Anda di halaman 1dari 322

CRYSTALLIZATION CHARACTERISTICS OF NaCl

CRYSTAL SIZE DISTRIBUTION ASSOCIATED


WITH A CMSMPR CRYSTALLIZER

by

Byung Sang Choi

A dissertation submitted to the faculty of


The University of Utah
in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in

Metallurgical Engineering

Metallurgical Engineering

The University of Utah

January 2005
Copyright
c Byung Sang Choi 2005

All Rights Reserved


THE UNIVERSITY OF UTAH GRADUATE SCHOOL

SUPERVISORY COMMITTEE APPROVAL

of a dissertation submitted by

Byung Sang Choi

This dissertation has been read by each member of the following supervisory committee
and by majority vote has been found to be satisfactory.

Chair: H. Y. Sohn

Terry A. Ring

Jan D. Miller

Siva Guruswamy

Edward Trujillo
THE UNIVERSITY OF UTAH GRADUATE SCHOOL

FINAL READING APPROVAL

To the Graduate Council of the University of Utah:

I have read the dissertation of Byung Sang Choi in its final form
and have found that (1) its format, citations, and bibliographic style are consistent and
acceptable; (2) its illustrative materials including figures, tables, and charts are in place;
and (3) the final manuscript is satisfactory to the Supervisory Committee and is ready
for submission to The Graduate School.

Date H. Y. Sohn
Chair: Supervisory Committee

Approved for the Major Department

Jan D. Miller
Chair/Director

Approved for the Graduate Council

David S. Chapman
Dean of The Graduate School
ABSTRACT

Compared to overwhelming technical data available in other advanced technolo-


gies, knowledge about particle technology, especially in particle synthesis from a
solution, is still poor due to the lack of available equipment to study crystallization
phenomena in a crystallizer. Recent technical advances in particle size measurement
such as Coulter counter and laser light scattering have made in/ex situ study of
some of particle synthesis, i.e., growth, attrition, and aggregation, possible with
simple systems. Even with these advancements in measurement technology, to
grasp fully the crystallization phenomena requires further theoretical and technical
advances in understanding such particle synthesis mechanisms. Therefore, it is
the motive of this work to establish the general processing parameters and to
produce rigorous experimental data with reliable performance and characterization
that rigorously account for the crystallization phenomena of nucleation, growth,
aggregation, breakage, and mixing including their variations with time and space
in a controlled continuous mixed-suspension mixed-product removal (CMSMPR)
crystallizer.
This project reports the results and achievements that are; (1) experimental pro-
grams to support the development and validation of the phenomenological models
and generation of laboratory data for the purpose of testing, refining, and validating
the crystallization process, (2) development of laboratory well-mixed crystallizer
system and experimental protocols to generate crystal size distribution (CSD) data,
(3) the effect of feed solution concentration, crystallization temperature, feed flow
rate, and mixing speed, as well as different type of mixer resulting in the evolution of
CSDs with time from a concentrated brine solution, (4) with statistically designed
experiments the effects of processing variables on the resultant particle structure
and CSD at steady state were quantified and related to each of those operating
conditions by studying the detailed crystallization processes, such as nucleation,
growth, and breakage, as well as agglomeration.
The purification of CaCl2 solution involving the crystallization of NaCl from
the solution mixture of CaCl2 , KCl, and NaCl as shipped from the Dow Chemical,
Ludington, in a CMSMPR crystallizer was studied as our model system because of
its nucleation and crystal growth tendencies with less agglomeration. This project
also generated a significant body of experimental data that are available at URL
that is http://www.che.utah.edu/ring/CrystallizationWeb.

v
To my wife, Misun, for her dedication...
CONTENTS

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xx

CHAPTERS
1. FUNDAMENTALS OF CRYSTALLIZATION . . . . . . . . . . . . . . . 1
1.1 Crystallization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Solutions to Population Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 CMSMPR Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Homogeneous Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.2 Heterogeneous Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.3 Secondary Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Crystal Growth Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.1 Surface Energy Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.4.2 Adsorption Layer Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.4.2.1 Surface Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
1.4.2.2 Two-dimensional Growth of Surface Nuclei . . . . . . . . . . . . 32
1.4.2.3 Surface Diffusion and Dislocation . . . . . . . . . . . . . . . . . . . 35
1.4.3 Diffusion-reaction Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.4.3.1 Two-component Diffusion with Different Concentrations . 44
1.4.4 Chemical Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
1.4.5 Impurity Effects on Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
1.4.6 Summary of Crystal Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
1.4.7 Size-dependent Crystal Growth . . . . . . . . . . . . . . . . . . . . . . . . . 52
1.5 Crystal Shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
1.5.1 Heat Flow and Interface Stability . . . . . . . . . . . . . . . . . . . . . . . 55
1.6 Aggregation of Crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
1.7 Attrition of Crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2. EXPERIMENTAL DESIGN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.1 General Consideration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.2 Process Chemistry of Crystallization . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.3 Layout and Design of a CMSMPR Crystallizer . . . . . . . . . . . . . . . . . 63
2.4 Reactor Agitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.5 Automatic System Control and Data Acquisition . . . . . . . . . . . . . . . 65
2.6 Summary of Overall Design Parameters . . . . . . . . . . . . . . . . . . . . . . 67
2.7 Statistical Experimental Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.7.1 Statistics in Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.7.2 Statistical Design of Experiments . . . . . . . . . . . . . . . . . . . . . . . 69
2.8 Design of Batch Breakage and Aggregation Experiments . . . . . . . . . 70
3. EXPERIMENTAL METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.1 Measurement of Solution Properties . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.1.1 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.1.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.1.3 Element Analysis of Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2 Residence Time Distribution Experiment . . . . . . . . . . . . . . . . . . . . . 80
3.3 Crystallization Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.3.1 Precipitation Point Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.3.2 Determination of Reactor Steady State . . . . . . . . . . . . . . . . . . . 84
3.3.3 Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.3.4 Calculation of Yield in a Crystallization Process . . . . . . . . . . . . 87
3.3.5 Kinetic Data Measurement and Utilization . . . . . . . . . . . . . . . . 87
3.4 Characterization of Particle Properties . . . . . . . . . . . . . . . . . . . . . . . 88
3.4.1 Element Analysis of Solid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.4.2 Crystal Size and Shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.4.2.1 SEM and Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.4.2.2 Crystal Size Distribution (CSD) . . . . . . . . . . . . . . . . . . . . 89
3.4.3 Crystal Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.4.3.1 XRD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.5 Measurement of Particle Size Distribution (PSD) . . . . . . . . . . . . . . . 91
3.5.1 Stabilization of NaCl Particles for ex situ PSD measurement . . 93
4. EXPERIMENTAL RESULTS AND DISCUSSION . . . . . . . . . . 94
4.1 Reactor Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.1.1 RTD measurement as a Reactor Characterization Tool . . . . . . . 94
4.2 Solution Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.1 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.2 Viscosity measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.2.3 Element Analysis of Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.3 Crystallization Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.3.1 Precipitation Point Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.3.2 Crystallization of NaCl from CaCl2 , KCl, and NaCl solution . . 99
4.3.3 Crystallization Condition Monitoring . . . . . . . . . . . . . . . . . . . . 101
4.4 Characterization of Particle Properties . . . . . . . . . . . . . . . . . . . . . . . 107
4.4.1 Elemental Analysis of Solid . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
4.4.2 Crystal Size and Shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.4.2.1 SEM and Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.4.3 Crystal Size Distribution (CSD) . . . . . . . . . . . . . . . . . . . . . . . . 114
viii
4.5 Characterization of CSD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.5.1 Reproducibility of CSD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.5.2 Effect of Solution Concentration . . . . . . . . . . . . . . . . . . . . . . . . 119
4.5.3 Effect of Crystallization Temperature . . . . . . . . . . . . . . . . . . . . 122
4.5.4 Effect of Flow Rate (Mean Residence Time) . . . . . . . . . . . . . . . 123
4.5.5 Effect of Mixing Speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.5.6 Effect of Different Mixer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.5.7 Overall Effect on Yield and Lmean . . . . . . . . . . . . . . . . . . . . . . . 129
4.6 Batch Breakage and Aggregation Experiments . . . . . . . . . . . . . . . . . 140
4.6.1 Temperature Effect on Breakage and Aggregation . . . . . . . . . . . 141
4.6.2 Mixing Speed Effect on Breakage and Aggregation . . . . . . . . . . 142
4.6.3 Effect of Different Mixer on Breakage and Aggregation . . . . . . . 144
4.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5. ANALYSIS AND MODEL VALIDATION . . . . . . . . . . . . . . . . . . 152
5.1 Evaluation of Nucleation and Growth Rates . . . . . . . . . . . . . . . . . . . 153
5.2 Application of Hounslows Equation . . . . . . . . . . . . . . . . . . . . . . . . . 154
5.3 Statistical Experimental Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
5.3.1 Analysis of T samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
5.3.2 Analysis of TS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
5.3.3 Analysis of IS sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.4 Influence of Different Mixer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
5.4 Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
5.4.1 Phenomenological Models for Crystallization Modeling . . . . . . . 197
5.4.2 Crystallization Model Validation . . . . . . . . . . . . . . . . . . . . . . . . 198
5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
6. GENERAL CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
6.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211

APPENDICES
A. CRYSTALLIZER GEOMETRY AND CONFIGURATION . . . 212

B. REACTOR CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . . . . 219

C. OPERATING PROCEDURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

D. CHARACTERIZATION OF NaCl CRYSTALS . . . . . . . . . . . . . . 255

E. LS230 SVM PLUS OPERATING SUMMARY . . . . . . . . . . . . . . 274

F. PARTICLE STABILIZATION FOR PSD MEASUREMENT . . 277

ix
G. THERMODYNAMIC MODEL (OLI SYSTEMS) . . . . . . . . . . . . 286

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299

x
LIST OF FIGURES

1.1 Solubility curves for various types of crystallization . . . . . . . . . . . . . . . 2


1.2 A continuous mixed-suspension, mixed-product removal (CMSMPR)
crystallizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Population plots characterizing the crystal size distribution, and the
nucleation and the growth kinetics for a CMSMPR crystallizer . . . . . . 6
1.4 Nucleation scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.5 Classical nucleation theory, dependence of nuclei size on Gibbs free
energy as a function of saturation ratio, S . . . . . . . . . . . . . . . . . . . . . . 13
1.6 Critical nuclei size as a function of saturation ratio, S . . . . . . . . . . . . . 14
1.7 Embryo size distribution function for various values of the saturation
ratio, S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.8 The energetics of crystal growth from solution . . . . . . . . . . . . . . . . . . . 22
1.9 A representation of a growing crystal interfaces . . . . . . . . . . . . . . . . . . 24
1.10 Dependence of fraction of the surface occupied on the Gibbs free
energy for various degrees of surface roughness . . . . . . . . . . . . . . . . . . 26
1.11 Generalized nucleation rate for homogeneous, heterogeneous, and sur-
face nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.12 Two-dimensional growth of surface nuclei . . . . . . . . . . . . . . . . . . . . . . 33
1.13 Two-dimensional growth of surface nuclei by birth and spread model . 34
1.14 Archimedian growth spiral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1.15 Schematic representation of the solute flux to a step site . . . . . . . . . . . 37
1.16 Distribution of solute concentration profile around each kink/step site
for the volume diffusion model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.17 Concentration driving forces in crystallization from solution . . . . . . . . 42
1.18 The effectiveness factor for crystal growth . . . . . . . . . . . . . . . . . . . . . . 46
1.19 Unit cell of NaCl crystal structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
1.20 Sodium chloride crystal grown from pure solution and from a solution
containing 10% urea. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
1.21 Temperature distribution for solidification . . . . . . . . . . . . . . . . . . . . . . 56
1.22 The development of thermal dendrites . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.1 Computer-interfaced experimental set-up . . . . . . . . . . . . . . . . . . . . . . . 66
3.1 A pycnometer with fixed volume (10 ml) . . . . . . . . . . . . . . . . . . . . . . . 77
3.2 Schematic of an atomic absorption spectroscopy (AAS) . . . . . . . . . . . . 80
3.3 Dominant aqueous species and precipitation point for the dominant
solid species crystallizing out from an aqueous solution (100 g) con-
taining CaCl2 (47.63 wt%), KCl (1.07 wt%) and NaCl (0.67 wt%) as
a function of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.4 Steady state measurements with pH and temperature response . . . . . . 85
3.5 Schematic of the PIDS assembly in the Beckman Coulter LS230 par-
ticle size analyzer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.6 Braggs law of diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.1 Viscosity of the solution with varying temperature . . . . . . . . . . . . . . . 97
4.2 NaCl particles after the crystallization at 30.60.18 C (SEM) . . . . . . 100
4.3 Monitoring crystallization condition for IS 18 sample . . . . . . . . . . . . . 102
4.4 EDS result for the comparison of NaCl particles after the crystalliza-
tion to the powder produced by evaporation of solution . . . . . . . . . . . 107
4.5 Microscopic photos taken at different mean residence time for T 05 . . 110
4.6 NaCl particles showing the crystal habit and size (SEM) . . . . . . . . . . . 111
4.7 Agglomerated or composite NaCl crystals known as twin or macle
produced by a CMSMPR crystallization (SEM) . . . . . . . . . . . . . . . . . 111
4.8 NaCl particles produced at an early stage of the crystallization (SEM) 112
4.9 A regular octahedron inscribed in a cube . . . . . . . . . . . . . . . . . . . . . . . 113
4.10 Temperature distribution at the tip of a growing thermal dendrite . . . 114
4.11 Differential and cumulative volume % of CSD at 1.5 for T 05 sample 115
4.12 Differential and cumulative volume % of CSD at 10.5 for T 05 sample116
4.13 A 3D plot of CSDs for T 05 sample . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.14 A 3D plot of CSDs with a view from z-direction for T 05 sample . . . . 118
4.15 Reproducibility of CSDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.16 Effect of feed solution concentration on CSDs. . . . . . . . . . . . . . . . . . . . 122
4.17 Crystallization temperature effect on CSDs . . . . . . . . . . . . . . . . . . . . . 125
4.18 Flow rate effect on CSDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.19 Mixing speed effect on CSDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
4.20 Effect of different mixer on CSDs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

xii
4.21 Effect of crystallization temperature, feed flow rate, and mixing speed
on the mean size of particles (Lmean ) at steady state for T samples . . . 129
4.22 Linear regression statics on the mean size of particles (Lmean ) vs.
mixing speed at steady state for T samples . . . . . . . . . . . . . . . . . . . . . 131
4.23 Yield of NaCl product at 10 for TS samples . . . . . . . . . . . . . . . . . . . 133
4.24 Effect of temperature, flow rate, and mixing speed on the mean size
of particles (Lmean ) at steady state for TS samples . . . . . . . . . . . . . . . 133
4.25 Linear regression statics on the mean size of particles (Lmean ) vs.
mixing speed at steady state for TS samples . . . . . . . . . . . . . . . . . . . . 134
4.26 Yield of NaCl product at 10 for IS samples . . . . . . . . . . . . . . . . . . . . 136
4.27 Linear regression statics on the product yield (g/cm3 ) vs. flow rate
(ml/min) at steady state (10 ) for IS samples . . . . . . . . . . . . . . . . . . . 137
4.28 Effect of temperature, flow rate, and mixing speed on the mean size
of particles (Lmean ) at steady state for IS samples . . . . . . . . . . . . . . . . 137
4.29 Linear regression statics on the product yield (g/cm3 ) vs. flow rate
(ml/min) at steady state (10 ) for IS samples . . . . . . . . . . . . . . . . . . . 138
4.30 CSDs in differential volume % as a function of time (hours) . . . . . . . . 147
4.31 Effect of reaction temperature on mean size of particles as a function
of time (hours) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.32 Effect of mixing speed on mean size of particles as a function of time
(hours) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.33 Effect of different mixer on CSDs as a function of time (hours) and
differential volume % . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.34 Effect of different type of mixer on mean size of particles as a function
of time (hours) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.1 A plot of population number density distribution . . . . . . . . . . . . . . . . 154
5.2 FEM predictions of number density function (n(v)) for aggregation
and growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
5.3 A typical plot of nonlinear regression data fit to the particle number
density (N ) using the Hounslows constant aggregation rate equation
at steady state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
5.4 The order of importance in main and interaction effects on growth
rate (G) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
5.5 The order of importance in main and interaction effects on nucleation
rate (B0 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.6 The order of importance in main and interaction effects on aggregation
rate (Ka ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

xiii
5.7 Influence of crystallization temperature and mixing speed only (main
effect) on dependent variables at steady state for T samples . . . . . . . . 175
5.8 Nucleation rate (B0 ) vs. number of nuclei (n0 ) at steady state for T
samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
5.9 Nucleation rate (B0 ) vs. particle mean size (Lmean ) at steady state
for T samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
5.10 Influence of feed flow rate and mixing speed only (main effect) on
number of nuclei(n0 ), growth rate (G), and nucleation rate (B0 ), as
well as aggregation rate (Ka ) at steady state for TS samples . . . . . . . . 179
5.11 Nucleation rate (B0 ) vs. number of nuclei(n0 ) at steady state for TS
samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
5.12 Nucleation rate (B0 ) vs. reciprocal of number of nuclei (n0 ) at steady
state for T samples in . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.13 Particle mean size (Lmean ) vs. nucleation rate (B0 ) at steady state
for TS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
5.14 Predicted B0 vs. measured B0 at steady state for TS samples . . . . . . . 185
5.15 Predicted G vs. measured G at steady state for TS samples . . . . . . . . 186
5.16 Predicted Ka vs. measured Ka at steady state for TS samples . . . . . . 188
5.17 Number of nuclei(n0 ) vs. nucleation rate (B0 ) at steady state for IS
samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.18 Nucleation rate (B0 ) vs. growth rate (G) at steady state for IS samples190
5.19 Influence of mixing speed only (main effect) on number of nuclei(n0 ),
growth rate (G), and nucleation rate (B0 ), as well as aggregation rate
(Ka ) at steady state for IS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
5.20 Particle mean size (Lmean ) vs. nucleation rate (B0 ) at steady state
for IS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
5.21 Influence of different type of mixer on population number density
distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
5.22 Measured and predicted volume fraction of particle size distribution
for T 16 sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
5.23 Measured and predicted volume fraction of particle size distribution
for TS 16 sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
5.24 Measured and predicted volume fraction of particle size distribution
for IS 16 sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
A.1 Geometry of a laboratory scale crystallizer (1.4 ` working volume)
with baffles and a turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
A.2 Dimension and configuration of a 1.4 ` crystallizer (sideview) . . . . . . 214

