under varied process conditions.

The metastable zone width was measured at the 500 mL scale, under various process

conditions including addition rate and location, and, agitation intensity. As expected

increasing the addition rate resulted in a wider metastable zone width, however, the

influence of addition location proved to be extremely important with variable and

sometimes negative metastable zone widths occurring when anti-solvent was added

close the wall of the vessel. This problem was exacerbated at low agitation intensities.

The inadequate incorporation of the antisolvent into the bulk solution was identified

as a possible cause of this variability and a study of the vessel using computational

fluid dynamics confirmed this to be the case. Nucleation kinetics were estimated for

the well mixed case where anti-solvent was added close to the impeller and the

influence of agitation intensity on the kinetics was quantified. A unique expression

relating the nucleation rate to the agitation intensity was developed by modifying

classic nucleation theory for the anti-solvent case. This study highlighted the sensitive

and sometimes counterintuitive nature of crystallization processes. By moving the

addition location a very short distance a significant difference in the repeatability and

robustness of the metastable zone width was observed. The sensitivity of this system

Page 161: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

151

to such small changes highlights some of the difficulties associated with scale-up and

achieving a consistent process and product performance.

In order to facilitate efficient scale-up a series of crystallizations experiments were

undertaken at the 500 mL scale to identify suitable operating parameters that would

ensure a repeatable and robust crystallization. The in situ tools offered the opportunity

to monitor each crystallization in real time for the full duration of the experiment and

allowed the comparison of each crystallization over the entirety of each run. High

agitation intensity and a slow addition rate resulted in a very repeatable crystallization

in terms of supersaturation, monitored using ATR-FTIR, and particle size, monitored

using FBRM. The in situ data also allowed the elucidation of growth rate kinetics for

the system. By combining the ATR-FTIR supersaturation information with growth

rate data, generated using the FBRM, a kinetic expression for growth was rapidly

estimated. This method highlights the opportunity to monitor growth rate kinetics in

situ and in real time for any crystallization process. The ability to do this may prove

useful for the control of crystallization processes in the lab as well as at production

scales. The ability to modify the growth rate in real time by tuning process parameters

would be a challenging but potentially valuable alternative to current production scale

crystallization practices.

Current crystallization practice requires the efficient and effective scale-up of the

crystallization from the lab to production, typically with the aid of an intermediate

pilot scale. In order to study the sale-up behaviour of this system three 70 L pilot scale

batches were conducted in a geometrically dissimilar vessel under different process

conditions. The process conditions implemented were chosen by combining the

Page 162: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

152

detailed information gathered over the course of this study - accurate solubility

measurement, the impact of agitation on nucleation kinetics, and the choice of suitable

process parameters to ensure repeatability. Computational fluid dynamics was also

used to model mixing patterns in the vessel allowing suitable addition locations to be

chosen. A successful scale-up was achieved based on repeatability, short batch time,

and a similar yield and particle size as the laboratory scale process. The FBRM

distributions and PVM images clearly showed the similarity in particle size and

morphology at both scales. However, significant segregation and premature

nucleation in the pilot scale vessel were observed indicating further work would be

needed before scaling up to the full production scale.

Page 163: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

153

8. NOMENCLATURE

A anti-solvent concentration (g antisolvent/ g solvent)

c concentration (g/g)

cs saturation concentration

f fugacity

g growth rate order

G growth rate

i component in UNIQUAC method

j component in UNIQUAC method

J nucleation rate

k component in UNIQUAC method

kn nucleation rate constant

kg growth rate constant

n nucleation order

N agitation intensity (rpm)

m mass ratio of anti-solvent to solvent (-)

P pressure

R anti-solvent addition rate (gs-1

)

s mass of solute (g)

T temperature

Tm melting point

t time (s)

w mass of solvent (g)

Page 164: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

154

Symbols

α shape factor

β agitation parameter for nucleation kinetics

ΔC supersaturation (g/g)

ΔAmax metastable zone width

ΔAmax latent heat of fusion

Γ liquid activity coefficient

ρ density

τ binary interaction parameter

Page 165: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

155

9: REFERENCES

Abbas, A., D. Nobbs, and J. A. Romagnoli, "Investigation of on-line optical particle

characterization in reaction, and cooling crystallization systems. Current state of the

art," Measurement Science & Technology 13 (3): 349-356 (2002)

Akiti, O. and P. M. Armenante, "Experimentally-validated micromixing-based CFD

model for fed-batch stirred-tank reactors," Aiche Journal 50 (3): 566-577 (2004)