xiv
A.3 Dimension and configuration of a 1.4 ` crystallizer (topview) . . . . . . 215
A.4 Dimension and configuration of a 6-bladed flat disc turbine (Rushton
type) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
A.5 Dimension and configuration of a 6-bladed pitch blade impeller . . . . . 218
B.1 Concentration profile measured with conductivity, H+ concentration,
and temperature. Operating condition: flow rate = 110 ml/min,
mixing speed = 0 rpm, and reactor volume = 1,300 ml for a tank
without baffles and with a turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
B.2 Concentration profile measured with conductivity, H+ concentration,
and temperature. Operating condition: flow rate = 110 ml/min,
mixing speed = 287 rpm, and reactor volume = 1,300 ml for a tank
without baffles and with a turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
B.3 RTD measured with H+ concentration. Operating condition: flow
rate = 110 ml/min, mixing speed = 287 rpm, and reactor volume =
1,300 ml for a tank without baffles and with a turbine . . . . . . . . . . . . . 229
B.4 Comparison of tm vs. mixer rpm at different flow rate for 1,300 ml
laboratory reactor without baffles and with a turbine . . . . . . . . . . . . . 229
B.5 Comparison of tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor with baffles and a turbine . . . . . . . . . . . . . . . . . . . . 230
B.6 Comparison of tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor with baffles and an impeller . . . . . . . . . . . . . . . . . . 230
B.7 Comparison of /tm vs. mixer rpm at different flow rates for 1.3 `
laboratory reactor without baffles and with a turbine . . . . . . . . . . . . . 231
B.8 Comparison of /tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor without baffles and an impeller . . . . . . . . . . . . . . . . 231
B.9 Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for 1.3
` laboratory reactor without baffles and with a turbine . . . . . . . . . . . . 232
B.10 Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for 1.4
` laboratory reactor with baffles and a turbine . . . . . . . . . . . . . . . . . . 232
B.11 Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for 1.4
` laboratory reactor with baffles and an impeller . . . . . . . . . . . . . . . . . 233
B.12 Velocity vector profile for turbulent flow in the 1.4 ` laboratory reactor
operating at 40 ml/min feed flow rate and a mixer speed of 80 rpm . . . 234
B.13 Output concentration as a function of time for the Fluent prediction
of the residence time distribution for various mixer speeds (i.e., 0, 20,
40, 80 and 200 rpm) and a feed flow rate of 40 ml/min . . . . . . . . . . . . . 236
B.14 Comparison of experimental result of residence time distribution func-
tion, E(t), experimental measurements, and calculated ideal E(t) for
a perfectly mixed CSTR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
xv
B.15 Comparison of experiment and Fluent simulation result for mean
residence time ( ) vs. mixing speed (rpm) at various flow rate for
a 1.4 ` laboratory reactor with baffles and a turbine . . . . . . . . . . . . . . 238
B.16 Comparison of experiment and Fluent simulation result for standard
deviation (/tm ) of RTD vs. mixing speed (rpm) at various flow rate
for a 1.4 ` laboratory reactor with baffles and a turbine . . . . . . . . . . . 239
C.1 Operating flow and control diagram for a CMSMPR crystallization
system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
D.1 A plot of differential Number % vs. particle size, L (m) . . . . . . . . . . . 255
D.2 A plot of population density distribution vs. particle size, L (m) . . . . 256
D.3 A plot of differential Number % vs. particle size, L (m) . . . . . . . . . . . 257
D.4 3D plot of PSDs for TS 04, TS 05, and TS 06 . . . . . . . . . . . . . . . . . . . 258
D.5 3D plot of PSDs for TS 07, TS 08, and TS 09 . . . . . . . . . . . . . . . . . . . 259
D.6 3D plot of PSDs for TS 10, TS 11, and TS 12 . . . . . . . . . . . . . . . . . . . 260
D.7 3D plot of PSDs for TS 13, TS 14, and TS 15 . . . . . . . . . . . . . . . . . . . 261
D.8 3D plot of PSDs for TS 16, TS 17, and TS 18 . . . . . . . . . . . . . . . . . . . 262
D.9 3D plot of PSDs for TS 19, TS 20, and TS 21 . . . . . . . . . . . . . . . . . . . 263
D.10 3D plot of PSDs for TS 22, TS 23, and TS 24 . . . . . . . . . . . . . . . . . . . 264
D.11 3D plot of PSDs for TS 25, TS 26, and TS 27 . . . . . . . . . . . . . . . . . . . 265
D.12 3D plot of PSDs for IS 04, IS 05, and IS 06 . . . . . . . . . . . . . . . . . . . . . 266
D.13 3D plot of PSDs for IS 07, IS 08, and IS 09 . . . . . . . . . . . . . . . . . . . . . 267
D.14 3D plot of PSDs for IS 10, IS 11, and IS 12 . . . . . . . . . . . . . . . . . . . . . 268
D.15 3D plot of PSDs for IS 13, IS 14, and IS 15 . . . . . . . . . . . . . . . . . . . . . 269
D.16 3D plot of PSDs for IS 16, IS 17, and IS 18 . . . . . . . . . . . . . . . . . . . . . 270
D.17 3D plot of PSDs for IS 19, IS 20, and IS 21 . . . . . . . . . . . . . . . . . . . . . 271
D.18 3D plot of PSDs for IS 22, IS 23, and IS 24 . . . . . . . . . . . . . . . . . . . . . 272
D.19 3D plot of PSDs for IS 25, IS 26, and IS 27 . . . . . . . . . . . . . . . . . . . . . 273
F.1 A schematic showing the pinning of an advancing step by absorbed
immobile impurities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
F.2 Dependence of step velocity on supersaturation in the presence of Fe3+
contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
F.3 Result of PSD measurements without the addition of CdCl2 . . . . . . . . 282
F.4 Result of PSD measurements with the addition of CdCl2 . . . . . . . . . . 283
F.5 SEM image of particles showing the crystal habit and size after crys-
tallization at 30 C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
xvi
LIST OF TABLES

1.1 Moments of the population density distribution . . . . . . . . . . . . . . . . . . 8


1.2 Various binding energies for different molecular additions . . . . . . . . . . 25
1.3 Particle growth rates and dissolution rates . . . . . . . . . . . . . . . . . . . . . . 50
1.4 Shape factors for various geometries . . . . . . . . . . . . . . . . . . . . . . . . . . 55
1.5 Aggregation Rate Kernels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.6 Other kinetic collision frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.7 Particle breakage on a mass basis (1) . . . . . . . . . . . . . . . . . . . . . . . . . . 60
1.8 Particle breakage on a mass basis (2) . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.1 Overall summary of design parameters used in the installation of 1.4
` working volume of a CMSMPR crystallizer . . . . . . . . . . . . . . . . . . . . 67
2.2 Statistical experimental design for CMSMPR crystallization experi-
ment (T samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.3 Statistical experimental design for CMSMPR crystallization experi-
ment (TS samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.4 Statistical experimental design for CMSMPR crystallization experi-
ment (IS samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
2.5 Statistical experimental design for batch breakage and aggregation
experiment (BT and BI samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.1 Output species existing at various temperature . . . . . . . . . . . . . . . . . . 84
4.1 Solution density and specific gravity of liquid . . . . . . . . . . . . . . . . . . . 96
4.2 Comparison of solution concentration by AAS and OLI stream analyzer 98
4.3 Measured values for a CMSMPR crystallization of IS 18 sample . . . . . 103
4.4 Monitored experimental conditions for CMSMPR crystallization ex-
periment (T samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.5 Monitored experimental conditions for CMSMPR crystallization ex-
periment (TS samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.6 Monitored experimental conditions for CMSMPR crystallization ex-
periment (IS samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.7 Comparison of concentration for dried powder and NaCl particles . . . 108
4.8 Comparison of experimental conditions and results for CSD repro-
ducibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
4.9 Yield and mean size of NaCl particles (T samples) . . . . . . . . . . . . . . . 130
4.10 Yield and mean size of NaCl particles (TS samples) . . . . . . . . . . . . . . 132
4.11 Yield and mean size of NaCl particles (IS samples) . . . . . . . . . . . . . . . 135
4.12 Correlation analysis of independent variables to yield and particle
mean size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
4.13 Monitored experimental conditions for batch breakage and aggrega-
tion experiment (BT samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.14 Monitored experimental conditions for batch breakage and aggrega-
tion experiment (BI samples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.15 Mean size of particles measured as a function of time (hour) for BT
samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.16 Mean size of particles measured as a function of time (hour) for BI
samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
5.1 Analytical data from CMSMPR crystallization experiment (T samples)158
5.2 Analytical data from CMSMPR crystallization experiment (TS samples)159
5.3 Analytical data from CMSMPR crystallization experiment (IS samples)160
5.4 Linear regression statistics for T samples (main effects) . . . . . . . . . . . 162
5.5 Multiple linear regression statistics for T samples (two-factor interac-
tions) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
5.6 Multiple linear regression statistics for T samples (three-factor inter-
actions) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
5.7 Correlation analysis of independent variables to dependent variables
for T, TS, and IS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5.8 Factorial design analysis for T samples . . . . . . . . . . . . . . . . . . . . . . . . 171
5.9 Factorial design analysis for TS samples . . . . . . . . . . . . . . . . . . . . . . . 172
5.10 Factorial design analysis for IS samples . . . . . . . . . . . . . . . . . . . . . . . . 173
5.11 Correlation analysis of independent variables to the dependent vari-
ables for TS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
5.12 Correlation analysis of independent variables to the dependent vari-
ables for IS samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
5.13 Inlet and outlet conditions and composition for T, TS, and IS samples 203
5.14 Kinetic model parameter estimation for T, TS, and IS samples . . . . . . 204
A.1 Dimension of vessel (1.4 ` working volume), standard baffle, and
mixer, as well as in and outlet tube . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
xviii
B.1 Experimental conditions and results of RTD measurements without
baffles and with a turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
B.2 Experimental conditions and results of RTD measurements with baf-
fles and a turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
B.3 Experimental conditions and results of RTD measurements with baf-
fles and an impeller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
C.1 Analysis of feed solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
C.2 Equipment list and specification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
F.1 Comparison of particle volume mean diameter statistics (Arithmetic) . 284

xix
ACKNOWLEDGMENTS

I would like to thank to my advisor, Professor Terry A. Ring, with my deepest


gratitude for his leadership, guidance, patience, and confidence. I cant thank him
enough for creating such a wonderful working environment that he let me explore
not only the world of particle science, but also the world you belong to.
The author also gratefully acknowledges the professional contribution from the
project teams and the financial support from the U.S. Department of Energy and
The Office of Industrial Technologies.
CHAPTER 1

FUNDAMENTALS OF CRYSTALLIZATION

The purpose of this chapter is to survey general theories of crystallization in


order to gain a better understanding of what steps must be taken to characterize the
evolution of crystals during crystallization and to control a final crystal size distri-
bution (CSD) and crystal morphology in a controlled continuous mixed-suspension
mixed-product removal (CMSMPR) crystallizer.

1.1 Crystallization
Crystallization is an important separation and purification process used in many
areas such as chemical, pharmaceutical, petrochemical, and biotechnological, as well
as electronic industries. Its wide use is due to the highly purified and attractive
form in which the compounds can be obtained from relatively impure solutions
by means of a single processing step. Crystallization can be performed at high or
low temperatures, and generally requires much less energy for separation of pure
materials than other commonly used methods of purification. While crystallization
may be carried on from vapor or a melt, the most common industrial method is
from a solution.
Crystallization may be defined as a phase change in which a crystalline product
is obtained from a solution. When a substance is transformed from one phase to
at constant pressure and
another, the change in the molar Gibbs free energy, G,
temperature is given by
= (2 1 )
G (1.1)

where is the chemical potential phase 1 (solute) and phase 2 (solid). When
G < 0, the transition from 1 to 2 is a spontaneous process. When G
> 0, it is
2

not thermodynamically possible, on a macroscopic scale. A necessary and sufficient


condition for equilibrium is when G = 0, given by the lines in Figure 1.1; above
< 0, and below the lines G
the lines, G > 0. A supersaturated solution can

be called undercooled if dCeq /dT > 0 (curve B in Figure 1.1), or superheated if


dCeq /dT < 0 (curve C in Figure 1.1). If T0 is the temperature where the solubility

Figure 1.1. Solubility curves for various types of crystallization: Line A, isother-
mal solubility; Line B, positive temperature coefficient of solubility; Line C, negative
temperature coefficient solubility [1].

is equal to the actual concentration, then at a temperature T ,


Z T T0 )
H(T
=
G =
SdT (1.2)
0 T0

where S is the molar entropy and H


is the molar enthalpy change for the phase

transformation. This equation is used for melt crystallization.


The molar Gibbs free energy can also be expressed as
 
a
G = Rg T ln = Rg T ln(S) (1.3)
a0

where Rg is the gas constant, T is the absolute temperature, a is the activity of the
solute, and a0 is the activity of the pure solute in equilibrium with a macroscopic
3

becomes equivalent to
crystal. Assuming the activity coefficients are one, G
Rg T ln(S), where S is the saturation ratio given by
C
S= (1.4)
Ceq
where C is the actual concentration in solution and Ceq is the solubility at the tem-
perature and pressure of the system. For ionic crystal precipitation, the solubility
is given by the solubility product as shown in the following example,
  2
A+2 + 2B AB2 (s), Ksp = A+2 0 B 0

(1.5)

where [A+2 ]0 and [B ]0 are the ionic concentration of A and B at equilibrium,


respectively. The expression used for the saturation ratio, S, for this example
becomes,
2
[A+2 ] [B ]
S= (1.6)
Ksp
where [A+2 ] and [B ] are the actual ionic concentrations of A and B, respectively.
In some complex cases, several salts can precipitate from the solution but only one
will be the least soluble at the specific temperature and pH of the system.
The crystal size distribution (CSD) produced during precipitation is a result of
the relative rates of reaction, nucleation, growth, and agglomeration, as well as the
degree of backmixing in the precipitator. The kinetics of each of these steps will be
discussed .

1.2 Solutions to Population Balance


Particle characteristics (e.g., crystal size, shape, and degree of agglomeration)
are vitally important for both process efficiency and product quality. When crys-
tallization processes involve a large number of crystals of varying sizes, a crystal
size distribution(CSD) is needed to fully characterize the system. The study of
such CSDs is fundamental to understanding the dynamics, structure, and prop-
erties of condensed matter. In crystallization processes, the population balance
equation(PBE), which governs the CSD, is solved together with mass/energy bal-
ances and crystallization kinetics such as nucleation, crystal growth, breakage and
agglomeration.
4

1.2.1 CMSMPR Model


A continuously operated mixed suspension mixed product removal (CMSMPR)
crystallizer is an excellent device to determine the main kinetic parameters, crystal
growth rate (G) and effective rate of nucleation (B0 ), which influence crystal size
distribution (CSD). The CMSMPR model will be briefly introduced and some lim-
iting conditions will be pointed out in order to establish the basis for crystallization
experiment analysis.
A comparison of the operating modes and the methods available to estimate
kinetic parameters from measured CSDs shows that the operation of CMSMPR
crystallizers is most appropriate for this purpose. However, it is necessary to satisfy
certain operating conditions, which impose limitations on the CMSMPR method,
and to apply recommended measuring and analyzing procedures in order to get
accurate results. An essential prerequisite is that the total crystallizer contents
(solution and crystals) are perfectly mixed. Thus there must be no differences
of concentration (C), temperature (T ), supersaturation (C), suspension density
(MT ), and crystal size distribution (CSD) within the suspension volume. Also, only
solids with the same specification as that in the crystallizer must be removed. As
a result of these limitations, a small crystallizer is recommended since it is easier
to obtain good mixing.
The population balance for analysis of crystallizers is demonstrated by its appli-
cation to a continuous crystallizer to predict the product CSD from a continuous,
stirred tank reactor (CSTR) [2]. Such a continuous crystallizer is called a continuous
mixed suspension, mixed product removal (CMSMPR) crystallizer (Figure 1.2) [3].

For a complete description of the CSD in a continuously operated crystallizer


it is necessary to quantify the nucleation and growth processes and to apply the
three conservation laws of mass, energy and crystal population. The importance
of the population balance, in which all particles must be accounted for, has been a
central feature in the pioneering work of Randolph and Larson [2, 4].
5

Figure 1.2. A continuous mixed-suspension, mixed-product removal (CMSMPR)


crystallizer.

The crystal population density, n (number of crystals per unit size per unit
volume of system) is defined by

N
lim =n (1.7)
L0 L

where N is the number of crystals in the size range L per unit volume. The
value of n depends on the value of L at which the interval dL is taken, i.e., n is a
function of L (Figure 1.3 (a)). The number of crystals in the size range L1 to L2 is
thus given by Z L2
N = ndL (1.8)
L1

Application of the population balance is most easily demonstrated with reference


to the case of a continuously operated MSMPR crystallizer, Figure 1.2. Several
assumptions should be made for the formalism.

1. No classification of particles in the crystallizer (well-mixed reactor) and


steady state operation.

2. No crystals in the feed stream.

3. Particles formed only by nucleation and increase in size only through growth
(negligible breakage, attrition, and agglomeration).

4. Crystal growth rate independent of crystal size.


6

Figure 1.3. Population plots characterizing the crystal size distribution, and the
nucleation and the growth kinetics for a CMSMPR crystallizer [3].

5. All crystals of the same shape, and shape factors for particles are not a
function of size (use single characteristic size as the dimension to describe
the entire size range).

Using the same approach as is used to write material and energy balances, a
population balance can also be constructed. Considering an arbitrary size within
the crystallizer, for size range L1 to L2 with population densities n1 and n2 ,
respectively, the mass balance is

Input = Output (1.9)

Particles may enter or leave the size range via either growth or flow. If G (= dL/dt)
is denoted as the growth rate of the characteristic dimension, V is the crystallizer
volume, and Q is the volumetric flow rate, the number balance equation yields,

V n1 G1 + Qi ni L = V n2 G2 + Q
nL (1.10)

where V n1 G1 is the number of particles entering the size range by growth, V n2 G2


is the number of particles leaving the size range by growth, Qi ni L is the number
of particles entering the size range by flow where ni is average in L1 to L2 and
7

L = L2 L1 , and Q
nL is the number of particles leaving the size range by flow.
Upon arrangement,  
n 2 G2 n i G2
V = Qi ni Q
n (1.11)
L
As L approaches zero,
d(Gn)
V= Qi ni Qn (1.12)
dL
the mean retention time, = V /Q, and assuming no particles in feed stream, i.e.,
ni = 0,
d(Gn)
+n=0 (1.13)
dL
Apply McCabes L law, G 6= G(L),
dn n
= (1.14)
dL G
which is an ordinary differential equation that can be separated and integrated.
If the boundary condition of n0 representing zero-sized particles is employed, i.e.,
n(0) = n0 , the result is 

L
n = n0 exp (1.15)
G
Equation 1.15 is the fundamental relationship between crystal size L and population
density n characterizing the CSD. The quantity n0 is the population density of
nuclei (zero-sized crystals). A plot of log n versus L should give a straight line with
an intercept at L = 0 equal to n0 and a slope of 1/G (Figure 1.3 (b)). Therefore,
if the residence time is known, the crystal growth rate, G, can be calculated.
Furthermore, CSD data can be used to give information on the nucleation and
growth kinetics. If the nucleation rate, B0 , is the rate of appearance of near zero-
sized particles, B0 can be expressed as a function of the supersaturation, c.
dN
B0 = = k1 cb (1.16)
dt L0

The crystal growth rate G can be expressed in a similar manner.


dL
G= = k2 cg (1.17)
dt
as  
dN dN dL
=
dt L0 dL dt L0

8

Therefore, the nucleation rate may be expressed in terms of the growth rate by

B0 = n0 G (1.18)

or
B0 = k3 Gi (1.19)

where
b
i= (1.20)
g
Consequently
n0 = k4 Gi1 (1.21)

so a plot of log n0 versus log G should give a straight line of slope i 1 (Figure 1.3
(c)) or a plot of log B0 versus log G should give a line of slope i. Thus the kinetic
order of nucleation, b, may be evaluated if the kinetic order of growth, g, is known.
The moments of the size distribution can be calculated for the exponential
number density distribution (Equation 1.15) obtained with the idealized CMSMPR
model. These are shown in Table 1.1. The nucleation rate B0 based on the

Table 1.1. Moments of the population density distribution.

R the number of crystals


m0 = 0 n(L)dL per unit suspension volume, Ntot = n0 G = B0
Ntot = m0
R total surface area per
m2 = 0 L2 n(L)dL unit suspension volume, aT = 2n0 (G )3
aT = m2
crystal volume per unit
R suspension volume, = 6n0 (G )4
m3 = 0 L3 n(L)dL = m3 ,
suspension density, mT = 6C n0 (G )4
mT = C m3 mT = C
mass median particle
m3 (0, L50 )
= 0.5 size L50 L50 = 3.67G
m3 R L50 3
L n(L)dL
0.5 = 0
m3
9

volumetric hold up is given by Equation 1.22 as a relationship between the crystal


growth rate G and the residence time or the median diameter L50 :
B0 1 3.673 3.674 G
= = = (1.22)
6G3 4 6L350 6L450
or s
G
L50 = 3.67 4
6B0 /
This is only valid for the idealized MSMPR model.

1.3 Nucleation
A supersaturated solution is required for crystallization to occur. A supersat-
urated solution is not at equilibrium. In order to relieve the supersaturation and
move towards equilibrium, the solution crystallizes. Once crystallization starts,
however, the supersaturation can be relived by a combination of nucleation and
crystal growth. It is the relation of the degree of nucleation to crystal growth that
controls the product crystal size and size distribution, and is, therefore, a crucial
aspect of industrial crystallization processes. The fundamentals of nucleation and
crystal growth will be discussed in this and the next section.
During crystallization new particles are born in the size distribution by nucle-
ation processes. The nucleation rate, which appears as a boundary condition at size
L = L 0 in the population balance, generally has a dominating influence on the
particle size distribution. Nucleation is also the least understood of the various
rate processes in precipitation. There are three main categories of nucleation,
i.e., primary homogeneous, primary heterogeneous, and secondary nucleation. The
mechanisms governing the various types of primary and secondary nucleation are
different and result in different rate expressions. Also the relative importance of
each type of nucleation varies with the crystallization conditions.
It is desirable to classify the various mechanisms of nucleation as shown in
Figure 1.4. Primary nucleation occurs in the absence of crystalline surfaces, whereas
secondary nucleation involves the active participation of these surfaces. Homoge-
neous nucleation rarely occurs in practice, however, it forms the basis of several
10

Figure 1.4. Nucleation scheme.

nucleation theories. Heterogeneous nucleation is usually induced by the presence of


dissolved impurities. Secondary nucleation involves the presence of crystals and its
interaction with environment, i.e., crystallizer walls, impellers, and other crystals.

1.3.1 Homogeneous Nucleation


Homogeneous nucleation occurs in the absence of a solid interface at higher
levels of supersaturation. Classical nucleation theory is given by the Gibbs-Kelvin
equations which were developed for vapor phase condensation of spherical droplets.
Classical theories [5, 6, 7] of primary homogeneous nucleation assume that in
supersaturated solution, solute molecules combine to produce clusters, or embryos
by an addition mechanism.

a + a = a2
a2 + a = a3

a3 + a = a4

a2 + a2 = a4
..
.

acm + am = ac (1.23)
11

that continues until a critical size is reached. The rate of nucleus formation by this
mechanism is given by an Arrhenius type of expression as
 
G
B0 = A exp (1.24)
kB T

where A is the preexponential factor and has a theoretical value of 1030 nuclei/cm3 s.
The thermodynamic considerations for homogeneous nucleation were developed
by Gibbs (1928), Volmer (1939), and others. The overall free energy change per
embryo, G[erg/embryo] , of the aggregate is a result of two terms, the free energy
due to the new surface (a positive quantity) and the free energy due to the formation
of new solid (a negative quantity) [8].

V
G = G + A (1.25)
V  
V HT T
= 1 + A (for cooling)
V T0

where

V = volume of the embryo aggregate(= V r3 )

A = area of the embryo aggregate(= A r2 )

V = molar volume of the precipitate

= surface free energy per unit area, and


= molar free energy for the phase transition
G

= kB T ln(S)
G (1.26)

or with cooling,  
HT 1 1
ln(S) =
kB T T0
where HT is the enthalpy of the phase transition (condensation for a gas or
fusion for a melt) occurring at the temperature T below the equilibrium temperature
T0 . kB is Boltzmanns constant. S = a/a0 is the saturation ratio, where a is the
activity of the condensing species and a0 is the activity in equilibrium with the
12

condensate. A generalized particle radius, r = 3V /A, can be used to calculate the


total free energy as
V r3
 
G(r) = kB T ln(S) + A r2 (1.27)
V
where

V = volume conversion factor


A = surface area conversion factor

For a sphere,
4
V =
3
A = 4

and for a cubic nuclei,

V = 8

A = 24

When the supersaturation, S < 1.0, G(r) is always positive and cluster formation
is non-spontaneous. When the supersaturation, S > 1.0, G(r) has a positive
maximum at the critical size, r , with a maximum Gibbs Free energy, Gmax =
G(r ) which is the energy barrier for nucleation as shown in Figure 1.5. Cluster
larger than the critical size will decrease their free energy by further growth, giving
stable nuclei which grow to form macroscopic particles. Below the critical size,
clusters will decrease their free energy by dissolving. The critical size, r , is obtained

by setting dG(r)/dr = 0 giving
2A V
r = (1.28)
3V kB T ln(S)
Nielsen [7] has adopted the terminology that an embryo is subcritical and a
nuclei is supercritical in size. This critical size corresponds to a value of the free
energy at the maximum of
A r2
Gmax = (1.29)
3
In general, the maximum free energy of a critical cluster, Gmax , includes strain
energy, magnetic energy, electrical energy, etc., in addition to the volume free-energy
13

Figure 1.5. Classical nucleation theory, dependence of nuclei size on Gibbs free
energy as a function of saturation ratio, S [1].

and surface free-energy terms. If they are relevant, these contributions should be
added to obtain the maximum free energy of the critical cluster. Figure 1.6 shows
the critical nuclei size as a function of saturation ratio, S. The standard critical size
of nuclei is given by
2A V
rstd = (1.30)
3V kB T
which occurs in the limit of S = e (=2.718).
For a given value of S, all particles with r r will grow and all particles
with r < r will dissolve. This phenomena is referred to as ripening [9]. At high
supersaturation, the critical size, r , approaches the size of an individual molecule,
where the theory is invalid. At such large supersaturation, the rate of nucleation
is limited by the collision of molecules by diffusion. Homogeneous nucleation is
difficult to observe in practice due to the presence of dissolved impurities and
physical features such as crystallizer walls, stirrers, and baffles.

Embryo Concentrations:
14

Figure 1.6. Critical nuclei size (i.e., G = Gmax ) as a function of saturation


ratio, S. For a given value of S, all r r will grow and all r < r will dissolve [1].

Using a Boltzman distribution and these free energy concepts, the equilibrium
number density of embryos of size, r, is given by
   
G(r) NA G(r)
Ne (r) = N0 exp = exp (1.31)
kB T V kB T

where n0 is number of molecules per unit volume of condensing material in the gas
phase, NA is Advogadros number, and kB is Boltzmanns constant. This expression
is only valid over a range of G values where G is greater than zero. This embryo
size distribution function is shown in Figure 1.7. Detailed experiments have been
carried out studying the nucleation of liquid droplets in clean vapor [10, 11]. The
surface tension of the liquid can be measured readily, and the clusters of liquid in
the vapor phase are likely to be spherical. The experimental data on these systems
has verified the validity of the nucleation equations.

Population Balance for Homogeneous nucleation rate [12]:


15

Figure 1.7. Embryo size distribution function for various values of the saturation
ratio, S. Plot generated for a sphere [1].

 
# 1 dN (r)
B0Homo (s) =
cm3 mmin r dt r=rs

2DV 1/3 N0 Nc G(r )


 
= exp
r kB T
! !2
2DN0 Nc V 1/3 2A V
" A
2A V
# " #
3V kB T ln(S)
= exp
3V kB T ln(S) 3kB T
(1.32)

where

D = solution diffusion coefficient


Nc = number concentration in solution
= surface energy for the crystal/solution interface

V = volume per atom or molecule in the crystal


V = volume conversion factors for homogeneous nucleation
16

A = area conversion factors for homogeneous nucleation


kB = Boltzmanns constant

The classical nucleation rate, J = dN (r )/dt, was divided by a critical nucleus size
of r = 2A V /3V kB T ln(S) to make use of the classical theory of nucleation rate
with population balance equation. It is clear that the nucleation rate increases
with increasing supersaturation and temperature, and decreases with an increase
in surface energy.

1.3.2 Heterogeneous Nucleation


Most primary nucleation is, in practice, likely to be heterogeneous nucleation
induced by other surfaces. Nucleation on a foreign surface, which has a lower
surface energy than that of a new solute particle, takes place at a lower critical
supersaturation. Volmer (1939) found that the decrease in free energy depended
on the contact (or wetting) angle () of the solid phase as

Ghet = Ghom (1.33)


1
= (2 + cos )(1 cos )2 (1.34)
4

where is referred to a shape factor. has a numerical value 1 dependent only


on , i.e., the shape of the nucleus. Thus, a lower supersaturation is necessary for
nucleation in a heterogeneous system. Energy considerations show that spontaneous
nucleation would occur for a system with zero contact angle, however, no such
systems exist in practice. Partial attraction is possible in a case where the foreign
substance and the crystal have almost identical atomic arrangement.
The heterogeneous nucleation rate is similar to the homogeneous nucleation rate
except that the interfacial energy is altered by the cosine of the contact angle, ,
among the crystal, liquid, and foreign particle as given by:
17
 
# 1 dN (r)
B0Hetero (s) =
3
cm mmin r dt r=rs

2DV 1/3 N0 Nc G(r )


 
= exp
r kB T
! !2
2DN0 Nc V 1/3 2A V cos()
" " A
2A V cos()
# #
3V kB T ln(S)
= exp
3V kB T ln(S) 3kB T
(1.35)

The f = geometry factor, which is given in formulations by Zeldovich [13], Schu-


bert [14], Eble [15] and Heyer [16], is not included in this equation.

1.3.3 Secondary Nucleation


Secondary nucleation in crystallization occurs either in the presence of a solute-
particle interface or causes mostly by collisions of crystals with crystallizer and
mixer. These parent crystals have a catalyzing effect on the nucleation phenomena,
and thus, nucleation occurs at a lower supersaturation than needed for spontaneous
nucleation. Recent reviews [17, 18] have classified secondary nucleation into three
categories as apparent, true, and contact nucleations. Apparent nucleation refers
to the small fragments washed from the surface of seeds when they are introduced
into the crystallizer. True secondary nucleation occurs simply due to the presence
of solute particles in solution. Contact nucleation occurs when a growing particle
contacts the walls of the container, the stirrer, the pump impeller, or other particles,
producing new nuclei.
The rate of secondary nucleation is governed by three processes: (1) the gener-
ation of secondary nuclei on or near a solid phase; (2) removal of the clusters; and
(3) growth to form a new solid phase. Several factors influence these processes: the
supersaturation, the rate of cooling, the degree of agitation, and the presence of
impurities.
The degree of supersaturation is the critical parameter controlling the rate of
nucleation. The effect of supersaturation on the nucleation rate is threefold. At
higher supersaturation, the absorbed layer is thicker and results in a large number
18

of nuclei. The size of the critical nucleus decreases with increasing supersaturation.
Thus, the probability of the nuclei surviving to form crystals is higher. As the su-
persaturation is increased, the micro-roughness of the crystal surface also increases,
resulting in a larger nuclei population. In general, the secondary nucleation rates
are enhanced with increasing supersaturation. However, the nucleation exponent
is found to be lower than that for primary nucleation.
Stirring the solution leads to a thinning of the absorbed layer and hence should
lead to lower nucleation rates. However, it was found out that the nucleation rate
increased with the degree of agitation for smaller crystals of magnesium sulfate. The
effect of hardness of the contact material and the crystal hardness also found out
to be important factor on the generation of secondary nuclei; a harder material is
more effective in enhancing the nucleation rates. In situ experiments have indicated
that micro-attrition is also an important source of nuclei. Crystal hardness also
affects nucleation behavior, i.e., a hard, smooth crystal is less effective. Irregular
crystals with some roughness are generally more active. Larger crystals yield more
secondary nuclei. Smaller crystals generally follow the path of the liquid and have
a smaller probability of coming in contact with the stirrer or walls to generate
secondary nuclei.
It is well known that a small amount of impurity can profoundly affect the
nucleation rate, however, it is impossible to predict the effect prior. The presence
of additives can either enhance or inhibit the solubility of a substance. Enhanced
solubilities would lead to lower supersaturations and lower growth rates. If it is
postulated that the impurity adsorbs on the crystal surface, then two opposing
effects come into play the presence of an additive would lower the surface tension
and lead to higher growth rates, however, the impurity adsorption blocks potential
growth sites and lowers nucleation rates. Thus, the effect of impurities is complex
and unpredictable.
Break-up of the crystals has been difficult to predict and it is hard to avoid.
It affects the CSD and the crystal growth rate by reducing the size of the large
crystals and increasing the number of small crystals. Unstable secondary nucleation
19

is in most cases an undesirable phenomenon and it can cause problems in product


quality. Generation of nuclei by secondary nucleation mechanisms makes operating
of a continuous crystallizer more difficult.
The secondary nucleation depends mostly on local CSD and local velocities
which change greatly in various parts of a crystallizer. Suspension density affects
secondary nucleation and crystal growth rate.
A review of contact nucleation, frequently the most significant nucleation mech-
anism, is presented by Garside and Davey [19] who give empirical evidence that
the rate of contact nucleation, B0 , depends on stirrer rotation rate (RPM), length
scale factor, , which defines a reactors size and geometry, particle mass density,
MT , and saturation ratio, S.
 
#
B0Secondary (s) 3
d (S 1)b MTj (RP M )h (1.36)
cm m min

Typical values of b lie between 0.5 and 2.5. These values are much lower than
the typical values for primary nucleation by either homogeneous or heterogeneous
mechanisms where B0 values between 6 12 are more common. The importance
of MT is first order, i.e., j = 1, suggesting that contact of the crystals with the
walls and impeller is the important phenomenon. However, some systems, i.e.,
K2 SO4 [20] and KCl [21], have much lower values of j 0.4. Typical values
of h range from 0 to 8 but most fall between 2 and 4, which are expected from
semi-theoretical models [22].
Gahn and Mersmann [23, 24] have presented a mechanical model for crystal
break-up due to crystal-impeller collisions. Their model describes the effect of the
impact energy and the material properties of the crystals. Impeller rotation speed
and the geometries of the impeller and the draft tube were used to obtain the
collision velocities and the probability of collision.
Secondary nucleation occurs when the impact energy in collision of a crystal
with a crystallizer or other crystals reaches above a critical value. Small fragments
are born and the volumes of the original crystals are reduced. The size of the
smallest fragments, La,min (Equation 1.37), depends only on the material properties
20

of the crystal. The size of the largest fragments, La,max (Equation 1.38), depends
additionally on the impact energy [23]:
32
La,min = (1.37)
3 H 2 Kr
1/3
1 H 2/3 Kr

La,max = Wp4/9 (1.38)
2
where

H = Vickers hardness (N/m2 )


Kr = constant of radial efficiency

= fracture surface energy (J/m2 ), and


= shear modulus (N/m2 )

It has been shown that only fragments with La > La,ef f possess such a high
statistically mean growth rate that they are effective fragments or secondary nuclei.
The ratio Na,ef f /Na,tot of effective fragments to the total number of fragments is
given [23] by
Z Z La,ef f
Na,ef f
= n(La )dLa = 1 n(La )dLa (1.39)
Na,tot La,ef f 0

The number density distribution n(La ) can be approximated by


 2.25
2.25L3.25 2.25L3.25

32
n(La ) = 2.25 a
2.25
a
2.25 2.25 2
L3.25
a (1.40)
La,min La,max La,min 3 H Kr

The effective fragment size, La,ef f , depends on supersaturation because the mean
growth rate of a size interval increases with supersaturation. It also has been shown
that only fragments larger than La,ef f 25m contribute significantly to growth
max . The effective size, La,ef f , is smaller when supersaturation is
 
G(La ) > 0.1G
higher and vice versa.

1.4 Crystal Growth Kinetics


As stable nuclei, i.e., particles larger than the critical size have been formed
in a supersaturated or supercooled system, they begin to grow into crystals of
21

visible size. For a comprehensive historical development of many crystal growth


theories, reference should be made to the critical reviews by Wells [25], Buckley [26],
Strickland-Constable [27], Lewis [28], Chernov [29, 30], and N
y vlt et al. [31]. Al-
though the number of theoretical publications is quit extensive, still there is serious
limitation to our present knowledge of the detailed atomic structure of solutions.
Therefore, any model of atomistic behavior in the neighborhood of a crystal-solution
interface is highly speculative.
For a crystal to grow in a solution, the principle that soluble atoms can be easily
dissolved into a solvate (solvation) must be used. Therefore, desolvation must take
place when dissolved atoms are included into a crystal. The density of atoms varies
greatly in solution due to the low diffusion coefficient of the atoms dissolved in the
solution, so the process by which the dissolved atoms are supplied from the growth
medium cannot be ignored, and the process of diffusion of dissolved atoms in a
solution must be examined.
Elwell and Scheel [12] described two types of surface sites where atoms (or
growth units) can be integrated into the growing crystal structure: a step and a
kink site. As schematically shown in Figure 1.8 (a) step site is a location where
two nearest-neighbor bonds can be made by an atom and a crystal. A kink site is
a location where three nearest-neighbor bonds can be made with the crystal. The
process of crystal growth occurs in following processes as (refer to Figure 1.8):

1. Transport of solute from the bulk solution to the crystal surface through an
interface between the crystals and the growth medium.

2. Adsorption on the crystal surface.

3. Diffusion over the surface.

4. Attachment to a step.

5. Diffusion along a step.

6. Integration into the crystal at a kink site.


22

Figure 1.8. The energetics of crystal growth from solution. (a) movement
of the solvated solute molecule and (b) corresponding energy changes for each
transformation [12].

7. Diffusion of coordination shell of solvent molecules away from crystal surface.

8. Liberation of heat of crystallization and its transport away from crystal.

The solute is often an ion which is solvated by a coordination shell of solvent


atoms (or other ions). At the crystal surface, a desolvation of one, two or three of
the solvent molecules in the coordination shell must occur before the solute can (1)
adsorb, (2) attach to a step or (3) integrate into the crystal of a kink, respectively.
The diffusion of solvent molecules (or other coordination ions) away from the crystal
surface may limit the diffusion of the new solvated solute toward the crystal surface,
23

and thus limit the growth rate. The solute does not become a part of the crystal
structure until the enthalpy of crystallization has been librated and desolvation is
complete. Figure 1.8 schematically presents this process along with the energetics
of each step [12].
The growth rate is determined by the rate at which the crystals pass through
each of these processes. The process that has the largest impact on the overall
rate of growth is clearly the process in which the rate of growth is the slowest.
This particular process is therefore known as the rate-determining process. Exactly
which of the processes becomes the rate-determining process depends on the growth
environments and the growth conditions.
If the growth medium is solution, many of the crystals in the medium will
grow as polyhedrons. In this case the growth interface is flat on an atomic scale.
Therefore, the rate-determining step is the interface kinetic process because the
orientation time for this particular process is sufficiently long compared with both
the atom-supply process and the heat removal process. For the crystal growth
in a solution, the concentration of solute atoms in the growth medium is lower
than that in the crystals, but the density of the solution itself is almost equal to
that of the crystal. Therefore, the volume diffusion coefficient is estimated to be
smaller by two orders of magnitude than that of the mixed gas phase. Because the
density of solute atoms in a solution phase is sufficiently low compared to that in
crystals, solution phase is called thin growth media. However, the density of atoms
in the environment surrounding the crystals is significantly different from that of
the atoms in crystals, thus one phase can be clearly distinguished from the other
at a boundary (surface).
The growth units can either remain at their initial point of contact, diffuse across
the surface, and eventually integrate into the structure of the crystal, or return to
solution. The nature of the crystalsolution interface is of critical importance. A
simple representation of both atomically smooth and rough surface is shown in
Figure 1.9. This figure illustrates two types of atomic structure of crystal surfaces,
i.e., smooth (atomically flat close-packed interface) and rough (atomically diffuse
24

(a) (b)

Figure 1.9. A representation of a growing crystal interfaces. (a) atomically smooth


surface and (b) rough surface.

interface) surfaces. On the smooth surface, all the atoms shown as a cube are
identical. Inside a crystal, any atom will have six neighbors with a binding energy
of three times the atomic bond energy (i.e., 3Eaa ), since each bond is shared
between two atoms; when the binding energy per atom pair is (1/2)Eaa and only
nearest-neighbor interactions are considered. If a single extra atom is to be added
to the smooth crystal interface, it forms a bond with only one nearest neighbor with
the atomic pair bond energy. Single or multiple atoms can be added to form extra
bonds with the first adatom stabilizing the cluster, but the bonding energy of the
first atom to the surface is clearly a major barrier to the crystal growth. Therefore,
in this case the transition from liquid to solid occurs over a rather narrow transition
zone approximately one atom layer thick. On the other hand, any atom added to
a rough surface has a greater probability of becoming part of that surface than a
crystal with a smooth surface since the probabilities of the binding energies will
depend on through terms exp[((1/2)Eaa )/kB T ], exp[(Eaa )/kB T ], and etc., where
T is the interface temperature and kB is Boltzmanns constant. An atom added
to a rough surface has several possible sites of attachment with different binding
energies as shown in Table 1.2. The growth rate of a crystal with a rough surface
will be higher than that of a crystal with a smooth surface due to the transition from
25

Table 1.2. Various bind-


ing energies for different
molecular additions.