Alfano, J. C. et al., "Polyelectrolyte-induced aggregation of microcrystalline

cellulose: Reversibility and shear effects," Journal of Colloid and Interface Science

223 (2): 244-254 (2000)

Barata, P. A. and M. L. Serrano, "Salting-out precipitation of potassium dihydrogen

phosphate (KDP) .2. Influence of agitation intensity," Journal of Crystal Growth 163

(4): 426-433 (1996)

Barata P.A., Serrano M.L., “Salting-out precipitation of potassium dihydrogen

phosphate (KDP). III. Growth Process”, Journal of Crystal Growth, 194, 1, 101-108

(1998)

Barata P.A., Serrano M.L., “Salting-out precipitation of potassium dihydrogen

phosphate (KDP). IV. Characterization of Final Product”, Journal of Crystal Growth,

194, 1, 109-118 (1998)

Barrett, P. and B. Glennon, "In-line FBRM monitoring of particle size in dilute

agitated suspensions," Particle & Particle Systems Characterization 16 (5): 207-211

(1999)

Page 166: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

156

Barrett, P. and B. Glennon, "Characterizing the metastable zone width and solubility

curve using lasentec FBRM and PVM," Chemical Engineering Research & Design 80

(A7): 799-805 (2002)

Barrett, P. et al., "A review of the use of process analytical technology for the

understanding and optimization of production batch crystallization processes,"

Organic Process Research & Development 9 (3): 348-355 (2005)

Beckmann, W., "Nucleation phenomena during the crystallisation and precipitation of

Abecarnil," Journal of Crystal Growth 199: 1307-1314 (1999)

Blanco, A. et al., "Focused beam reflectant measurement as a tool to measure

flocculation," Tappi Journal 1 (10): 14-20 (2002)

Bohlin, M. and A. C. Rasmuson, "Importance of macromixing in batch cooling

crystallization," Aiche Journal 42 (3): 691-699 (1996)

Borissova, A. et al., "Examination of the semi-batch crystallization of benzophenone

from saturated methanol solution via aqueous antisolvent drowning-out as monitored

in-process using ATR FTIR spectroscopy," Crystal Growth & Design 4 (5): 1053-

1060 (2004)

Charmolue, H. and R. W. Rousseau, "L-Serine Obtained by Methanol Addition in

Batch Crystallization," Aiche Journal 37 (8): 1121-1128 (1991)

Choi, YC., Morgenroth, E., “Monitoring biofilm detachment under dynamic changes

in shear stress using laser-based particle size analysis and mass fractionation”, Water

Sci Technol., ;47, 5, 69-76 (2003)

Page 167: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

157

Clarke, M. A. and P. R. Bishnoi, "Determination of the intrinsic rate constant and

activation energy of CO2 gas hydrate decomposition using in-situ particle size

analysis," Chemical Engineering Science 59 (14): 2983-2993 (2004)

David, R., "General rules for prediction of the intensity of micromixing effects on

precipitations," Powder Technology 121 (1): 2-8 (2001)

De Clercq, B., P. A. Lant, and P. A. Vanrolleghem, "Focused beam reflectance

technique for in situ particle sizing in wastewater treatment settling tanks," Journal of

Chemical Technology and Biotechnology 79 (6): 610-618 (2004)

Derdour, L. et al., "Real-time evaluation of the concentration of impurities during

organic solution crystallization," Powder Technology 129 (1-3): 1-7 (2003)

Doki, N. et al., "Process control of seeded batch cooling crystallization of the

metastable alpha-form glycine using an in-situ ATR-FTIR spectrometer and an in-situ

FBRM particle counter," Crystal Growth & Design 4 (5): 949-953 (2004)

Dowding, P. J., J. W. Goodwin, and B. Vincent, "Factors governing emulsion droplet

and solid particle size measurements performed using the focused beam reflectance

technique," Colloids and Surfaces A-Physicochemical and Engineering Aspects 192

(1-3): 5-13 (2001)

Dunuwila, Dilum D., I. I. Carroll, and Kris A. Berglund, "An investigation of the

applicability of attenuated total reflection infrared spectroscopy for measurement of

solubility and supersaturation of aqueous citric acid solutions," Journal of Crystal

Growth 137 (3-4): 561-568 (1994)

Page 168: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

158

Dunuwila, Dilum D. and Kris A. Berglund, "ATR FTIR spectroscopy for in situ

measurement of supersaturation," Journal of Crystal Growth 179 (1-2): 185-193

(1997)

Feng, L. L. and K. A. Berglund, "ATR-FTIR for determining optimal cooling curves

for batch crystallization of succinic acid," Crystal Growth & Design 2 (5): 449-452

(2002)

Franke, D. and W. Gosele, "Modeling precipitation with a hydrodynamic approach,"

Chemie Ingenieur Technik 71 (11): 1245-1252 (1999)

Fujiwara, M. et al., "Paracetamol crystallization using laser backscattering and ATR-

FTIR spectroscopy: Metastability, agglomeration, and control," Crystal Growth &

Design 2 (5): 363-370 (2002)

Gee, E.A., Cunnigham, William K., Heindl, R., “Pilot Plants: Production of Iron-Free

Alum”, Ind. Eng. Chem., 39, 9, 1178 - 1188 (1947)

Gong, J. B. et al., "Simulation and scale-up of barium sulphate precipitation process

using CFD modeling," Chinese Journal of Chemical Engineering 13 (2): 167-172

(2005)

Gong, J. B. et al., "Simulation and scale-up of barium sulphate precipitation process

using CFD modeling," Chinese Journal of Chemical Engineering 13 (2): 167-172

(2005)

Granberg, Roger A. et al., "Primary nucleation of paracetamol in acetone-water

mixtures," Chemical Engineering Science 56 (7): 2305-2313 (2001)

Page 169: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

159

Granberg R. A., Rasmuson, A.C., “Solubility of Paracetamol in Pure Solvents”, J.

Chem. Eng. Data, 44 (6), 1391 -1395 (1999)

Groen, H., Roberts, K.J., “Application of ATR FTIR spectroscopy for on-line

determination of solute concentration”, Proceedings of the 14th International

Symposium on Industrial Crystallization (1999)

Groen, H. and K. J. Roberts, "Nucleation, growth, and pseudo-polymorphic behavior

of citric acid as monitored in situ by attenuated total reflection Fourier transform

infrared spectroscopy," Journal of Physical Chemistry B 105 (43): 10723-10730

(2001)

Heath, A. R. et al., "Estimating average particle size by focused beam reflectance

measurement (FBRM)," Particle & Particle Systems Characterization 19 (2): 84-95

(2002)

Heffels, C. et al., "Control of particulate processes by optical measurement

techniques," Particle & Particle Systems Characterization 15 (5): 211-218 (1998)

Holmback, X. and A. C. Rasmuson, "Size and morphology of benzoic acid crystals

produced by drowning-out crystallisation," Journal of Crystal Growth 199: 780-788

(1999).

Hukkanen, E. J. and R. D. Braatz, "Measurement of particle size distribution in

suspension polymerization using in situ laser backscattering," Sensors and Actuators

B-Chemical 96 (1-2): 451-459 (2003)

Page 170: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

160

Jeffers, P. et al., "Focussed beam reflectance measurement (FBRM) monitoring of

particle size and morphology in suspension cultures of Morinda citrifolia and

Centaurea calcitrapa," Biotechnology Letters 25 (23): 2023-2028 (2003)

Kaneko, S. et al., "Effect of supersaturation on crystal size and number of crystals

produced in antisolvent crystallization," Journal of Chemical Engineering of Japan 35

(11): 1219-1223 (2002)

Kitamura, M. and M. Sugimoto, "Anti-solvent crystallization and transformation of

thiazole-derivative polymorphs--I: effect of addition rate and initial concentrations,"

Journal of Crystal Growth 257 (1-2): 177-184 (2003)

Kougoulos, E., A. G. Jones, and M. W. Wood-Kaczmar, "Modelling particle

disruption of an organic fine chemical compound using Lasentec focussed beam

reflectance monitoring (FBRM) in agitated suspensions," Powder Technology 155

(2): 153-158 (2005)

Kougoulos, E., A. G. Jones, and M. W. Wood-Kaczmar, "Estimation of crystallization

kinetics for an organic fine chemical using a modified continuous cooling mixed

suspension mixed product removal (MSMPR) crystallizer," Journal of Crystal

Growth 273 (3-4): 520-528 (2005)

Kougoulos, E. et al., "Use of focused beam reflectance measurement (FBRM) and

process video imaging (PVI) in a modified mixed suspension mixed product removal

(MSMPR) cooling crystallizer," Journal of Crystal Growth 273 (3-4): 529-534 (2005)

Kougoulos, E., A. G. Jones, and M. Wood-Kaczmar, "CFD modelling of mixing and

heat transfer in batch cooling crystallizers - Aiding the development of a hybrid