Legneda Binding energy


A (1/2)Eaa b
B Eaa
C (3/2)Eaa

a
These sites are equivalent
to the sites in Figure 1.9
b
This is the same as the ad-
dition to an atomically smooth
surface

liquid to solid occurs over several atom layers. Rough surfaces tend to maintain
their roughness during growth; atoms attach themselves at sites which are preferred
sites for subsequent additions. On a smooth surface, the rate-limiting growth step
is the addition of the first atom to the surface, since the subsequent addition of
other atoms to the newly created corners is energetically favored. Consequently,
this new layer is quickly completed, yielding another smooth surface with a higher
activation energy. Therefore the crystal growth on a smooth surface is referred to
as a lateral growth whereas a continuous growth refers to the addition of atoms
to a rough surface. As well defined crystal faces are considered to be planar as
shown in Figure 1.8, new solute molecules arriving on these planes must find a
suitable incorporation site. This sequence then allows for surface attachment with
subsequent surface diffusion to a step and incorporation into a kink-site. Growth
then occurs by a flow of steps across the surface. With a step velocity of , a
step spacing , and step height of a (equivalent to a cubic embryo of height), it is
immediately evident that the growth rate normal to the surface, R, is given simply
as:
a
R= (1.41)

26

Temkin [32] and Jackson [33] have characterized the roughness of a crystal
surface with a surface entropy factor, , defined as
 
2
= (Ess + Ef f Esf ) (1.42)
kB T

where E is the bond energy. Subscripts f and s correspond to the fluid and
solid, respectively. The free energy change due to solidification, Gs , is given
in Figure 1.10 as a function of the fractional occupation of a single layer, x, and
the surface entropy factor, . From this plot it is shown that when < 2, the

Figure 1.10. Dependence of fraction of the surface occupied on the Gibbs free
energy for various degrees of surface roughness [33].

minimum of Gs occurs at x = 0.5, yielding a rough surface. However, when


> 2, the minimum of Gs occurs near x = 0+ or x = 1 , giving a smooth
surface.
27

1.4.1 Surface Energy Theories


In 1878 Gibbs [34] suggested that the growth of a crystal could be considered as
the total free energy of a crystal in equilibrium with its surroundings at constant
temperature and pressure would be a minimum for a given volume. If the volume
free energy per unit volume is assumed to be constant throughout the crystal, then
n
X
ai gi = minimum (1.43)
1

where ai is the area of ith face of a crystal bounded by n faces, and gi is the surface
free energy per unit area of the ith face. Therefore, if a crystal is allowed to growth
in a supersaturated medium, it should develop into a equilibrium shape, i.e., the
development of the various faces should be in such a manner as to ensure that
the whole crystal has a minimum total surface free energy for a given volume. Of
course, a liquid droplet is very different from a crystalline particle; in the former
the constituent atoms or molecules are randomly dispersed, whereas in the latter
they are regularly located in a lattice structure. The surface energy and the rate of
growth of a face, however, should be inversely proportional to the reticular or lattice
density of the respective lattice plane, so that faces having low reticular densities
would grow rapidly and eventually disappear. In other words, high index faces grow
faster than low. So far there is no general acceptance of the surface energy theories
of crystal growth, since there is little quantitative evidence to support them; their
failure to explain the well-known effects of supersaturation and solution movement
on the crystal growth rate.

1.4.2 Adsorption Layer Theories


Gibbs-Volmer theory [35] based on the existence of an adsorbed layer of solute
atoms or molecules on a crystal face is started from thermodynamic reasoning.
When units of the crystallizing substance arrive at the crystal face they are not
immediately integrated into the lattice, but merely lose one degree of freedom and
are free to migrate over the crystal face (surface diffusion). Therefore, there will
be a loosely adsorbed layer of integrating units at the interface, and a dynamic
28

equilibrium is established between this layer and the bulk solution. The adsorption
layer, or third phase, plays an important role in crystal growth and secondary
nucleation. The thickness of the adsorption layer probably dose not exceed 10 nm,
and may even be nearer 1 nm.

1.4.2.1 Surface Nucleation


The origin of kinks are not absolutely necessary in the theoretical development
in any case, but it can be shown from thermodynamic reasoning that along any
step, there must be a concentration of kink-sites other than zero at any finite
temperature. As expected, the number of kinks per length of step also increases
with temperature. The origin of steps is a much more subtle. Steps are not
expected to form spontaneously. For atomically smooth, close-packed interfaces,
the minimum free energy also corresponds to the minimum internal energy, i.e.,
a minimum number of broken solid bonds. Early theories postulated that surface
adsorbed molecules migrating across crystal surface undergo binary, ternary, . . .
collisions. Such clusters as shown in Figure 1.9 (a) that the number of broken
bonds associated with the interface, i.e., the interfacial energy, will be increased
were unstable and soon broke up. In the very rare case where a large number of
molecules were present simultaneously at some site, a stable grouping could form.
This grouping was termed a critical-sized two-dimensional nucleus.
Expressions for the energy requirement of two-dimensional nucleation and the
critical size of a two-dimensional nucleus can be derived in a similar manner to those
for homogeneous three-dimensional nucleation. Therefore, if this is a circular disc
of radius r (cross-sectional area/circumference = 2) and height a corresponding to
the height of one growth unit (i.e., an atom or a molecule), then the overall excess
free energy of surface nucleation is given as
A r2 a
 
s
G = kB T ln(S) + e L r (1.44)
v
where e is the edge energy per unit length surrounding the embryo and Gs is on
a per embryo basis. For a cylindrical embryo, A = and L = 2 (for a cubic
embryo of height, a, and effective radius, r = a/2, A = 1 and L = 4). The first
29

term in Equation 1.44 corresponds to the change in free energy due to the volume
of the embryo. The second term represents the free energy due to the length of the
edge of surrounding the embryo. Therefore, for the case of spontaneous nucleation
(i.e., S > 1), Gs has a maximum value of
e2 L2 v e2 L2 a2
Gsmax = = (1.45)
4A akB T ln(S) 4A kB T ln(S)
occurring at
e L v e L a2
rS = = (1.46)
2A akB T ln(S) 2A kB T ln(S)
The last term in the previous two equations was obtained by noting that v ' a3 .
Therefore, at high temperature and solution supersaturation, low value of rS is
expected. The number of molecules in a critical nucleus, n , can also be written as
aA (rS )2 A (rS )2 e2 L2 a2
n = = = (1.47)
v a2 4A (kB T ln(S))2
The mole fraction of surface embryos, Xr , of size r is given by
Gs (r)
 
Xr = exp (1.48)
kB T
Equation 1.48 can also be written in terms of the surface concentration of embryo
by dividing by the unit area of the crystal. Recall that in three-dimensional
nucleation theory the nucleation rate was on a per volume of the embryo basis,
but for two-dimensional nucleation theory JS is on an embryo per surface basis
(i.e., embryo/cm2 s). From this, the simplified surface nucleation rate [7], JS , on a
unit area per unit time is estimated as
Gsmax
   
D
JS = exp (1.49)
d4 kB T

where d = v 1/3 (molecular mean diameter), D is the molecular diffusion coefficient


in the liquid and k2 = D/d2 = (kB T /h) exp[GD /kB T ] for homogeneous nucle-
ation, and GD is the activation energy for diffusion in the bulk liquid. Therefore,
JS is very sensitive to S. A typical value of this pre-exponential factor D/d4
s
(= Jmax ), is 1021 cm2 s1 (i.e., D = 105 cm2 s1 , and d = 3107 cm). Considering
homogeneous, heterogeneous, and surface nucleation at the same time, there is a
30

critical value of the supersaturation ratio, Sc . These critical saturation ratios are
presented in decreasing order as

Sc,homo > Sc,hetero > Sc,surf ace > 1 (1.50)

This order of critical saturation ratios is indication to the order of the nucleation
activation energy which controls the nucleation rates. A plot of each entity in the
total nucleation rate equation is given in Figure 1.11. It shows that heterogeneous

Figure 1.11. Generalized nucleation rate for homogeneous, heterogeneous, and


surface nucleation [36].

nucleation (A = 2) takes place at a lower critical saturation ratio, with a maximum


nucleation rate of 1025 , which corresponds to the consumption of all the foreign
nuclei for this particular example. At higher levels of supersaturation, homogeneous
nucleation (A = 200) takes place [36].
The critical size of embryos form, grow, and expand across the surface by adding
new molecules. These steps are not straight but as rS increases much beyond rS ,
it may not be a bad approximation to neglect the effect of curvature. Thus, the
step formation reveals a serious fault. Comparing experimental growth data with
31

this surface nucleation, it is easily noticed that to maintain a satisfactory rate of


step generation, JS must be large, i.e., S must be large. But many experimental
data showed that the real growth rates are considerably higher than predicted,
particularly at low values of S. In other words, this spontaneous generation and
growth of two-dimensional nuclei cannot be made sufficiently large to agree with
many experimental data.

Surface nucleation rate [12]:

 
# dN (r)
Js =
cm2 min dt r=rs

Gs (rs )
 
= 4DN0 rs Nc V 1/3 exp
kB T
!
2 L2 V 2/3
! " #
L V 2/3 2A RT ln(S)
= 4DN0 Nc V 1/3 exp
A RT ln(S) kB T
(1.51)

where

rs = critical radius for homogeneous surface nucleation


h i
= edge energy per unit length V 1/3

L = edge length conversion factor

A = surface area conversion factor for surface nucleation

(1.52)

for spherical nuclei, V = 4/3 and A = 4, for cubic nuclei, V = 1 and A = 6,


and for a cylindrical embryo of height, d/2 V 1/3 /2, above the surface, A =
and L = 2; for cubic embryo with an effective radius = L/2, A = 1 and L = 4.
Note that surface nucleation has different units as it occurs at a surface typically
causing scale on that surface and not in the bulk solution.
32

1.4.2.2 Two-dimensional Growth of Surface Nuclei


Several growth models based on two-dimensional nucleation, followed by the
spread of the monolayers have been developed in recent years (OHara and Reid [37],
van der Eerden [38]). The polynuclear growth, i.e., birth and spread (B+S) model
or nuclei on nuclei (NON), can also be seen in literature to describe virtually the
same behavior.
Atoms entering and leaving the surface are in equilibrium; adatoms exist in a
certain density and they are moving about on the surface. Therefore, they can
collide and attach to each other. As these collisions and attachments take place
repeatedly, many adatoms congregate and form a two-dimensional cluster. A small
cluster might lose atoms and disappear, but if a cluster is exposed to fluctuations
and increase in size, the probability that it will survive becomes higher. If the
probability that a certain nucleus could increase in size is equal to the probability
that it could decrease, then such a nucleus is called a critical nucleus. The growth of
such surface nuclei is achieved by either the surface or bulk diffusion of a growth unit
to the step or kink site at the edge of the growing nucleus with little or without
any increase in the number of broken bonds, i.e., the interfacial energy remains
unchanged when atoms arrives at kink sites. For the crystal growth in solution,
bulk diffusion is the rate-determining step yielding a surface growth rate [7] of

dr
= d2 NA Ceq D(S 1) (1.53)
dt

where r is the size of the growing surface nuclei and d is the molecular diameter
of the diffusing species. The two-dimensional growth rate must be compared to
the surface nucleation rate, JS , to have further insight into the rate-determining
step of crystal growth. To make this comparison, the time between two nucleation
events, (JS A)1 , on a given surface area, A, is compared to the time necessary to
grow a two-dimensional layer over the whole surface, [L/(dr/dt)s ], where L is the
maximum length from the surface nuclei to the edge of the crystal.

Mononuclear Two-dimensional Growth:


33

When (JS A)1  [L/(dr/dt)S ], each layer, on an average, is deposited as a


result of one nucleation site, thus the growth rate normal to the interface will be
governed by the surface nucleation rate. This rate of deposition is described by a
monocular two-dimensional crystal growth (see Figure 1.12 (a)), where the change
in crystal size, R, is given by

A R2 D GSmax
 
dR
= dAJS = exp (1.54)
dt m d3 kB T

A being a crystal shape factor given by A/R2 for a cylinder and d is the mean
molecular diameter. This is extraordinary simple concept that the rate limiting

(a) (b)

Figure 1.12. Two-dimensional growth of surface nuclei. (a) mononuclear growth


and (b) polynuclear growth.

step is the formation of a critical size of nuclei. The odd result of this theory is
that (dR/dt) A R2 . Therefore, for crystals comprised of [100] planes, even if
no symmetric initially, should grow eventually into perfect cubes. No experiments
have not carried out yet to show any relation between these variables.

Polynuclear Two-dimensional Growth:

To apply the two-dimensional nucleation rate concept further, the assumption


of infinite lateral spreading velocity should be removed. In fact, if a zero spreading
velocity is assumed, growth occurs simply by the accumulation of critical size of
embryos. With such a base, the polynuclear two-dimensional crystal growth is
developed. When (JS A)1  [L/(dr/dt)S ], the surface nucleation is so fast that
34

each layer of the crystal is the result of the intergrowth from many individual surface
nuclei. This rate of crystal growth is described by a polynuclear crystal growth (see
Figure 1.12 (b)). Again with a nucleation rate per unit area and time the change
in crystal size, R,as

GSmax
   
dR Dd 2/3 2/3
= (NA Ceq ) (S 1) exp (1.55)
dt P 3 kB T

There is no area dependency, but R is a strange function of S; e.g., as S increases


nucleation rate increases but rS decreases and a maximum in R would be expected
at some value of S. The R S dependency is not observed.

Two-dimensional Birth and Spread Growth:

Between the above two extreme cases, the nucleation occurs and the nuclei
can grow and spread at finite velocities [7] as illustrated in Figure 1.13. The

Figure 1.13. Two-dimensional growth of surface nuclei by birth and spread model.

exact rules relating lateral spreading velocities to supersaturation and other system
parameters represent the basic difference between the different birth and spread
models. However, the above two growth mechanisms are of equal importance when
the time between two nucleation events is 60% of the time necessary to complete
the entire surface. That is

L
(JS A)1 = 0.6 (1.56)
dr
dt S

35

This corresponds to the cross-over point between mononuclear and polynuclear


layer growth. At this point of intersection, the crystal growth rate is given by

dR 0.2V DCeq (S 1)
= (1.57)
dt R

which is 0.2 times the diffusion-controlled growth rate, which will be discussed later.
V is the molar volume of the precipitate in Equation 1.57. This result indicates
that when the crystal growth rate is less than 20% of the diffusion-controlled growth
rate, the monocular layer growth rate dominates; above 20% of the rate of diffusion-
controlled growth, polynuclear layer growth dominates.
In general, the final equation shows that R does not depend upon crystal area
but is a complex function of supersaturation and temperature, i.e., R increases
both with S and T . However, no maximum of R with S is generally predicted.
These three models briefly discussed above are not satisfactory in predicting or
correlating crystal growth rates from solution. Some of the unusual dependencies
of rate with crystal area or supersaturation have already received comment. Also,
with reasonable values of the constants and system parameters, it is found that
at low values of S, predicted rates are much less than those generally observed
experimentally.

1.4.2.3 Surface Diffusion and Dislocation


A step in the form of a screw has its origin as a screw dislocation which is
a crystal fault known to exist in most crystals. No new steps then have to be
generated; solute molecules from the bulk adsorb on the flats between steps, migrate
under a surface concentration gradient to the steps, and become incorporated into
the lattice. This is the basis of the Burton-Cabrera-Frank (BCF) model of crystal
growth. In fact, at very low values of the saturation ratio (below that for surface
nucleation), layered growth has been experimentally observed [7] (i.e., SC,layered <
SC,surf ace 1.5). These new layers are produced by a crystal dislocation, which is
a continuous source of step and kink sites. Figure 1.14 shows a crystal face with
such type of dislocation emerging from point A. Molecules are quickly integrated
36

(a) (b)

Figure 1.14. Archimedian growth spiral. (a) schematic presentation of Archime-


dian growth [39] and (b) spiral growth shown with atomic force microscopy images
(5 5 mm2 ) [40].

into the crystal at the monocuclear step AB. At all points on the curve AB, the
step moves such that the angular velocity decreases with the distance away from
the starting point A. This produces a screw or spiral growth pattern on the crystal
surface, as illustrated in Figure 1.14. This type of growth has been shown in a
large variety of crystals growth from vapor [41], from aqueous solution [42], and
from melt [12]. The presence of growth spirals gives evidence for Franks screw
dislocation model [43].
Burton, Cabrera, and Frank [39] and Bennema and Gilmer [44] developed a
theory to predict the crystal growth rate for screw dislocations, where the growth
rate depends on the shape of the growth spiral. For an Archimedian spiral, as
shown in Figure 1.14, the distance between the steps of the spiral, y0 , is given as

y0 = 4rS (1.58)

where rS = e L d2 /[2A kB T ln(S)] is the critical radius for surface nucleation at a


given saturation ratio, S. A more exact treatment [39, 45] suggests that

19e L h2
y0 = 19rS = (1.59)
2A kB T ln S
37

The base angle, , of the growth cone formed by the spiral is given by
 
1 h
= tan (1.60)
y0
where h is the new layer thickness deposited with each turn of the spiral. The rate
of solute deposition is a result of a volume diffusion flux, jv , through the bulk, and
surface diffusion, js , to the step, where integration into the crystal takes place. For
simplicity, suppose that the curvature of the spiral can be neglected and that the
growth of the spiral is slow compared to surface diffusion. Figure 1.15 is a schematic
representation of this surface. At steady state, the two fluxes must be equal, giving

Figure 1.15. Schematic representation of the solute flux to a step site [36].

djs (y)
jv = 0 (1.61)
dy
where the surface flux, js , can also be expressed in terms of a surface diffusion
coefficient, Ds , and a surface concentration, nS .
dnS ds
js = Ds = Ds nse (1.62)
dy dy
where nse is the equilibrium surface concentration and s is the local saturation
ratio at the step surface. The volumetric flux, jv , can be expressed in terms of a
38

volume diffusion coefficient, Dv , a solute surface concentration (CS = nse s /h) and
a bulk solution concentration (C = nse S/h), as follows:
Dv nse (S s ) Dv (C CS )
jv = = (1.63)
h
where again h is the thickness of each layer, is the diffusion boundary layer
thickness and S is the saturation ratio in the bulk solution. Substituting these
two fluxes into the flux balance Equation 1.61, the following differential equation
results:
d2 s
yS2 + s = S (1.64)
dy 2
where ys = (Ds h/Dv )1/2 is the mean distance traveled by the solute molecules on
the surface to arrive at a step, which is always less than or equal to y0 (the average
distance between steps). This second-order ordinary differential equation has the
solution as
cosh (y/ys )
S s (y) = (S) (1.65)
cosh (y0 ys )
for the following boundary conditions:

S s = S at y = y0

s = S at y = 0

where = 1 s /S at y = y0 .
Note that the above solution is only valid in the regime where the distance
between each kink site is much smaller than the distance between the steps (i.e.,
x0  ys , as shown in Figure 1.15. If x0 > ys , it is necessary to introduce an extra
factor into this equation to account for the nonplanar growth fields around the
kinks. For this condition, the flux of molecules to the step may be written as
ds Ds nse S tanh(y0 /ys )
js |y0 = Ds nse = (1.66)
dy y0 ys
The crystal growth rate (in the z-direction), dR/dt, can be calculated from the flux
of steps in the y-direction times the step height, h. The step flux, jstep (i.e., the
flux of one step per unit time), is described by
js |y0
jstep = (1.67)
y0 d
39

where is the crystal density. Making these substitutions, the crystal growth rate
can be shown to be
   
dR Ds nse S y0
= tanh
dt y0 ys ys
  2  
Ds nse S S1
= 2
tanh (1.68)
ys S1 S

where  
y0 9.5e d
S1 = S= (1.69)
ys ys kB T
Due to complications associated with parameter determination, the first term in
the above equation is usually reduced to an experimentally determined constant.
This theory predicts a second-order dependence of supersaturation on the particle
growth rate (i.e., dR/dt S 2 ) when S  S1 (low supersaturation condition), and a
linear dependence (i.e., dR/dt S) when S S1 (high supersaturation condition).
If many screw dislocations emerge on the same dislocation line of the crystal
surface, multiple spirals will result. To account for the cooperative effect of multiple
spirals, a factor, , must be introduced into the growth equation, giving
  2  
dR Ds nse S S1
= 2
tanh (1.70)
dt ys S1 S

The diffusion equationfor a two-dimensional fieldhas been solved by Cher-


nov [46]; his results show that the concentration profile associated with each step
is cylindrical around each kink site (see Figure 1.16). When there is significant
overlap of these diffusion fields, because of step bunching, the flux for the central
steps is reduced, which slows down the crystal growth of the interior steps. This
allows subsequent steps to catch up with one another and stay together. This step
bunch then has its own bunch flux. Step bunching leads to imperfections (e.g.,
vacancies, defects, and inclusions) in the crystal structure. In fact, the screw step
height observed experimentally is between 50 and 150
A, which corresponds to 10
to 50 molecular steps bunched together. Another observation is that step bunching
increases with increasing supersaturation, since the number of step and kink sites
increases accordingly.
40

Figure 1.16. Distribution of solute concentration profile around each kink/step


site for the volume diffusion model. The dotted line is the isosbestic line [46].