Page 171: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

161

predictive compartmental model," Chemical Engineering Research & Design 83

(A1): 30-39 (2005)

Kougoulos, E., A. G. Jones, and M. W. Wood-Kaczmar, "A hybrid CFD

compartmentalization modeling framework for the scaleup of batch cooling

crystallization processes," Chemical Engineering Communications 193 (8): 1008-

1023 (2006)

Langston, P. A. et al., "Particle and droplet size analysis from chord measurements

using Bayes' theorem," Powder Technology 116 (1): 33-42 (2001)

Law, D. J., A. J. Bale, and S. E. Jones, "Adaptation of focused beam reflectance

measurement to in-situ particle sizing in estuaries and coastal waters," Marine

Geology 140 (1-2): 47-59 (1997)

Lewiner, F. et al., "Improving batch cooling seeded crystallization of an organic

weed-killer using on-line ATR FTIR measurement of supersaturation," Journal of

Crystal Growth 226 (2-3): 348-362 (2001)

Lewiner, F. et al., "On-line ATR FTIR measurement of supersaturation during

solution crystallization processes. Calibration and applications on three solute/solvent

systems," Chemical Engineering Science 56 (6): 2069-2084 (2001)

Lewiner, F. et al., "An Online Strategy To Increase the Average Crystal Size during

Organic Batch Cooling Crystallization," Industrial & Engineering Chemistry

Research 41 (5): 1321-1328 (2002)

Page 172: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

162

Li, M., D. Wilkinson, and K. Patchigolla, "Comparison of particle size distributions

measured using different techniques," Particulate Science and Technology 23 (3):

265-284 (2005)

168. Li, M. Z., D. Wilkinson, and K. Patchigolla, "Determination of non-spherical

particle size distribution from chord length measurements. Part 2: Experimental

validation," Chemical Engineering Science 60 (18): 4992-5003 (2005)

Li, M. Z. and D. Wilkinson, "Determination of non-spherical particle size distribution

from chord length measurements. Part 1: Theoretical analysis," Chemical Engineering

Science 60 (12): 3251-3265 (2005)

Liotta, V. and V. Sabesan, "Monitoring and feedback control of supersaturation using

ATR-FTIR to produce an active pharmaceutical ingredient of a desired crystal size,"

Organic Process Research & Development 8 (3): 488-494 (2004)

Liu, Weidong, Nigel N. Clark, and Ali Ihsan Karamavruc, "Relationship between

bubble size distributions and chord-length distribution in heterogeneously bubbling

systems," Chemical Engineering Science 53 (6): 1267-1276 (1998)

Loan, M. et al., "Iron oxy-hydroxide crystallization in a hyd ro metallurgical residue,"

Journal of Crystal Growth 235 (1-4): 482-488 (2002)

Löffelmann, M., Mersmann, A., “A New Method for Supersaturation Measurement”,

International Journal of Modern Physics B, 16, 01-02, 391-398 (2002)

Madras, G. and B. J. Mccoy, "Mixing effects on particle precipitation," Industrial &

Engineering Chemistry Research 44 (14): 5267-5274 (2005)

Page 173: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

163

Marcant, B. and R. David, "Experimental-Evidence for and Prediction of

Micromixing Effects in Precipitation," Aiche Journal 37 (11): 1698-1710 (1991)

Marchisio, D. L. et al., "Effect of fluid dynamics on particle size distribution in

particulate processes," Chemical Engineering & Technology 29 (2): 191-199 (2006)

McDonald, K. A., A. P. Jackman, and S. Hurst, "Characterization of plant suspension

cultures using the focused beam reflectance technique," Biotechnology Letters 23 (4):

317-324 (2001)

Monnier, O. et al., "Model identification of batch cooling crystallizations through

calorimetry and image analysis," Chemical Engineering Science 52 (7): 1125-1139

(1997)

Mori, Y. et al., "CFD study of local crystal growth rate in a continuous suspension

crystallizer," Journal of Chemical Engineering of Japan 35 (11): 1178-1187 (2002)

Mullin, J.W., Garside, J., Unahabhokha, R., “Diffusivities of Ammonium and

Potassium Alums in Aqueous Solutions”, Journal of Applied Chemistry of the USSR,

(1965)

Mullin, J., Teodossiev, N., Söhnel, O., “Potassium sulphate precipitation from

aqueous solution by salting-out with acetone”, Chemical Engineering and Processing,

26, 2, 93-99 (1989)

Mydlarz, J. and A. G. Jones, "Crystallization and Agglomeration Kinetics During the

Batch Drowning-Out Precipitation of Potash Alum with Aqueous Acetone," Powder

Technology 65 (1-3): 187-194 (1991)

Page 174: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

164

Nyvlt. J., “Kinetics of nucleation in solutions”, Journal of Crystal Growth, 3, 377-

383, (1968)

O‟Grady D., et al, 'The Effect of Mixing on the Metastable Zone Width and Nucleation

Kinetics In the Anti-solvent Crystallization of Benzoic Acid, Chemical Engineering Research

and Design, Transactions IChemE part A, 85 (7) 945-953, 2007

O‟Grady D. & Glennon B.,, „Solubility Measurement for an Anti-solvent Crystallization

System Using Gravimetric Analysis, ATR-FTIR and FBRM', Crystal Growth and Design, in

press

O'Rourke, A. M. and P. F. MacLoughlin, "A comparison of measurement techniques

used in the analysis of evolving liquid-liquid dispersions," Chemical Engineering and

Processing 44 (8): 885-894 (2005)

O'Sullivan, B. et al., "In situ monitoring of polymorphic transitions," Organic Process

Research & Development 7 (6): 977-982 (2003)

O'Sullivan, B. and B. Glennon, "Application of in situ FBRM and ATR-FTIR to the

monitoring of the polymorphic transformation of D-mannitol," Organic Process

Research & Development 9 (6): 884-889 (2005)

Oosterhof, H., G. J. Witkamp, and G. M. van Rosmalen, "Some antisolvents for

crystallisation of sodium carbonate," Fluid Phase Equilibria 155 (2): 219-227 (1999)

Paul, E. L., H. H. Tung, and M. Midler, "Organic crystallization processes," Powder

Technology 150 (2): 133-143 (2005)

Page 175: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

165

Pearson, A. P., B. Glennon, and P. M. Kieran, "Comparison of morphological

characteristics of Streptomyces natalensis by image analysis and focused beam

reflectance measurement," Biotechnology Progress 19 (4): 1342-1347 (2003)

Pearson, A., Glennon, B., Kieran, P., “Monitoring of cell growth using the focused

beam reflectance method”, Journal of Chemical Technology & Biotechnology, 79, 10,

1142 -1147 (2004)

Pettersson, M. and A. C. Rasmuson, "Hydrodynamics of suspensions agitated by

pitched-blade turbine," Aiche Journal 44 (3): 513-527 (1998)

Pina, C. M., L. FernandezDiaz, and M. Prieto, "Crystallization of beta''-LiNH4SO4

and (NH4)(2)SO4 in gels: Growth morphology and epitaxy phenomena," Journal of

Crystal Growth 177 (1-2): 102-110 (1997)

Pina, C. M. et al., "Metastability in drowning-out crystallisation: precipitation of

highly soluble sulphates," Journal of Crystal Growth 222 (1-2): 317-327 (2001)

Pohorecki, R. and J. Baldyga, "New Model of Micromixing in Chemical Reactors .1.

General Development and Application to a Tubular Reactor," Industrial &

Engineering Chemistry Fundamentals 22 (4): 392-397 (1983)

Rohani, S. and J. Baldyga, "Micromixing Described in Terms of Inertial Convective

Disintegration of Large Eddies and Viscous Convective Interactions Among Small

Eddies .2. Semibatch and Continuous Stirred Tank Reactors," Chemical Engineering

Science 42 (11): 2611-2619 (1987)

Rohani, S. and J. Baldyga, "Mathematical-Modeling of Micromixing in Liquid

Liquid-Systems," Hungarian Journal of Industrial Chemistry 16 (1): 85-97 (1988)

Page 176: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

166

Ruf, A., J. Worlitschek, and M. Mazzotti, "Modeling and experimental analysis of

PSD measurements through FBRM," Particle & Particle Systems Characterization 17

(4): 167-179 (2000)

Scott, C. and S. Black, "In-line analysis of impurity effects in crystallization,"

Organic Process Research & Development 9 (6): 890-893 (2005)

Sha, Z. et al., "Application of CFD simulation to suspension crystallization-actors

affecting size-dependent classification," Powder Technology 121 (1): 20-25 (2001)

Shi, B., W. J. Frederick, and R. W. Rousseau, "Effects of calcium and other ionic

impurities on the primary nucleation of burkeite," Industrial & Engineering

Chemistry Research 42 (12): 2861-2869 (2003)