1.4.3 Diffusion-reaction Theories


The origin of the diffusion theories dates back to the work of Noyes and Whit-
ney [47] who considered that the deposition of solid on the face of a growing crystal
was essentially a diffusional process. They also assumed that crystallization was
the reverse of dissolution, and that the rates of both processes were governed by
the difference between concentration at the solid surface and in the bulk of the
solution. An equation for crystallization was proposed in the form
dm
= km A(C C ) (1.71)
dt
where

m = mass of solid deposited in timet

A = surface area of the crystal


C = solute concentration in the solution (supersaturated)
C = equilibrium saturation concentration

km = coefficient of mass transfer

On the assumption that there would be a thin stagnant film of liquid adjacent to the
growing crystal face, through which molecules of the solute would have to diffuse,
Nernst (1904) modified Equation 1.71 to the form
dm D
= A(C C ) (1.72)
dt
41

where D is the coefficient of diffusion of the solute and is the length of the diffusion
path.
The thickness of the stagnant film would obviously depend on the relative solid-
liquid velocity, i.e., on the degree of agitation of the system. Film thicknesses up
to 150 m have been measured on stationary crystals in stagnant aqueous solution,
but values rapidly drop to virtually zero in vigorously agitated systems. As this
implies an almost infinite rate of growth in agitated systems, it is obvious that the
concept of film diffusion alone is not sufficient to explain the mechanism of crystal
growth. Furthermore, crystallization is not necessarily the reverse of dissolution.
A substance generally dissolves at a faster rate than it crystallizes under the same
conditions of temperature and concentration. Another important finding was made
by Miers [48], who determined, by refractive index measurements, the solution
concentrations near the faces of crystals of sodium chlorate growing in aqueous
solution; he showed that the solution in contact with a growing crystal face is not
saturated but supersaturated.
In the light of these facts, a considerable modification was made to the diffusion
theory of crystal growth by Berthoud [49] and Valeton [50], who suggested that there
were two steps in the mass deposition (i.e., the diffusion process), whereby solute
molecules are transported from the bulk of the fluid phase to the solid surface,
followed by a first-order reaction when the solute molecules arrange themselves
into the crystal lattice. These two stages, occurring under the influence of different
concentration driving forces, can be represented as

dm
= kd A(C Ci ) (diffusion) (1.73)
dt

and
dm
= kr A(Ci C ) (reaction) (1.74)
dt
where kd is a coefficient of mass transfer by diffusion, kr is a rate constant for the
surface reaction (integration) process, and Ci is the solute concentration in the
solution at the crystal-solution interface. Figure 1.17 shows these two stages with
the various concentration driving forces. It must be clearly understood that this is
42

Figure 1.17. Concentration driving forces in crystallization from solution accord-


ing to the simple diffusion-reaction model.

only diagrammatic: the driving forces will rarely be of equal magnitude, and the
concentration drop across the stagnant film is not necessarily linear. Furthermore,
there appears to be some confusion in recent crystallization literature between this
hypothetical film and the more fundamental boundary layers.
Equations 1.73 and 1.74 are not easy to apply in practice because they involve
interfacial concentrations that are difficult to measure. It is usually more convenient
to eliminate the term Ci by considering overall concentration driving force, C C ,
which is quite easily measured. A general equation for crystallization based on this
overall driving force can be written as

dm
= KG A(C C )g (1.75)
dt

where KG is an overall crystal growth coefficient. The exponent g is usually referred


to as the order of the overall crystal growth process. In crystallization work the
exponent, which is applied to a concentration difference, has no fundamental species
involved in the growth process. If g = 1 and the surface reaction (Equation 1.74)
is also first-order, the interfacial concentration, Ci , may be eliminated from Equa-
tions 1.73 and 1.74 to give
dm A(C C )
= (1.76)
dt 1/kd + 1/kr
43

where
1 1 1
= + (1.77)
KG kd kr
or
kd kr
KG = (1.78)
kd + kr
For the cases of extremely rapid reaction, i.e., large kr , KG kd and the
crystallization process is controlled by the diffusional operation. Similarly, if the
value of kd is large, i.e., if the diffusional resistance is low, KG kr , and the process
is controlled by the surface integration. It is worth pointing out that whatever the
relative magnitude of kd and kr they will always contribute to KG .
The diffusional step (Equation 1.73) is generally considered to be linearly de-
pendent on the concentration driving force, but the validity of the assumption of a
first-order surface reaction (Equation 1.74) is highly questionable. Many inorganic
salts crystallizing from a aqueous solution give an overall growth rate order, g, in
the range 1 to 2. The rate equations, therefore, can be written as
1 dm
RG = = kd (C Ci ) (diffusion) (1.79)
A dt
= kr (Ci C )r (reaction) (1.80)

= KG (C C )g (overall) (1.81)

The reverse process of dissolution may be represented by the overall relationship

RD = KD (C C)d (1.82)

where d is generally, but not necessary, unity. From equation


RG
Ci = C (1.83)
kd
so Equation 1.80 representing the surface integration step, may be written
 r
RG
RG = kr C (1.84)
kd
where C = C C and r 1. If r = 1,
 
kd kr
RG = C (1.85)
kd + kr
44

as in Equation 1.78. However, if r 6= 1, the surface integration step is dependent


on the concentration driving force in non-linear manner. For example, if r = 2,
Equation 1.84 can be solved to give
v( )
  u  2
kd u kd
RG = kd 1 + t 1+ 1 C (1.86)
2kr C 2kr C

However, apart from such simple cases, Equation 1.84 cannot be solved explicitly
for RG and the relationship between the coefficients KG , kd , and kr remains obscure.
Recently, however, Sobczak [51] has proposed an integral method, based on a
linearization of Equation 1.84, which allows reasonable values of kd and kr to be
estimated.

1.4.3.1 Two-component Diffusion with Different Con-


centrations
If two components are needed to complete the crystal structure (e.g., Na+ and
Cl ), then each species has its own diffusive flux. That is, for each ith component
from Ficks first and second laws
i i

Di C Ceq
Ji = (1.87)
R
Therefore, for a precipitation reaction given as

aA+ + bB
Aa Bb (s) (1.88)

the following solubility product, Ksp , results:


 A a  B b
Ksp = Ceq Ceq (1.89)

For the crystal Aa Bb (s) being formed, the flux of A (i.e., JA ) and the flux of B
(i.e., JB ) must be related to the crystal stoichiometry by
A A
 B B

JA JB DA C Ceq DB C Ceq
= = = (1.90)
a b aR bR
A B
However, both Ceq and Ceq are related through the solubility product. For the case
of a 1:1 (a : b) stoichiometrywith similar diffusion coefficients (i.e., DA = DB =
45

A B
D, an analytical solution is obtained by eliminating Ceq and Ceq in Equation 1.90,
giving s
B 2
A B
 A


 
D C + C C C
JA = JB = Ksp + (1.91)
R 2 4

which can be written in terms of the particle growth rate as


! s

B 2
A B
 A
dR VD C + C C C
= V JA = Ksp +
(1.92)
dt R 2 4

A B A B A B
For the case of C  C (or equivalently, C C  Ksp = Ceq Ceq ) this reduces
to !
dR V DC
B
= (1.93)
dt R
which is the limiting single component case when the equilibrium concentration
B B
is much less than the bulk (i.e., C  Ceq ). However, Equation 1.92 accounts
for the varying concentrations of two reactants within a solution. It should be
A B
noted here that the situation of C  C , corresponding to the solution given in
Equation 1.93, relates to high saturation ratios.

1.4.4 Chemical Reaction


A quantitative measure of the degree of diffusion or surface integration control
may be made through the concept of effectiveness factors. The concept of effec-
tiveness factors, well established in reaction engineering, is useful in characterizing
the extent to which the diffusional resistance influences the growth rate, A crystal
growth rate effectiveness factor, c , defined by the expression:

grwoth rate at the interface conditions


C = (1.94)
grwoth rate when the interface exposed to the bulk conditions

is given by [52, 53]:


C = (1 C Da)r (1.95)
46

where r is the order of the surface integration process and Da is the Damkohler
number for crystal growth, which represents the ratio of the pseudo-first-order rate
coefficient at the bulk conditions to the mass transfer coefficient, defined by

(1 w)
Da = kr (C C )r1 (1.96)
kd

where w is the mass fraction of solute in solution. Equation 1.95 is plotted in


Figure 1.18. When Da is large, growth is diffusion controlled and C Da1 .

Figure 1.18. The effectiveness factor for crystal growth.

Conversely when Da is small C 1 and growth is controlled by the integration


step.

1.4.5 Impurity Effects on Growth


Sometimes the growth rate of a solute crystal is significantly affected by rela-
tively small quantities of foreign ions. These impurities invariably reduce growth
rates compared to pure solutions, but the mechanism of this growth retardation is
still not completely clear. It has been suggested that impurity adsorption increases
47

the interfacial free energy of the solute crystal. The theories developed from
either two-dimensional nucleation or surface diffusion bases would predict a growth
retardation since an increase in would increase the size of the critical nucleus
and decrease the nucleation rate J, the step spacing y0 , and the step velocity js .
A more interesting idea assumes that impurities adsorb tightly on the surface (in
a regular array) and hinder the smooth flow of steps [54]. That is, an advancing
step is somehow forced to flow between impurity atoms. The steps can then have a
local, high curvature. This results consequently a decrease in step velocity. In fact,
if two impurity atoms are separated by a distance less than 2c , the curvature is so
large that step movement completely ceases (refer to Figure F.1 in Appendix F).
Comparing growth rates for two solution, with and without the impurity, it
can be shown that in the impurity case, there is a critical supersaturation (i.e., d
shown in Figure F.2 in Appendix F) below which no growth should occur. If there
is an impurity flux Ji to the surface, the rate of growth without impurity has to
exceed about 142c Ji h in order for any finite growth to occur in the impurity case.
The critical supersaturation for an impure solution, d , is:

Jia
d (1.97)
Tb

where a is roughly 1/3 to 1/4 and b, 3/4 to 2/3 and the proportionality constant
is given in terms of the system parameters. The significance of Equation 1.97 is
that growth in an impure solution should not occur if < d . When > d , two
equally good answers are predicted, that is, there appear to be two mechanisms for
growth through the impurity array [37]. If both routes are equally likely (or nearly
so), unstable growth could result.
With respect to the movement of the steps, Chernov [46] suggests that there are
two effects of impurities: (1) reduction of the number of kink sites, if the impurities
are relatively small and mobile, and (2) action as an obstacle for step movement
(i.e., step pinning) if the impurities are large and immobile. Reducing the number
of kink sites will also effectively decrease crystal growth. Chernov estimates that an
impurity concentration of 103 M will drastically reduce the crystal growth rate.
48

Slavnova [55, 56, 57] has observed qualitative confirmation of Chernovs theory.
The effects of poisons on nucleation kinetics and the spiral shape are discussed by
Sears [58].
The effect of additives on the rates of nucleation can be accounted for with
the combination of the Gibbs adsorption equation and the Langmuir adsorption
isotherm as they effect the interfacial energy, , resulting in [59]:

= m kB T ln (1 + KLang aa ) + 0 (1.98)

where

aa = activity of the surface active additive


m = mono-layer coverage

KLang = Langmuir adsorption equilibrium constant

kB = Boltzmanns constant

T = temperature, and

0 = the interfacial energy of the crystal without additive

1.4.6 Summary of Crystal Growth


Several quite different crystal growth models have been presented in previous
sections. In all cases, the desired end result was an expression for the growth rate R
perpendicular to a face in terms of the independent system variables. The variables
found to be of most importance was supersaturation, temperature, crystal size, and
fluid dynamics, as well as impurity type and level.
A power law approximation for the growth rate of each face as a function of
saturation ratio S has been suggested by many authors.
Growth Rate,
dl h m i
G(l, S) = = K f (S) g(l) (1.99)
dt s
Dissolution Rate,

dl h m i
D(l, S) = = K f (S) g(l) (1.100)
dt s
49

which is written in terms of a rate constant K, a function of supersaturation ratio


f (S) and a size-dependent growth rate function g(l) which depend on the growth
mechanism and the actual growth conditions.
Growth rates of different faces often exhibit different dependencies on the su-
persaturation ratio S, and the crystal size l, as shown in Table 1.3. The derivation
of equations is given throughout this chapter, otherwise it can be found with
appropriate references. The supersaturation ratio, S 1, in Table 1.3 can be
replaced by a new supersaturation ratio, S S (R), where S (R) comes from the
Kelvin equation, i.e., Equation 1.28. This new supersaturation ratio accounts for
the effect of ripening during a precipitation reaction, suggesting that the equilibrium
saturation ratio does not asymptotically approach one but a higher value, S , for
small particles. When S S is positive, particle growth occurs and for S S less
than zero dissolution occurs.

Effect of Supersaturation:

For all the models except for those included in the two-dimensional nucleation,
the prediction of the derivative ( ln R/ ln S) is either 1 or 2. For the Chernov
bulk diffusion and the BCF surface diffusion (screw dislocation) models, the slope
of this derivative changes from 2 at low values of S to 1 at high S (For the BCF
bulk diffusion the slope is slightly greater than 2 at low S and decreases to 1 at
high S.). For other models, it remains equal to unity at all values of S.
For the two-dimensional nucleation theories, the slope changes with S and
the key group is 1/T 2 . For any reasonable supersaturation, the two-dimensional
nucleation can be reduce to:
ln R 1
' 2 (1.101)
ln S T S
The predicted variation of ln R with ln S is quite large. For example, for KCl-water
if S ' 0.01, ln R/ ln S ' 3, 300 if = 100 erg/cm2 or 200 if = 25 erg/cm2 .
Variations of this magnitude are not observed.
Most models predict a dependence of R on S as either proportional to S or
S 2 . The exponential dependence of the two-dimensional nucleation theories seems
50

Table 1.3. Particle growth rates and dissolution rates.

Growth Mechanism K f (S) g(l)



GSmax

A D
Monosurface Nucleation [7] exp l2
d3 kB T
1
Mono/poly crossover point [7] 0.2Vb DCeq S1
l
GSmax
 2/3  
Polysurface Nucleation [7] Dd Vb Ceq 2/3
(S 1) exp l
 3k
B T
S1
 2
DS nse S
Screw Dislocation [39] tanh l
yS 2 S1 S
1
Bulk Diffusion [35] Vb DCeq S1
l
1
Chemical Reaction [52] Vb DCeq S1
l
Vb kH Teq 1
Heat Conduction [1] ln(S)
"Hf l
Bulk Diffusion D b  Vb D
2 V Ceq (S 1) +
l 2lksi
and Surface v
u Vb D 2
u ! ! #
Integration Combined [60] Vb D n  o
t + Vb Ceq (S 1)
2lksi lksi

GS L2 e2 d2 9.5e d2 R T Hf
, S1 = , yS = DS deads , ln S = T eq
p
max = dT
4a kB T ln S kB T yS Rg T 2
where D is the diffusion coefficient (sub-s surface),
Vb is the molar volume of the crystal,
Nav is Avogadros number,
Ceq is the equilibrium concentration [mole/volume],
Hf is the heat of fusion,
is the Damkohler number,
kH is the thermal conductivity,
a is the area shape factor for surface nuclei,
L is the length shape factor for the nuclei,
is the ratio of the surface to the bulk supersaturation ratio,
ys is the mean distance traveled by solute molecules at the surface,
ys2 = Ds deads where deads = the time for de-adsorption of the solute
molecule from the surface,
e is the edge energy per unit length, h i
nse is the surface equilibrium concentration gm/cm2 ,
is the crystal density,
Teq is the equilibrium freezing temperature.
51

incorrect. However, it also does not seem possible to delineate a particular theory
if only isothermal rate data as a function of supersaturation are available.

Effect of Crystal Size:

Most mass transfer theories predict that ln R/d = (n/d), i.e., R dn


where n is typically 1/2 to 1. The mononuclear two-dimensional theory predicts
R d3 . Particle size appears to be important only if mass transfer is the
controlling rate step. If there is a low relative velocity between the crystal and
solution, R d1 ; at higher velocities R d1/2 . Unfortunately, there are no data
to support these predictions.

Effect of Fluid Velocity:

Fluid velocities relative to the crystal affect growth rates only when mass trans-
fer controls. At low relative velocities, the rate is independent of velocity. At high
relative velocities, R U 1/2 where U is the relative velocity between the crystal
and the bulk (length/time).

Effect of Temperature:

A discussion of the effect of temperature on growth rate is quite difficult. As


noted earlier, temperature may affect many of the dependent variables in the rate
expression. As the temperature increases, the surface diffusivity Ds should increase,
the mean diffusion distance on the surface can either become larger or smaller, the
equilibrium surface concentration will probably decrease. However the magnitude
of the presumed changes are not known. Thus, the inadequate statement is usually
made that growth rates increase with temperature. Though probably true, the
statement cannot be proved from theory. For the two-dimensional nucleation
models, since 1/T S is a large number, the predicted effect of temperature is very
large.
52

1.4.7 Size-dependent Crystal Growth


The simplifying assumption of size-independent crystal growth rate, utilized
in the crystal population analysis so far, cannot always be justified. For one
reason or another, many crystals exhibit size-dependent growth and this leads to
the generation of a different product CSD from that described previously under
CMSMPR conditions.
In order to accommodate the concept of size-dependent growth, several empirical
relationships between G (dL/dt) and L have been proposed, but the one currently
most widely used is the ASL equation [61].