Simmons, M. J. H., P. A. Langston, and A. S. Burbidge, "Particle and droplet size

analysis from chord distributions," Powder Technology 102 (1): 75-83 (1999)

Sistare, F., St. Pierre Berry, L., Mojica, C., “Process Analytical Technology: An

Investment in Process Knowledge”, Org. Process Res. Dev., 9, 3, 332 -336 (2005)

Sparks, R.G., Dobbs, C.L., “The use of laser backscatter instrumentation for the on-

line measurement of the particle size distribution of emulsions” 10, 5, 279-289 (1993)

Tadayyon, A., Rohani, S., “Determination of Particle Size Distribution by Par-Tec®

100: Modeling and Experimental Results” Particle and Particle Systems

Characterization, 15, 3, 127-135 (1998)

Tadayyon, A. and S. Rohani, "Control of fines suspension density in the fines loop of

a continuous KCI crystallizer using transmittance measurement and an FBRM (R)

probe," Canadian Journal of Chemical Engineering 78 (4): 663-673 (2000)

Page 177: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

167

Takiyama, H., T. Otsuhata, and M. Matsuoka, "Morphology of NaCl crystals in

drowning-out precipitation operation," Chemical Engineering Research & Design 76

(A7): 809-814 (1998)

Tavare, N. S., "Mixing, Reaction, and Precipitation - Interaction by Exchange with

Mean Micromixing Models," Aiche Journal 41 (12): 2537-2548 (1995)

Togkalidou, T. et al., "Solute concentration prediction using chemometrics and ATR-

FTIR spectroscopy," Journal of Crystal Growth 231 (4): 534-543 (2001)

Togkalidou, T. et al., "Solution concentration prediction for pharmaceutical

crystallization processes using robust chemometrics and ATR FTIR spectroscopy,"

Organic Process Research & Development 6 (3): 317-322 (2002)

Torbacke, M. and A. C. Rasmuson, "Influence of different scales of mixing in

reaction crystallization," Chemical Engineering Science 56 (7): 2459-2473 (2001)

Torbacke, M. and A. C. Rasmuson, "Mesomixing in semi-batch reaction

crystallization and influence of reactor size," Aiche Journal 50 (12): 3107-3119

(2004)

Vittorf, N.M. “Separation of aluminium-sulphate from iron by means of alcohol”,

Institut prikladnoi mineralogii i petrografii, Moscow. Transactions (1924)

Wei, H et al., "Computational fluid dynamics modeling of the precipitation process in

a semibatch crystallizer," Industrial & Engineering Chemistry Research 40 (23):

5255-5261 (2001)

Page 178: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

168

Woo, X. Y. et al., "Simulation of mixing effects in antisolvent crystallization using a

coupled CFD-PDF-PBE approach," Crystal Growth & Design 6 (6): 1291-1303

(2006)

Worlitschek, J. and M. Mazzotti, "Choice of the focal point position using lasentec

FBRM," Particle & Particle Systems Characterization 20 (1): 12-17 (2003)

Worlitschek, J. and M. Mazzotti, "Model-based optimization of particle size

distribution in batch-cooling crystallization of paracetamol," Crystal Growth &

Design 4 (5): 891-903 (2004)

Worlitschek, J., Hocker, T., Mazzotti, M., “Restoration of PSD from Chord Length

Distribution Data using the Method of Projections onto Convex Sets”, Particle &

Particle Systems Characterization, 22, 2, 81-98 (2005)

Wynn, E. J. W., "Relationship between particle-size and chord-length distributions in

focused beam reflectance measurement: stability of direct inversion and weighting,"

Powder Technology 133 (1-3): 125-133 (2003)

Yu, Z. Q., R. B. H. Tan, and P. S. Chow, "Effects of operating conditions on

agglomeration and habit of paracetamol crystals in anti-solvent crystallization,"

Journal of Crystal Growth 279 (3-4): 477-488 (2005)

Zauner, R. and A. G. Jones, "Scale-up of continuous and semibatch precipitation

processes," Industrial & Engineering Chemistry Research 39 (7): 2392-2403 (2000)

Zauner, R. and A. G. Jones, "Scale-up of Continuous and Semibatch Precipitation

Processes," Industrial & Engineering Chemistry Research 39 (7): 2392-2403 (2000)

Page 179: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

169

Zauner, R. and A. G. Jones, "Mixing effects on product particle characteristics from

semi-batch crystal precipitation," Chemical Engineering Research & Design 78 (A6):

894-902 (2000).

Zauner, R. and A. G. Jones, "On the influence of mixing on crystal precipitation

processes - application of the segregated feed model," Chemical Engineering Science

57 (5): 821-831 (2002)

Zhu, Y. et al., "Study of the solubility and the metastable zone of 1,3-

dihydroxyacetone for the drowning-out process," Journal of Crystal Growth 257 (3-

4): 370-377 (2003).

Zijlema, T. G. et al., "The production of pure NaCl by the antisolvent crystallisation

of NaCl center dot 2H(2)O and a consecutive recrystallisation step," Journal of

Crystal Growth 199: 789-795 (1999)

Zijlema, T. G. et al., "Antisolvent crystallization as an alternative to evaporative

crystallization for the production of sodium chloride," Industrial & Engineering

Chemistry Research 39 (5): 1330-1337 (2000)

Page 180: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

170

Page 181: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

171

APPENDIX A

APPLICATION OF COMPUTATION FLUID DYNAMICS TO EXPLAIN

VARIATION IN THE MSZW FOR ANT-SOLVENT ADDITION

CRYSTALLIZATION

Des O‟ Grady, Mark Barrett, Eoin Casey, Brian Glennon

Chapter 3 focuses on the study of the impact of process parameters on the MSZW of

an anti-solvent system and chapter 5 looks at the scale up of the crystallization to the

pilot-plant. This technical note will focus on the application of velocity profiles within

crystallization vessels at both scales, predicted using Computational Fluid Dynamics

(CFD), to model crystallization behaviour. This work was carried out by Mark Barrett

and Eoin Casey in conjunction with Des O‟Grady and Brian Glennon. A more in-

depth analysis of the use of CFD to model mixing for this crystallization system can

be found elsewhere (O‟Grady et al., 2007)

1. Abstract

Computational Fluid Dynamics is used to model mixing conditions in a 500 mL

vessel and a 70L vessel used to crystallize benzoic acid from ethanol water mixtures

using water as the anti-solvent. At the 500mL scale, anti-solvent addition close to the

impeller results in consistent MSZWs at all addition rates. However, when the

addition location is close to the vessel wall the MSZW is variable and premature

nucleation is common. CFD modelling indicated high downward velocities close to

Page 182: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

172

the impeller but upward velocities close to the wall. This suggests that anti-solvent

added close to the impeller is more rapidly incorporated into the vessel, limiting the

potential for locally high supersaturation. However, close to the wall the anti-solvent

is less rapidly incorporated and local regions of high supersaturation may be expected.

This provides an explanation for the variations in MSZW behaviour observed. Similar

velocity profiles were calculated for the 70 L pilot plant vessel in order to predict

optimal addition location prior to crystallization. The profiles indicate that the optimal

addition location changes as the volume in the reactor increases.

2. Introduction

Computational Fluid Dynamics have been used to model semi-batch reactive

crystallizations and has been used to predict the effect of mixing on the

supersaturation distribution and the resulting size of the crystals (Wei et al., 2001,

Baldyga & Orciuch, 2001). It has also been used to study the effect of mixing on

product yield (Akiti & Armenante, 2004). CFD has also been used in conjunction

with crystallization kinetics and solubility data to simultaneously solve the mass and

population balances for a reactive crystallization, allowing the impact of feed rate,

agitation intensity, feed point and feed tube diameter on nucleation rate and crystal

size to be illustrated (Zauner and Jones, 2001). While the focus of CFD research has

been on reactive crystallization there is clearly scope for the use of the technique to

model mixing in anti-solvent crystallization processes.

In this study, CFD is used to generate theoretical velocity profiles for the vessel prior

to nucleation, in an effort to understand the observed trends in metastable zone width.

Page 183: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

173

Velocity in the z-direction was chosen as a suitable parameter for assessing the

mixing regime, as movement of the liquid in this plane is vital for the incorporation of

anti-solvent into the bulk solution.

3. Computation Fluid Dynamics Model

To investigate the velocity profiles prior to nucleation, the system is treated as a

single phase and the possible solid-liquid interactions after nucleation are not

considered. Initially, the 500 ml vessel and its complex baffling system (the FBRM,

REACT-IR and temperature probes) are constructed within GAMBIT 2.1.6. GAMBIT

is a software package designed to aid in the construction and meshing of systems for

computational fluid dynamics (CFD). The precise geometry and dimensions of the

system (pitch blade impeller, probes etc.) are created in the graphical user interface

allowing for the meshing of the vessel and its internals, along with the assigning of

zone types and system specifics (i.e. fluid viscosity, internal reactor temperature and

agitation speed).