G(L) = G0 (1 + L)b b<1 (1.102)

where , b and G0 are experimentally determined constants. Equation 1.102 satisfies


a number of essential requirements, e.g., it is a continuous function of L and b
including L = 0 and b = 0, it acknowledges that crystal nuclei (L = 0) grow at
a finite rate, G0 , and all positive moments of the CSD generated by the model
converge. It does, however, suggest that as L increases to infinity so does G, but
although this is unrealistic, it does not present any serious problems in dealing
with crystal distributions in the size ranges normally encountered in industrial
crystallizers.
An interesting critical analysis of the validity of empirical relationships for the
assessment of size-dependent growth has been made by Rojkowski [62] who proposes
alternative relationships which merit consideration if a more rigorous assessment of
the problem is considered necessary.
The population equation for an CMSMPR crystallizer operated at steady state,
with crystal growth rate independent of size (dG/dL) = 0, may be written

dn n
G = (1.103)
dL

For the case of size-dependent growth, this must be modified to

(nG) n
= (1.104)
L
53

which, when combined with equation 1.102 and integrated, equating with 1/G0 ,
gives
1 (1 + L)1b
 
b
n = n0 (1 L) exp (1.105)
1b
For the case b = 0, equation 1.105 reduces to equation 1.15 and the population
density plot is linear (Figure 1.3). For b > 0 the plot begins to exhibit curvature
which becomes very pronounced at values of b > 0.5. Considerable care must be
taken, however, when attempting to interpret curvature in population density plots
based on experimental data since it is impossible to distinguish the independent ef-
fects of growth rate dispersion and size-dependent growth. Furthermore, curvature
at the extremities of a population density plot can indicate a departure from ideal
CMSMPR behavior. These and other aspects of the problem have been discussed
in some detail by Jan
ci
c and Grootscholten [63] and Randolph and Larson [2].

1.5 Crystal Shape


Crystals are solids in which the atoms are arranged in a periodic repeating
pattern that extends in three dimensions. While all crystals are solids, not all solids
are crystals. Materials that have shortrange rather than longrange ordering,
like glass, are noncrystalline solids. A noncrystalline solid is often referred to
as an amorphous solid. Many materials can form solids that are crystalline or
amorphous, depending on the conditions of growth. In addition, some materials
can form crystals of the same composition but with differing arrangements of the
atoms forming different threedimensional structures. Other materials can have
the threedimensional structure but appear different in shape when viewed under
the microscope. To make sense of this, and to understand the nature of crystals
and how they are identified requires some knowledge of crystals and their structure.
The study of crystal structure is called crystallography and is described in a number
of standard references [64, 65].
Many inorganic molecules form ionic crystals. An ionic structure of sodium
chloride is shown in Figure 1.19. Ionic crystals are made up of the individual
ionized atoms that make up the species in their stoichiometric proportion. They
54

Figure 1.19. Unit cell of NaCl crystal structure.

are held in place by electronic forces. The sodium chloride structure is face centered
cubic and the unit cell contains four sodium ions and four chloride ions. Because the
unit cells contain two types of atoms some additional constraints on the structure
exit. For example, a symmetry operation on the crystal must superimpose atoms
of the same type.
The external appearance of the crystals is also of great importance and is
referred to as the crystals habit, crystal shape, or crystal morphology. Crystal
habit refers to the external appearance of the crystal. A quantitative description of
a crystal means knowing the crystal faces present, their relative areas, the lengths of
the axes in the three directions, the angles between the faces, and the shape factor
of the crystal. Shape factors are a convenient mathematical way of describing the
geometry of a crystal. If the size of a crystal is defined in terms of a characterization
dimension, L, two shape factors can be defined: the volume shape factor (V ) and
the area shape factor (A ).
V = V L3 (1.106)

A = A L2 (1.107)

Shape factors for common geometries are given in Table 1.4.


55

Table 1.4. Shape factors for various


geometriesa .

Body V A F = A /V
Sphere 0.524 3.142 6.00
Tetrahedron 0.182 2.309 12.7
Octahedron 0.471 3.464 7.35
Six-sided prism 2.60 11.20 4.31
Cube 1.00 6.00 6.00
Prism 10.00 2.00 5.00
Platelet 0.20 2.80 14.00
Needle 10.00 42.00 4.20

a
Data adapted from N`
y vlt et al., 1985 [31].

The habit of crystals obtained from an industrial crystallization process can have
a major impact on a number of important properties relating to the slurry and the
dry product. Crystal habit will affect the rheological properties of the suspension,
the filtration or centrifugation efficiency, the bulk density of the solid, and the
flow properties of the solid. The control of crystal habit (along with crystal size
distribution) is, therefore, an important part of industrial crystallization processes.
Crystal habit can vary dramatically with the rate of crystal growth and nu-
cleation. Also changes in the solvent used or the presence of an impurity can
profoundly affect the crystal habit. Figure 1.20 demonstrates the effect of urea on
sodium chloride crystals.

1.5.1 Heat Flow and Interface Stability


The crystallization is controlled by the rate at which the latent heat of solidifi-
cation can be conducted away from the solid/liquid interface. Conduction can take
place either through the solid or the liquid depending on the temperature gradients
at the interface. Consider for example solid growing at a velocity with a planar
interface into a supercooled liquid, Figure 1.21 (a). The heat flow away from the
interface through the liquid must balance that from the solid plus the latent heat
56

(a) (b)

Figure 1.20. Sodium chloride crystal grown from pure solution and from a solution
containing 10% urea.

(a) (b) (c)

Figure 1.21. (a) Temperature distribution for solidification when heat is con-
ducted into the liquid. Isotherms (b) for a planar S/L interface, and (c) for a
protrusion [66].

generated at the interface, i.e.

KL TL0 = KS TS0 + Lv (1.108)

where K is the thermal conductivity, T 0 is the temperature gradient (dT /dx), the
subscripts S and L stand for solid and liquid, is the rate of growth of the solid,
and Lv is the latent heat of fusion per unit volume. If a protrusion forms on the
solid the negative temperature gradient in the liquid becomes even more negative.
Therefore heat is removed more effectively from the tip of the protrusion than from
the surrounding regions allowing it to grow preferentially. A solid/liquid interface
advancing into supercooled liquid is thus inherently unstable.
Heat flow into the liquid arise if the liquid is supercooled below crystallization
temperature (Tc ). Such a situation can arise at the beginning of crystallization if
57

nucleation occurs at impurity particles in the bulk of the liquid. Since a certain
supercooling is required before nucleation can occur, the first solid particles will
grow into supercooled liquid and the latent heat of solidification will be conducted
away into the liquid. An originally spherical solid particle will therefore develop
arms in many directions as shown in Figure 1.22. It is found experimentally that

(a) (b) (c)

Figure 1.22. The development of thermal dendrites. (a) a spherical nucleus, (b)
the interface becomes unstable, and (c) primary arms develop in crystallographic
directions [67].

the primary arms are always in certain crystallographic directions, each of which
is the axis of a pyramid whose sides are the most closely packed planes with which
a pyramid can be formed: e.g., < 100 > in cubic materials, and < 1100 > in hcp
materials [68]. Dendritic growth only continues until the heat of fusion removes
the constitutional supercooling. When this happens, freezing is completed by the
filling in of the spaces between the dendrite arms.

1.6 Aggregation of Crystals


Particles can aggregate by either Brownian motion or shear-induced aggrega-
tion. With Brownian aggregation, diffusion of particles by Brownian motion causes
particle collisions. With shear-induced aggregation, fluid movement induced by an
external source (e.g., a stirrer) causes particle collisions.
Aggregation rate,
58
Z
BA DA = n(l) (l, x, s)n(x)dx
0
1
Z l   l
 

3 3 3
l3 x3 , x, s
3 3
n l x3 n(x)dx
2 0 l x3
(1.109)

where the aggregation kernel, or kinetic collision frequency for different character-
istic conditions are given in Tables 1.5 and 1.6. (ai , aj , s) is a measure of the
frequency with which particles of size ai and aj collide, adhere, and form stable
aggregates, s.

Table 1.5. Aggregation Rate Kernels.

Regime Size (ai + aj ) (ai , aj , s)


    
2 kB T 1 1
Brownian Diffusion [69] 0 > ai + aj > (ai + aj ) +
3 ai aj

Table 1.6. Other kinetic collision frequency.

Size Collision Frequency Factor


Turbulent Subrange [70]
(ai + aj ) (ai , aj , s)
 1/2
1.29 (ai + aj )3

Viscous > ai + aj > 6 Equivalent
 to
4
(ai + aj )3
3
Transition 6 ai + aj < 25 2.365/12 1/4 (ai + aj )8/3
Inertial 25 ai + aj < L/2 6.871/3 (ai + aj )7/3
Macro L/2 > ai + aj L 7.09(L)1/3 (ai + aj )2

1.7 Attrition of Crystals


In a crystallizer, particles are kept in suspension by means of a pump or a
stirrer. This results in attrition of crystals owing to contact with either the high
59

speed impeller of the pump or the stirrer. Attrition occurs also as the result of
collisions of crystals with each other and with the wall. When the impact energy
exceeds the crystal strength, the consequence is the fracture of the crystals. Thus
the crystal geometry changes substantially and it seldom grows to its geometric
shape. Ang and Mullin [71] measured the volume shape factor of nickel ammonium
sulphate crystals and found a value of 0.58 0.02. In an industrial crystallizer a
wide variation in shape factor may be expected. This impact may occur with crystal
faces, edges, and corners. Because of the fluctuating turbulence in the liquid phase,
the particles are almost randomly oriented in the vicinity of the impeller. The
contacts with corners are therefore most likely to result in high local stress due
to the small area of contact. Hence repeated contacts would result in rounded-off
crystals of reduced size, provided that there is no competing mechanism of crystal
growth.
The shape factor of the crystals may vary widely depending on the attrition
process and the nature of the substance [72]. The detailed procedure of generating
random samples of volume shape factor from the normal density function is given
by Dey and Sen Gupta [73].
Particle breakage on a mass basis
Z
BB DB = S(l, s)n(l, t) S(l, s)P (l, x, s)n(x, t)dx (1.110)
l

where breakage rate term, S(l, s), is given in Table 1.7 and primary progeny
function, P (x, y) is given in Table 1.8 with different theories and references.
60

Table 1.7. Particle breakage on a mass


basis (1).

Breakage Rate Term S(l, s)


 n
L
Probability Theory [74] k
 L0 n
L
L0
Probability Theory [75, 76] k
L m

1
L1

Table 1.8. Particle breakage on a mass basis (2).

Reference P (x, y)
  1   2
x1 x1
Prasher [77] j + (1 j )
y   y
x
1 exp
y
Broadbent and Callcott [78]
1 exp(1)
 y 
Klimpel and Austin [79] P1 > 0, not normalized
x

Primary Progeny Function: P (x, y) Rx


Cumulative Progeny Function:
R y P (x, y) = 0 P (x, y)dz
Normalization Condition: 0 P (z,
R yy)dz = 1
Mass Conservation Condition: 2 0 zP (z, y)dz = y
CHAPTER 2

EXPERIMENTAL DESIGN

2.1 General Consideration


Crystallization systems are typically designed based on experience and scale-up
rules of thumb for the crystallizer, mixer, internal heat transfer surfaces, and
baffle geometry. After design, the crystallizer is fine-tuned by altering operating
conditions to produce filterable particles with little control of the particle size
distribution. Typically, significant extra effort has been expended on getting a
crystallizer running properly after it is commissioned. Many times the crystallizer
would produce too many fines requiring that the fines be separated and re-dissolved
and returned to the feed for a second pass. This can result in considerable expense
and loss in efficiency.
For the design consideration of this project, the first things to be known are
the basic level of physical properties and thermodynamics for the system of in-
terest. OLI solution chemistry technology is well suited for modeling crystalliza-
tion processes because of its comprehensive, rigorous thermodynamic modeling
framework that allows us to predict the solubility and supersaturation behavior
including complete speciation and activity coefficients under complex, real solution
conditions [80, 81].
Once the basic level of physical properties and thermodynamics are estab-
lished, the next level of crystallization phenomenology becomes practical. To
initiate crystallization analysis one must be able to predict the supersaturation
conditions of solutions at the full range of industrial conditions [82]. The OLI
thermodynamic models and associated databank have been developed to allow the
prediction of solution conditions in multi-component mixtures at supersaturation.
62

This includes accurate prediction of mean-activity coefficients of ions in solution at


the supersaturation concentration. Supersaturation is the primary driving force in
crystallization and its prediction is crucial to further development of crystallization
kinetics models. The OLI technology also provides for the prediction of transport
properties for aqueous and mixed solvent environments, relevant to hydrodynamics,
and heat and mass transfer [83].

2.2 Process Chemistry of Crystallization


The chemical reaction chosen to study in this work is the purification involving
precipitation of KCl and NaCl from the aqueous solution of CaCl2 , KCl, and NaCl
mixture based upon their characteristic nucleation and crystal growth mechanism
with less agglomeration.
The mixture was at about 62 C when it was drummed at the Dow Chemical,
Ludington. The sample was unfiltered, unsettled process liquor from the evaporator
train. The KCl crystals cannot be driven back into solution simply by heating to
70 C because the process liquor is a slurry of sodium chloride in calcium chloride
solution. Therefore, the precipitation of NaCl only will be studied and analyzed
for this system.
Of the solution chemistry for this precipitation reaction, an aqueous species for
the solution of CaCl2 (47.63 wt%), KCl (1.07 wt%) and NaCl (0.67 wt%) at 70 C
can be written as

H2 O + CaClo2 + KClo + HClo


+ CaOH+ + Ca+2 + CaCl+ + Cl + H+ + OH + K+ + Na+ (2.1)

As temperature cools down from 70 to 30 C, the species existing at 30 C can be


written as

H2 O + CaClo2 + KClo + NaCl(s) + HClo


+ CaOH+ + Ca+2 + CaCl+ + Cl + H+ + OH + K+ + Na+ (2.2)

which should result in the precipitation of NaCl particles out from the solution
leaving small amount of Na+ in aqueous solution.
63

2.3 Layout and Design of a CMSMPR Crystallizer


When considering the successful design of any continuous crystallization sys-
tem, the system must have certain well-defined characteristics and the following
requirements should be incorporated [20, 84].

Overall Design Criteria:

1. The vessel must be small enough so that minimum feed is required, but
large enough that sampling does not result in an appreciable disturbance to
the system.

2. The crystallizer should be designed so that both the suspension and mother
liquor are well mixed, i.e., the liquor composition and the crystal size
distribution must be uniform throughout the volume of the crystallizer.
This perfect mixing condition should be accomplished with a minimum
power input so as to minimize particle breakage.

3. Slurry discharge should be accomplished so that the discharge suspension


density, size distribution, and liquid-phase composition are the same as they
are in the crystallizer. There should be no size classification at the discharge
point.

4. Other features are necessary if transient experiments are to be undertaken.


These include effective control and regulation of rate of energy and rate of
mass input or removal.

Technical as well as economic considerations are incorporated with the results


from the preceding design criteria in order to make a decision on the crystallization
system as to the size of the experimental crystallizer and its operating conditions.
Once these main operating guidelines are established, the specific operating condi-
tions for this unit operation can be established.
A typical laboratory scale jacketed crystallizer, with working volume 1.4 ` in size,
equipped with automatic temperature control by circulating a solution of ethylene
glycol (50 vol%) and water (50 vol%) through the outer jacket of the crystallizer
64

vessel. PolyscienceT M digital temperature controller and bath were used to cool
or heat the solution, making it possible to keep the temperature constant within
0.1 C. The production rate will be established once the choice of the mean
residence time, , and solution concentrations, before and after crystallization, are
determined. The choice of the mean residence time, , will be made based on
reasonable run times when steady state is imposed. Also one needs to keep in
mined that is a function of the specific reaction kinetics associated with each
process. Randolph and Larson [20] have shown that longer residence time can
often be detrimental to the resultant particle size distribution.
The crystallized material can be conserved by recycling the product to a redis-
solving storage container maintained at 70 C and recycling it to the crystallizer.
When this is done over long periods, care must be taken to ensure that evaporation
does not take place, that no crystals remain undissolved, and that extraneous
impurities do not build up in the feed. However, product recycle to reconstitute
the feed may often be advantageous or necessary for high value products.
The feed concentration will be measured beforehand for each of these experi-
ments since the resultant CSD will depend on the level of supersaturation at certain
temperature. Also care must be taken so that no crystallization takes place before
the feed enters the crystallizer. This could be checked with measuring temperature
at the end-tip of feed-point inside the crystallizer.

2.4 Reactor Agitation


The power input per unit volume, crystallizer geometry, and layout should
be carefully considered to achieve the vessel to be well mixed so that powders
are suspended uniformly. One more consideration is that the shear should be
low enough to minimize secondary nucleation by particle breakage. An empirical
correction for the additional power needed for homogeneity, proposed by Yamasaki
et al. [85], was incorporated to estimate the minimum power requirement for our
1.4 ` working volume crystallizer vessel with the configuration of baffles and the
different type of mixer.
65

The geometries of the laboratory scale crystallizer with working volume (1.4 `)
are shown in Figure A.1 through A.3 in Appendix A. These figures present the
strategic placement of all the important flow lines and baffles, also the position
of impeller. The feed is injected 4/5 of the way down the crystallizer to aid well
mixing before going out to the outlet at the top surface. The ratio of liquid level
to the crystallizer diameter (L/D) of the vessel in Figure A.2 shows 1.27 which is
within the general range of 1.0 to 1.5 for most industrial precipitator [20, 86].
To reduce the energy input to the system while maintaining uniformity, a
standard baffle design [87] is used. This device induces turbulence at a lower
energy input while decreasing the nucleation zone resulting in narrower nuclei
size distribution. The use of standard baffle consisting of four flat vertical plates,
radially-directed (i.e., normal to the vessel wall), spaced at 90 around the vessel
periphery running the length of the vessels straight side. Standard baffle width
is 1/10 or 1/12 of the vessel diameter (D/10 or D/12). The gaps with the vessel
wall and base are left to allow the flow to clean the baffles. Recommended gaps are
equal to 1/72 of the vessel diameter (D/72) between the baffles and the vessel wall,
and 1/4 to one full baffle width between the bottom of the baffles and the vessel
base. More of the detailed baffling information could be found in Myers et al. [88].
The dimension of standard baffle (1/12 width of the vessel diameter) and vessel is
given in Table A.1 along with the dimension of mixer in Appendix A. Stirring was
achieved by a 6-bladed flat disc turbine (Rushton type) or a 6-bladed pitch blade
impeller.

2.5 Automatic System Control and Data Acquisition


The system designed and constructed requires precise control of equipment thus
acquiring dependable data. This was accomplished by using automatic system
control and data acquisition with the aid of OPTO 22T M . The computer-interfaced
experimental set-up by OPTO 22T M is shown in Figure 2.1. All parameters, such as
mixing speed (rpm), coolant circulator (temperature and flow rate), inlet feed pump
speed (rpm or flow rate), and water bath (temperature), are controllable through
66

Figure 2.1. Computer-interfaced experimental set-up.

the computer-interface by one simple click and modification. Only thing that is not
controllable is the temperature for storage container which is maintained constant
at 70 C with the aid of drum heater (see more details in Figure C.1 for an operating
flow and control diagram and in Table C.2 for equipment list and specification of a
CMSMPR crystallization system in Appendix C).
The feeding and withdrawing of solution and product from the crystallizer,
respectively, are established by peristaltic pumps controlled by OPTO 22T M . Flow
rate was obtained by the careful calibration of the speed of pump and tube size,
therefore ensuring the desired feed and withdrawal rates as well as a constant
residence time. To ensure constant solution level (within 1.5 mm) inside the
reactor, a Small Tank Alphasonic Level TransmitterT M was used to control the
outlet pump speed while maintaining constant feed flow rate.
Also, the OPTO 22 unit system measures the temperature (solution feed and
outlet point, solution at the end-tip of feed, solution in the tank, coolant in and
outlet of the jacked vessel, coolant in circulator, water bath, and source solution),
pH, conductivity, ionic concentration, mixing speed (rpm), flow rate, and solution
67

level inside the tank on-line in this system. All of these measurements were used
to monitor the reaction and the steady state condition inside the crystallizer.
The calibration for all of these systems was done before each run to increase the
measurement accuracy, see equipment calibration procedures in Appendix C.

2.6 Summary of Overall Design Parameters


Since the choice of variables and the reasonable operating range of these vari-
ables are extremely sensitive to obtain successive experimental results it is wise to be
well studied for this system beforehand. A summary of overall design parameters
are listed in Table 2.1. The operating procedure for this purification involving

Table 2.1. Overall summary of design parameters used in the installation


of 1.4 ` working volume of a CMSMPR crystallizer.

Design Parameter Selected Value Basis


Crystallizer 1.4 ` Working Volume Ha /Db Adjustment
Mean Residence Time, Run times and
35 min.
Reaction Kinetics
CaCl2 (47.63 wt%) Dow Chemical
Feed Solutionc , C01 +KCl (1.07 wt%) and
+NaCl (0.67 wt%) OLI Stream Analyzer
Production Rate 0.2 g NaCl/min Feed Solution and
Nucleation Zone
Mixing Design Standard Baffle
Minimization
Agitator Type Turbined Solid Suspension
Minimum Power
Stirrer Speed, w 400 rpm
for Solids Suspension
Feed Solution Temperature, Solubility and
70 C
TS OLI Stream Analyzer
Reaction pH 7.0 7.5 pH measurement
Crystallization Temperature,
30 C OLI Stream Analyzer
Tc

a
H: Vessel Height.
b
D: Vessel Diameter.
c
The feed solution concentration, C01 = CaCl2 (47.631.76 wt%), KCl (1.070.16
wt%), NaCl (0.670.08 wt%) with 95% confidence level was obtained by AAS.
d
6-bladed flat disc turbine (Rushton type).
68

precipitation of NaCl was established after the completion of trials, see Appendix C
for more details of experimental setup and operating procedures.

2.7 Statistical Experimental Design


The relative importance of variables affecting a chemical process, as well as the
importance of their interactions, can be found by planning and expediting research
experiments according to factorial-design principles. Since the purpose of research
is to obtain information, research efficiency might be defined as the amount of
useful information obtained per unit cost. One technique for increasing this effi-
ciency is that of the statistically designed experiment. This systematic, economical
technique speeds up the solution of research projects by permitting evaluations to
be made before completing all experiments. The method also indicates the relative
importance of process variables and interactions, something not ordinarily possible
with other techniques.

2.7.1 Statistics in Research


Statistics come into the research picture in the design of experiments and the
analysis of data. Design is concerned with how experiments are planned, and
analysis with the method of extracting all relevant information from the data that
has been collected.
Of these two applications, design is undoubtedly of greater importance. The
damage caused by poor design is irreparable because, no matter how ingenious the
analysis, little information can be salvaged from poorly planned experiments. On
the other hand, if the design is sound, then even quick methods of analysis can
yield a great deal of pertinent information.
In this section, it will be emphasized the statistical three-level factorial design
method, but also mention the fractional factorial design procedure because these
two systems are very powerful tools in any kind of research, and have been found
to be of special value in industry. These methods are applicable to any field, such
as chemical, petroleum, food, biology, engineering, business, economics, etc.
69

2.7.2 Statistical Design of Experiments


A logical experimental program ideally suited for practical study of any physical
system or situation is factorial design. Here, experimental conditions are chosen by
selection of a fixed number of levels for each variable, after which experiments are
run at all possible combinations. In our experiments we decide to study first the
effect on yield of the variables of crystallization temperature (TC ), flow rate (Q),
and mixing speed (MS ) assuming all other variables are negligible and measure
the effect of the chosen ones on yield with experimental condition and parameter
stated in Tables 2.2. The design in Table 2.2 contains three processing variables
which affect the solubility, structure, and composition of the precipitated powders.
Feed solution temperature (TS ) and the range of crystallization temperature (TC )
were determined from the use of OLI Lab AnalyzerT M based upon the same solution
concentration as stated in Table C.1 (see Appendix C). However, the concentration
will not be exactly the same for each experiment due to evaporation of H2 O and
HCl so that it is necessary to verify the concentration for each run. If necessary
the concentration also could be adjusted for a certain value.
Essentially, what we want to know from the statistical design is which variables
(or combinations of variables) are important for the process. It can be determined
quantitatively, or perhaps only qualitatively, how the individual variables rank with
respect to their influence on the process. This information may be obtained by
computing average effects of each variable. For instance, one way to proceed to
evaluate the influence of temperature on yield is to observe in Table 2.2 that
for T 01, T 10, and T 19 the conditions of flow rate and mixing speed are the
same, but the temperature conditions are different. Therefore, the difference in the
results of these three combinations of tests can be attributed solely to the effect of
temperature. Similarly, the individual effect of mixing speed and flow rate on yield
can be evaluated. Also, the statistical design allows us to evaluate the combined
effect of two or three variables on yield.
The design in Table 2.3 as compared with Table 2.2 is to examine the change
in crystallization kinetics and particle size distribution due to the difference in feed
70

solution concentration of C02 = CaCl2 (46.88 wt%), KCl (0.93 wt%), NaCl (0.50
wt%) as compared to C01 = CaCl2 (47.63 wt%), KCl (1.07 wt%), NaCl (0.67 wt%).
All other experimental conditions are same as stated in each table. The comparison
of design in Table 2.4 with respect to Table 2.3 is to see the mechanical effect of
different type of mixer (turbine and impeller) on the resultant structure of particles
and CSD. Also from the comparison among Tables 2.2, 2.3, and 2.4 one can easily
have an access to see the every possible effects of three variables (TC , Q, and MS )
in addition to the effect of solution concentrations (C01 , C02 , and C03 ) and the
different type of agitator (turbine and impeller) on the resultant particle size and
shape as a result of different nucleation and crystal growth kinetics.
A total of 81 experiments were performed with some experiments taking as much
as 12 hours to run from the setup to the completion. All the rpms chosen give well
mixed behavior as determined by the residence time distribution measurements, see
Section 4.1. For the impeller, higher rpms than those for the turbine are needed to
give well-mixed behavior.

2.8 Design of Batch Breakage and Aggregation


Experiments
The experimental setup used for the batch experiments was similar to that used
for the continuous experiments except for the inlet and outlet which were shut down
right after the continuous crystallization experiment. The statistical designs shown
in Table 2.5 for the batch breakage and aggregation experiment were designed so
as to see the effect of variables, i.e., reaction temperature (RT ), mixing speed
(MS ), and the type of mixer (TM ), to the resultant shape of particles and CSD
assuming all other variables are negligible and measure the effect of the chosen ones
on yield with experimental conditions and parameters stated in Tables 2.5. For the
comparison to see the effect of different type of mixer, it will be only possible with
mixer speed of 500 rpm since the turbine will make bubbles above 500 rpm and
the impeller will not create enough mixing power for particles to be suspended
in solution below the mixing speed of 500 rpm. Also, the difference in solution
71

concentration will be ignored since the starting CSD for this experiment will be
known from a CMSMPR crystallization which will be performed ahead of each
breakage and aggregation experiment as mentioned earlier.
72

Table 2.2. Statistical experimen-


tal design for CMSMPR crystalliza-
tion experiment (T samples)a .

Experimental Condition
Test No. TC b Qc MS d

( C) (ml/min) (rpm)
T 01 25 20 300
T 02 25 20 400
T 03 25 20 500
T 04 25 40 300
T 05 25 40 400
T 06 25 40 500
T 07 25 60 300
T 08 25 60 400
T 09 25 60 500
T 10 30 20 300
T 11 30 20 400
T 12 30 20 500
T 13 30 40 300
T 14 30 40 400
T 15 30 40 500
T 16 30 60 300
T 17 30 60 400
T 18 30 60 500
T 19 35 20 300
T 20 35 20 400
T 21 35 20 500
T 22 35 40 300
T 23 35 40 400
T 24 35 40 500
T 25 35 60 300
T 26 35 60 400
T 27 35 60 500

a
Constant experimental parameters:
feed solution concentration (C01 ) =
CaCl2 (47.631.76 wt%), KCl (1.070.16
wt%), NaCl (0.670.08 wt%) with 95%
confidence level, feed solution tempera-
ture (TS ) = 70 C, mixing design = stan-
dard baffle + turbine.
b
TC : Crystallization temperature
c
Q: Flow rate
d
MS : Mixing speed
73

Table 2.3. Statistical experimen-


tal design for CMSMPR crystalliza-
tion experiment (TS samples)a .

Experimental Condition
Test No. TC b Qc MS d

( C) (ml/min) (rpm)
TS 01 25 20 300
TS 02 25 20 400
TS 03 25 20 500
TS 04 25 40 300
TS 05 25 40 400
TS 06 25 40 500
TS 07 25 60 300
TS 08 25 60 400
TS 09 25 60 500
TS 10 30 20 300
TS 11 30 20 400
TS 12 30 20 500
TS 13 30 40 300
TS 14 30 40 400
TS 15 30 40 500
TS 16 30 60 300
TS 17 30 60 400
TS 18 30 60 500
TS 19 35 20 300
TS 20 35 20 400
TS 21 35 20 500
TS 22 35 40 300
TS 23 35 40 400
TS 24 35 40 500
TS 25 35 60 300
TS 26 35 60 400
TS 27 35 60 500

a
Constant experimental parameters:
feed solution concentration (C02 ) =
CaCl2 (46.881.73 wt%), KCl (0.930.15
wt%), NaCl (0.500.05 wt%) with 95%
confidence level, feed solution tempera-
ture (TS ) = 70 C, mixing design = stan-
dard baffle + turbine
b
TC : Crystallization temperature
c
Q: Flow rate
d
MS : Mixing speed
74

Table 2.4. Statistical experimen-


tal design for CMSMPR crystalliza-
tion experiment (IS samples)a .

Experimental Condition
Test No. TC b Qc MS d

( C) (ml/min) (rpm)
IS 01 25 20 500
IS 02 25 20 700
IS 03 25 20 900
IS 04 25 40 500
IS 05 25 40 700
IS 06 25 40 900
IS 07 25 60 500
IS 08 25 60 700
IS 09 25 60 900
IS 10 30 20 500
IS 11 30 20 700
IS 12 30 20 900
IS 13 30 40 500
IS 14 30 40 700
IS 15 30 40 900
IS 16 30 60 500
IS 17 30 60 700
IS 18 30 60 900
IS 19 35 20 500
IS 20 35 20 700
IS 21 35 20 900
IS 22 35 40 500
IS 23 35 40 700
IS 24 35 40 900
IS 25 35 60 500
IS 26 35 60 700
IS 27 35 60 900

a
Constant experimental parameters:
feed solution concentration, (C03 ) =
CaCl2 (47.090.725 wt%), KCl
(1.030.04 wt%), NaCl (0.440.034
wt%) with 95% confidence level, feed
solution temperature (TS ) = 70 C,
mixing design = standard baffle +
impeller.
b
TC : Crystallization temperature
c
Q: Flow rate
d
MS : Mixing speed
75

Table 2.5. Statistical experimental design for batch


breakage and aggregation experiment (BTa and BIb
samples).

Experimental Experimental
Condition Condition
Test No. Test No.
RT c MS d RT MS

( C) (rpm) ( C) (rpm)
BT 01 25 300 BI 01 25 500
BT 02 25 400 BI 02 25 700
BT 03 25 500 BI 03 25 900
BT 04 30 300 BI 04 30 500
BT 05 30 400 BI 05 30 700
BT 06 30 500 BI 06 30 900
BT 07 35 300 BI 07 35 500
BT 08 35 400 BI 08 35 700
BT 09 35 500 BI 09 35 900

a
Constant experimental parameters: Starting solution con-
centration (C01 ) = CaCl2 (47.631.76 wt%), KCl (1.070.16
wt%), NaCl (0.670.08 wt%) with 95% confidence level, stan-
dard baffle + turbine configuration.
b
Constant experimental parameters: Starting solution con-
centration (C03 ) = CaCl2 (47.091.76 wt%), KCl (1.030.16
wt%), NaCl (0.440.08 wt%) with 95% confidence level, stan-
dard baffle + impeller configuration.
c
RT : Reactor temperature
d
MS : Mixing speed
CHAPTER 3

EXPERIMENTAL METHODS

3.1 Measurement of Solution Properties


Virtually, all industrial crystallization processes involve solutions. The develop-
ment, design, and control of any of these processes involve knowledge of a number
of properties of the solution.

3.1.1 Density
Density is a significant part of measurement and instrumentation. Density mea-
surements are made for at least two important reasons: (1) for the determination
of mass and volume of products, and (2) the quality of the product. In many
industrial applications, density measurement ascertains the value of the product.
The density of the solution is often needed for mass balance, flow rate, and product
yield calculations.
Pycnometers are static devices. A fixed volume vessel (10 ml) is used and be
filled with the sample liquid. The density of the fluid is measured by weighing the
sample. The simplest version consisting a vessel in the shape of a bottle with a long
stopper containing a capillary hole is shown in Figure 3.1. The capillary is used to
determine the exact volume of the liquid, thus giving high resolution when filling
the pycnometer. The bottle is first weighed empty, and then with distilled-aerated
water to determine the volume of the bottle. The bottle is then filled with the
process fluid and weighed again. The density is determined by dividing the mass
by the volume. The specific gravity of the liquid is found by the ratio of the fluid
mass to water mass.
77

Figure 3.1. A pycnometer with fixed volume (10 ml) is filled with liquid and
weighted accurately; capillary is used to determine the exact volume of the liquid.

When pycnometers are used, for good precision, ultimate care must be exercised
during the measurements; that is, the bottle must be cleaned after each measure-
ment, the temperature must be kept constant, and precision balances must be used.
Commonly accepted sets of conditions are normal temperature and pressure (NTP)
and standard temperature and pressure (STP).
The density measurement for the solution containing the mixture of CaCl2 , KCl,
and NaCl was performed by first taking the solution sample at 70 C which was
heated more than 24 hours and then the solution sample was diluted with distilled
water (1:1 by vol%) since the temperature at which the density measurement of
solution should be made causes particles formation. After the measurement at
20 C with the aid of precision balance, the mass and volume of water added were
subtracted to get the solution density of interest.

3.1.2 Viscosity
The design of any equipment that involves the flow or stirring of liquids re-
quires a knowledge of the fluids viscosity. Since crystallization operations involve
the stirring and movement of suspensions of particles in fluids, the viscosity of
suspensions is important in crystallization design and operation. Viscosity is a
property of particular material defined as the ratio of the shear stress and the shear
78

rate. Viscosity can be thought of as a measure of the resistance of a fluid to flow.


The ratio of the viscosity and the density is another commonly used term that
is known as the kinematic viscosity. The kinematic viscosity has units of length
squared per unit time.
Viscosity increases with increasing concentration in solutions and decreases with
increasing temperature. Recent work [89, 90] has shown that the viscosity of
supersaturated solutions increases with increasing concentration much more rapidly
than in undersaturated solution. This rise in viscosity has been attributed to
the formation of precritical molecular clusters in the solution. The formation of
clusters in solution is a time-dependent process with the cluster size increasing
with increasing time. This would indicate a possible dependence of viscosity on
solution age.
In crystallization operations, the viscosity of the slurry of solution and crystals
is of importance. The viscosity of a slurry of solution and crystals usually does not
obey Newtons law of viscosity, which is called non-Newtonian fluid, but instead
it follows other more complex empirical relations that must be obtained from
experimental data.The viscosity of slurries is a function of the solution and solid
involved, as well as the slurry density. The viscosity can also be significantly affected
by the particle size, size distribution, and particle shape.
Instruments used to measure viscosity are called viscometers. A number of
techniques and configurations are available for viscosity measurement. In rotational
viscometers, some part of the viscometer is rotated imparting movement to the fluid
that is transferred through the fluid to a measuring device. In capillary viscometers,
the fluid flows through a capillary under the force of gravity and the time required
for the fluid to flow through the capillary is measured.
BrookfieldT M DV-II+ programmable viscometer measuring viscous traction on
SC 21 spindle rotating in sample solution is used to measure dynamic viscosity
with varying temperature range of 30 70 C. For the temperature control of
sample, water jacket assembly is added to the viscometer. Viscosity measurement
of S60 solution at 40 C is used for calibration of this equipment and showed the
79

measurement accuracy within the range of 0.45 cP, i.e., 1% error, which confirms
the equipment accuracy. However, the measurement of solution of interest with
varying temperature will precipitate particles at lower temperature thus lowering
measurement accuracy. The solution sample was taken at 70 C from the source
tank the same way as solution sampling for the density measurement except for
dilution.
The SI physical unit of dynamic viscosity is the Pascal-second (Pas), which is
identical to 1 Ns/m2 or 1 kg/ms. The cgs physical unit for dynamic viscosity is
the poise (P). It is more commonly expressed, particularly in ASTM standards, as
centipoise (cP).

1 poise = 100 centipoise = 1 g/cms = 0.1 Pas

3.1.3 Element Analysis of Solution


The element analysis of solution was performed by atomic absorption spec-
troscopy (AAS) to measure the difference in solution concentration of before and
after crystallization thus allowing us to find out the production rate and yield of
NaCl particles, also to define property of solution and solid as well. Since each set
of experiment in Tables 2.2, 2.3, and 2.4 will have different yield of particles, all
the solution samples, before and after crystallization, are analyzed by AAS and the
results were compared with OLI analysis.
Atomic absorption spectroscopy (AAS) uses the absorption of light to measure
the concentration of gas-phase atoms. Since samples are usually liquids or solids, the
analytic atoms or ions must be vaporized in a flame or graphite furnace. The atoms
absorb ultraviolet or visible light and make transitions to higher electronic energy
levels. The analytic concentration is determined from the amount of absorption.
Concentration measurements are usually determined from a working curve after
calibrating the instrument with standards of known concentration. The schematic
of an atomic absorption spectroscopy (AAS) is shown in Figure 3.2.
The light source is usually a hollow-cathode lamp of the element that is being
measured. The disadvantage of these narrow-band light sources is that only one
80

Figure 3.2. Schematic of an atomic absorption spectroscopy (AAS).

element is measurable at a time. AAS requires that the analytic atoms be in the
gas phase. Ions or atoms in a sample must undergo desolvation and vaporization
in a high-temperature source such as a flame or graphite furnace. Flame AAS can
only analyze solutions, while graphite furnace AAS can accept solutions, slurries, or
solid samples. Flame AAS uses a slot type burner to increase the path length, and
therefore to increase the total absorbance. Sample solutions are usually aspirated
with the gas flow into a nebulizing/mixing chamber to form small droplets before
entering the flame. AA spectrometers use monochromators and detectors for uv and
visible light. The main purpose of the monochromator is to isolate the absorption
line from background light due to interferences. Photomultiplier tubes are the most
common detectors for AA spectroscopy.