A large number of individual volumes created in the construction of the geometry are

then meshed. The system contains 823 individual volumes and 475,000 cells. An

unstructured hexahedral meshing scheme is applied, as less numerical diffusion errors

are evident then in a tetrahedral-based mesh. The mesh created in GAMBIT is then

exported to FLUENT 6.1.22 for solution of the momentum and continuity equations

for the turbulent flow within the crystallizer. The „Multiple Reference Frame‟ (MRF)

approach is applied to the system and the flow equations solved using the SIMPLE

algorithm. The Reynolds number for the agitation speeds assessed lie between

12,500 (325 rpm) and 18,500 (475rpm). The standard k-epsilon turbulence model

Page 184: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

174

was chosen to model the flows. In addition, a no-slip boundary condition was

imposed on all walls and the free liquid surface is modelled with no vortex, a zero-

flux and zero-stress conditions. To validate the model the tip speed of the impeller is

calculated and compared to the tip speed calculated using the model. The model

underestimated the tip speed by about 3% at 325 and 475 rpm indicating the model is

suitable.

4. Results

4.1 500mL Scale

The difference in the behaviour of the MSZW can be explained in terms of mixing at

different regions of the vessel. Figures 1 and 2 depict velocity in the z-direction

through a cross section of the vessel at 475 rpm and 325 rpm respectively. Velocity in

this direction was chosen, as it will impact on the incorporation of the anti-solvent

into the bulk of the vessel. It is clear from Figure 1 that close to the impeller the

velocity is in the downward direction, but close to the wall the velocity is in the

upward direction. This means that when anti-solvent is added close to the impeller it

is incorporated easily into the bulk solution. However, when it is added close to the

wall this incorporation is more difficult. This may lead to an area of local

supersaturation close to the feed point and premature, variable nucleation as observed

experimentally. Furthermore, analysis of the same velocity profiles at lower agitation

levels indicate that velocity in the downward direction is less intense and velocity in

the upward direction is more pronounced. This explains the greater variability and

Page 185: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

175

narrower MSZWs when anti-solvent is added close to the wall at lower agitation

levels.

Page 186: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

176

Figure 1: Velocity profiles in the z-direction at 475 rpm through a central cross section of the vessel

Figure 2: Velocity Profiles in the z-direction at 325 rpm through a central cross section of the vessel

Page 187: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

177

4.2 70 L Scale

At the 70 L scale CFD models indicate that the optimum feed location changes over

the course of the crystallization. Initially, when the liquid level is close to the plane of

the impeller the most intense downward velocities, and hence optimal addition

location, are found close vessel wall (Figure 3). However, as the volume in the vessel

increases the optimal location moves closer to the impeller shaft (Figure 4). These

results assisted in the choice of addition location in Chapter 5. Two batches were

conducted with the feed point close to the wall for the entire run and for the final

batch the feed point was changed after some time to take advantage of the change in

the optimal feed location.

Page 188: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

178

Figure 3: Velcoity profiles in the z-direction at (a) 32 L and (b) 64 L showing change

in optimum feed location

Page 189: OGrady2007-Multi-Scale Characterization of Anti-Solvent Crystallization By

179

5. Discussion

The experimental results outlined in Chapter 3 can be readily accounted for by

considering the different mixing regimes depending on the process conditions. CFD is

a very useful tool for modelling mixing conditions within a crystallizer and in this

case the CFD model explains the experimental results. Close to the impeller, mixing

conditions favour anti-solvent incorporation leading to consistent MSZW results,

whereas close to the wall anti-solvent incorporation is hindered leading to variability

in the nucleation mechanism and non-repeatable MSZW measurements. This effect is

exacerbated at lower agitation intensities where mixing conditions become even less

favourable.

6. References

Akiti O., & Armenante P., “Experimentally-validated micromixing-based CFD model

for fed-batch stirred-tank reactors”, AIChE, 50 (3) 566-577 (2004)

Bakldyga J., & Orciuch W., „Barium Sulphate precipitation in a pipe – and

experimental study and modelling“, Chemical Engineering Science, 58 (7) 2435-2444

(2001)

Wei H., et al., "Computational Fluid Dynamics Modelling of the Precipitation Process

in a Semibatch Crystallizer," Industrial & Engineering Chemistry Research 40 (23):

5255-5261 (2000)