3.2 Residence Time Distribution Experiment


The continuous stirred tank is used ubiquitously in the chemical process industry
for mixing, reaction, and crystallization. The mixing in a continuous stirred tank
is often not ideal. The residence time distribution (RTD) is one of the ways to
characterize the non-ideal mixing in the tank. Comparison of the measured RTD
with that of an ideal reactor allows the process engineer to diagnose the ills of the
tank and mixer design. The engineer can then use an appropriate mixing model
for the tank in combination with the kinetics of the reaction to be performed in the
tank to develop an appropriate model for the reactor [91, 92].
The RTD is determined experimentally by injecting an inert chemical (mixing
solution of blue dye, NaCl, and HCl with water), called a tracer which is easily
81

detectable, in one shot into the feed stream entering the reactor at some time
t = 0 and then measuring the tracer concentration, C, in the outlet stream as
a function of time. On-line analyzers detected each of the species in the tracer;
blue dye with the aid of a video camera and UV-vis spectrometer, temperature
with a thermocouple, salt with conductivity, and H+ concentration with pH meter
were measured simultaneously with various conditions such as different flow rates
and mixer rpm values. Great care was used in synchronizing the pulse input with
the initiation of data accumulation for RTD analysis and to calibrate the pumps
to ensure that the reactor space-time [= Vtank /Q] was accurately measured. The
reactor control system has the capability of controlling the feed source and reactor
temperature to 0.25 C, the feed and product flow rate to 0.5 ml/min, liquid
level to 3.9 mm and mixing speed to 3 rpm.
For detailed RTD measurement method along with results and simulations refer
to Appendix B. The configuration of reactor with turbine and impeller are shown in
Figure A.2 through A.3. Also for the details of dimension of reactor configuration
and type of mixer see Table A.1 in Appendix A.

3.3 Crystallization Experiment


A feed solution of CaCl2 , KCl, and NaCl mixture which contains large amount
of crystals at room temperature can be solubilized by heating to 70 C. To ensure
complete solubilization, the solution was maintained at 70 C for more than 24 hr.
Then the solution is fed to the crystallizer which is maintained at 70 C. When the
volume of solution reaches 1400 ml, the temperature is kept for another 30 min to
have steady state at 70 C and then the reactor is cooled to 25 35 C with cooling
rate of 0.7 1.0 C/min to induce crystallization. The CaCl2 and KCl are soluble
at around 30 C, and NaCl crystallizes out of solution in a CMSMPR crystallizer
with four baffles and a 6-bladed flat disc turbine (Rushton type) or 6-bladed pitch
blade impeller in the middle of reactor. Also, to see the effect of supersaturation
with varying temperature the crystallization is carried out at 25 and 35 C (see
Tables 2.2, 2.3, and 2.4 for more details in experimental conditions and designs).
82

The detailed operating procedures of crystallization, i.e., the preparation of starting


materials, operating flow and control diagram, and crystallization operation and
sampling preparation for various analyses, etc., are described in stepwise for this
crystallization experiment in Appendix C. A cooling crystallization to purify the
CaCl2 is required of this crystallizer.

3.3.1 Precipitation Point Analysis


The crystallization chosen to study in this work is the purification involving
the precipitation of KCl and NaCl from an impure aqueous solution of CaCl2 by
cooling. From this system, the precipitation of NaCl is studied and analyzed by
heating the solution at 70 C and cooling down to 25, 30, and 35 C, respectively.
Using the OLI Stream Analyzer the various salts are predicted as a function
of temperature as shown in Figure 3.3. These predictions were confirmed by
experiments from the solution and solid analyses with AAS and EDS, respectively.
The detailed information for the ionic complexes existing at various temperatures
is shown in Figure 3.3 (a). The step change in the solubility for Ca at 27 C
is due to the presence of the CaCl+ above this temperature. The precipitation
point for dominant solid species crystallizing out of an aqueous solution is shown
in Figure 3.3 (b) indicating that upon cooling NaCl is the first to be crystallized at
34 C followed by CaCl2 6H2 O at 27 C and KCl at 16 C. The detailed information
for the output species existing on various temperature is shown in Table 3.1. The
reaction scheme based on Table 3.1 was taken as the basis for the crystallization of
sodium chloride such that solution concentrations, crystallization conditions, and
design parameters are determined from it. The chemical conditions necessary for
this particle production were chosen to be

1. Source Temperature, TS = 70 C.

2. Crystallization Temperature, TC = 25 35 C.

3. Reaction pH = 7.0 7.5.


83

(a)

(b)

Figure 3.3. (a) Dominant aqueous species and (b) precipitation point for the
dominant solid species crystallizing out from an aqueous solution (100 g) con-
taining CaCl2 (47.63 wt%), KCl (1.07 wt%) and NaCl (0.67 wt%) as a function
of temperature at 1 atm pressure [Calculated and plotted using Stream Analyzer
(OLI Systems, Inc.)].
84

Table 3.1. Output species existing at vari-


ous temperature. Analysis was done by OLI
Stream Analyzer with a feed solution concen-
tration, (C01 )a .

Temperature CaCl2 6H2 O KCl NaCl



C g g g
20 70.494 0.229 0.565
22 66.823 0.058 0.554
24 61.190 0 0.533
26 50.911 0 0.492
28 22.682 0 0.377
30 0 0 0.263
32 0 0 0.233
34 0 0 0.201
36 0 0 0.168
38 0 0 0.135
40 0 0 0.102
42 0 0 0.068
44 0 0 0.034
46 0 0 0
48 0 0 0
50 0 0 0

a
Feed solution concentration (C01 ) = CaCl2
(47.631.76 wt%), KCl (1.070.16 wt%), NaCl
(0.670.08 wt%) with 95% confidence level, feed so-
lution temperature (TS ) = 70 C, and total solution
mass = 100 g. The calculation was performed at 1
atm pressure.

4. Feed Solution Concentration, C01 = CaCl2 (47.631.76 wt%), KCl


(1.070.16 wt%), NaCl (0.670.08 wt%) with 95% confidence level.

3.3.2 Determination of Reactor Steady State


If an ideal continuous stirred tank reactor is assumed, which is defined as a
reactor in which fluid enters and leaves solely by plug flow, the time to arrive at
steady state is given by Levenspiel [93] as the percentage steady state (SS) stated
as
85

Figure 3.4. Steady state measurements with pH and temperature response.


Crystallization experiment was done at 34.9 C, 40 ml/min, and 380.5 rpm.

  
t
SS = 1 exp 100 (3.1)

For the non-ideality of a particular system, each reactor performance should be
measured. In order to characterize steady state for our system, pH, conductivity,
and ionic concentration as well as temperature are monitored with time. Figure 3.4
shows a result obtained with a crystallization experiment at 34.9 C, 40 ml/min,
and 380.5 rpm. From the figure it is possible to determine a reasonable steady
state condition for the system. The relationship for steady state given in Equa-
tion 3.1 predicts 99.62% of steady state at 5.6 . The time t was set to 0 when the
crystallization temperature reaches 35 C in this case.

3.3.3 Sampling
For continuous processes, the output flow rate must be equal to the input flow
rate such that no mass accumulates in the reactor with time. Allen [94] reviewed
86

many sampling schemes for a variety of sampling problems. By taking advantage of


this continuous processes an overflow line used as a sample stream at some pulsed
frequency. This way the representative sample from the reactor volume without
disturbing the system could be taken. For the general rules of sampling, powder
should be sampled when in motion and the whole of the stream of powder should
be taken for many short increments of time in preference to part of the stream
being taken for the whole of the time to minimize the errors induced by sampling.
Also care must be taken to assure that all representative crystals are removed
from the mother liquor and, once dewatered that agglomeration does not occur.
The slurry volume is needed so that the population density can be determined on a
clear liquor or slurry volume basis. This is done with measuring the filtrate volume
immediately after separation of the crystalline solids. Only the filtrate volume
and weight of dry crystals are needed in order to calculate slurry and population
densities. This result is compared with AAS, ICP, and EDS, as well as XRD to
assure the measurement accuracy.
The sample bottles should be the same temperature as the slurry from the crys-
tallizer. A 0.2 m Supor-200 GelmanSciencesT M filter is used. After the filtration,
samples should be dried at a temperature well below the crystal decomposition
temperature. Special care must be taken with hydrated crystals. For more details
in sampling procedure and handling, see Appendix C.
For the CSD measurement with Beckman Coulter LS230T M , the slurry sample
is diluted with saturated solution, i.e., the filtrate solution itself. Extreme care
should be taken to assure that the temperature must be the same for the sampling,
diluting, and measuring as the slurry in the crystallizer all the time. Details of
CSD measurement procedure are given in Section C.6.3 in Appendix C along with
the sampling procedure and preparation.
The most important concerns in sample treatments in every processes are that
none of the distribution is lost and that no CSD alteration takes place because of
the treatments.
87

3.3.4 Calculation of Yield in a Crystallization Process


In order to calculate the yield in a crystallization process, it is necessary that the
concentration of feed, mother liquor, and any change in solvent inventory (evap-
oration) be known. In most crystallization processes, the supersaturation in the
residual mother liquor is relatively small and can be ignored when calculating the
yield. With some materials, such as sugar, a substantial amount of supersaturation
can exist, and under such circumstances the exact concentration of the solute in the
final mother liquor must be known in order to make a yield calculation. The product
crystal may be hydrated, depending on the compound and temperature at which
the final crystal is separated from the mother liquor. Shown below is a formula
method for calculating the yield of a hydrated crystal from a feed solution [95].
100Wo S(Ho E)
P =R (3.2)
100 S(R 1)
where

P = weight of product,

R = mole weight of hydrate crystal/mole weight of anhydrous crystal,

for anhydrous crystals, R = 1,

S = solubility at the mother liquor (final) temperature in units/100 units

of solvent,
Wo = weight of solute in feed,

Ho = weight of solvent in feed, and

E = evaporation.

3.3.5 Kinetic Data Measurement and Utilization


Brief reference has already been made to the experimental measurement of
crystal nucleation and growth rates in Section 1.2, and a CMSMPR crystallizer
suitable for the purpose is depicted in Figure A.1 (see Appendix A).
To measure data useful for the design or scale-up of industrial crystallizers,
however, it is necessary to operate on a larger scale and at solution densities near
88

the industrial level (> 100 kg/m3 ). Large working volumes, however, can present
problems with the maintenance of good mixing throughout the vessel and the need
to store large volumes of solution at the required feedstock entry conditions. Since
steady-state conditions are not normally achieved in much less than ten residence
times, even the minimum recommended sized vessel would need 30 ` of solution
to be stored to enable meaningful runs to be carried out.
After reaching steady state, which usually obtained after a period of 8 to 15
residence times, product can be used to determine the crystal size distribution and
kinetic parameters, crystal growth rate, and nucleation rate rate. There should
be no crystal size classification at the discharge and sampling points. Every effort
should be made to sample isokinetically, i.e., with the sample withdrawal velocity
equal to that of the solution flowing at the point of removal. For crystallization
vessels operating with solution densities around industrial values it is helpful, for
design and scale-up purposes, if measurements are made of the power input to
the stirrer. Once all the kinetic data measurements are made with precision for
every set of crystallization and batch breakage and aggregation experiments stated
in Tables 2.2, 2.3, 2.4, and 2.5, its utilization and characterization are attempted
with the application of well-known theoretical approach for better understand of
crystallization kinetics.

3.4 Characterization of Particle Properties


3.4.1 Element Analysis of Solid
The element analysis of solid is performed by XRD and energy dispersive x-ray
spectrometer (EDS) in conjunction with the SEM. Combining the EDS system
with the SEM allows the identification, at microstructural level, of compositional
gradients at grain boundaries, second phases, impurities, inclusions, and small
amounts of material. In the scanning mode, the SEM/EDS unit can be used to
produce maps of element location, concentration, and distribution. The standard
guide for the performance of energy dispersive x-ray spectroscopy is covered in
ASTM E 1508, Quantitative Analysis by Energy-Dispersive Spectroscopy.
89

3.4.2 Crystal Size and Shape

3.4.2.1 SEM and Microscopy


During the crystallization experiment, at every the wet crystal size and shape
are monitored and recorded by using a light microscopy with the aid of video cam-
corder. The crystals at 10 are also filtered and stored in a dry container for later
examination of SEM, XRD, and TEM. See operating procedures in Appendix C for
stepwise crystallization procedure and sample preparation.

3.4.2.2 Crystal Size Distribution (CSD)


The definition of size distribution depends entirely on the method used to deter-
mine crystal size distribution (CSD); therefore, the method used for determining the
size must be clearly stated. The CSD is measured with Beckman Coulter LS230T M
which has an experimental reproducibility of 1% and verified for the shape and size
range with optical microscope and scanning electron microscopy (SEM).
The Beckman Coulter LS230T M is a system of multi-functional particle char-
acterization tools. Its laser-based technology permits analysis of particles without
the risk of missing either the largest or the smallest particles in a sample based
on the Fraunhofer and Mie theories of light scattering with 126 optical detectors.
The LS230 uses reverse Fourier lens optics incorporated in a binocular lens system
enabling the LS230 to optimize light scattering across the widest dynamic range in
a single scan with better than 1% reproducibility. The LS230 has a Polarization
Intensity Differential Scattering (PIDS) system that uses three wavelengths of light,
filtered for polarization in the vertical and the horizontal planes. Figure 3.5 shows a
schematic of PIDS measurement system in LS230. Six detectors (in addition to the
126 detectors used for measuring scattered light) are positioned at around 90 degrees
to the direction of the light path to measure the differential intensity between
scattered light of vertical and horizontal polarizations. A total of 42 measurements
are made at six scattering angles and three wavelengths, each at two polarizations.
The combination of multiple wavelengths and two polarizations provides informa-
90

Figure 3.5. Schematic of the PIDS assembly in the Beckman Coulter LS230
particle size analyzer.

tion that differentiates between sub-micron particle sizes and dramatically increases
resolution.
Although PIDS uses a second light source split into flavors, the scattering of these
light beams by particles is described by the same Mie theory as laser scattering,
so all scattering information is converted to particle size using the same algorithm
in a single operation. The performance of the Beckman Coulter LS230 Particle
Size Analyzer relative to traceable standards indicates that the instrument can be
used with confidence, and that the particle size distributions it produces accurately
reflect the sample material.
The theories to calculate particle size based on the Fraunhofer and Mie light
scattering are beyond the scope of this research but the principle is; when a particle
is illuminated by a laser light source, scattering from this beam occurs. The pattern
resulting from light scattering from a suspension of solid particles can be interpreted
to produce a size distribution of the particles present. High suspensions cannot
be handled due to multiple scatter, that is, light scattering off of more than one
particle surface. Multiple scatter is a function of particle number, and thus the
allowable slurry density (above which multiple scatter invalidates the measurement)
91

is a strong function of particle size. This could be avoided by making particle


containing solution diluted with saturated solution with care not to disturb the CSD
in solution. Figure 3.5 shows typical setup with polarization intensity differential
scattering (PIDS) assembly in the Beckman Coulter LS230 particle size analyzer.

3.4.3 Crystal Structure

3.4.3.1 XRD
To find out the crystallographic information of NaCl crystals, X-ray diffraction
(XRD) of dried particles was performed. Below is the brief description of X-ray
diffraction principle.
When X-ray radiation passes through matter, the radiation interacts with the
electrons in the atoms, resulting in scattering of the radiation. If the atoms are
organized in planes (i.e., the matter is crystalline) and the distances between the
atoms are of the same magnitude as the wavelength of the X-rays, constructive
and destructive interference will occur. This results in diffraction where X-rays are
emitted at characteristic angles based on the spaces between the atoms organized
in crystalline structures called planes. Most crystals can have many sets of planes
passed through their atoms. Each set of planes has a specific interplanar distance
and will give rise to a characteristic angle of diffracted X-rays. The relationship
between wavelength, atomic spacing (d) and angle was solved as the Bragg Equa-
tion, see Figure 3.6. If the illuminating wavelengh is known (depends on the type
of X-ray tube used and if a monochromator is employed) and the angle can be
measured (with a diffractometer) then the interplanar distance can be calculated
from the Bragg equation. A set of d-spaces obtained from a single compound will
be represent the set of planes that can be passed through the atoms and can be
used for comparison with sets of d-spaces obtained from standard compounds.

3.5 Measurement of Particle Size Distribution (PSD)


Initial experiments to find an inert liquid in which to disperse the NaCl particles
produced in these crystallization experiments were unsuccessful. Those liquids that
92

Figure 3.6. Braggs law of diffraction.

allow for particle stabilization against aggregation were ones that also dissolved the
particles. Those liquids that were inert (e.g., no dissolution) did not disperse the
particles well enough for PSD analysis.
Alternatively, a slurry sample ( 10 ml) was taken directly from the reactor
to the sample bottle with vacuum pump in a very short time (< 1 s) for a PSD
measurement. The temperature was kept constant during sampling to avoid any
change in the crystals. This sample was diluted in the LS230 to lower the number
density of particles so that light scattering measurement could be made without
multiple scattering. A saturated solution at the temperature of the crystallizer was
used for dilution. It was prepared by filtering the product of the same crystallization
experiment with a 0.2 m Supor-200 GelmanScienceT M filter making sure that
the temperature remained constant. Thus the temperature of the solution and
suspension for sampling, diluting, and measuring of sample was kept constant at
all times at 30 C, the same temperature as the slurry in the crystallizer. This is
feasible for all of the steps except measurement in the LS230 where the sample
is heated by 2 to 5 C by the light source and mixer inside the sample chamber.
Due to this heating the measurement of the PSD was not stable over a 10-minute
period as is shown in Figure F.3 see Section F for more details. The particles tend
to dissolve slowly thus requiring a special technique to make the particles stable
and the PSD measurements successful. This will be discussed later.
93

3.5.1 Stabilization of NaCl Particles for ex situ PSD measurement


The sample preparation for PSD with diluted solution requires special treatment
since the particles can either grow or dissolve during the measurement due to the
different environment. The application of CdCl2 impurity to limit zero-growth rate
and to stabilize particle was discussed with underlying theory. For more detailed
mechanism and application of stabilizing NaCl particles see Appendix F.
CHAPTER 4

EXPERIMENTAL RESULTS AND


DISCUSSION

4.1 Reactor Characterization


4.1.1 RTD measurement as a Reactor Characterization Tool
Residence time distributions (RTD) are measured for both a baffled and an
unbaffled laboratory reactor of the same size with several internal pipes and a
turbine or an impeller operating at different feed flow rates and mixer rpms. Ideal
behavior as determined by the mean and the variance of the RTD was observed at
a mixer Reynolds number of 2,327 for the baffled tank and 3,878 for the unbaffled
tank both in the turbulent transition range.
As an affiliation of this research project team and with their contributions,
the experimental results for the baffled tank are compared to computational fluid
dynamics (CFD) predictions of the RTD using the k turbulence model in
FluentT M for transitional flow regime in the tank, i.e., mixer Reynolds number
between 10 and 10,000. The comparisons are presented in Section B.1.3 (see
Appendix B). All the qualitative aspects of the predicted RTDs are similar to
those measured experimentally. The mean residence times as well as the variances
of the residence time are accurately predicted by CFD in the transition flow regime.
See Appendix B for more details of principle, experiment, and result, as well as
simulation result of RTD by FluentT M .
95

4.2 Solution Properties


The solution which contains large amount of solid at room temperature was
heated to 70 C or above for more than 24 hours to make the solid dissolved
into solution. However, small amount of solid, mostly KCl, cannot be driven
back to solution simply by the heating as explained in Section 2.2. These solid
particles settled down at the bottom of drum and were not considered as part of
solution mixture. The solution in drum was transferred to the preheating solution
container (see Section C.3 in Appendix C for operating flow and diagram) and
heated again at 730.4 C to have a fully homogeneous solution before feed into
the crystallizer. As a result, the solution properties were affected by the change in
solution concentration due to the unavoidable particles setting down at the bottom
of drum. Thus, the characterization of solution properties at 730.4 C was needed
and will be discussed later in this chapter.

4.2.1 Density
The feed solution sample was taken from the solution in preheating solution
container at 730.4 C and the solution after crystallization was a filtrated solution
from the crystallizer outlet. Due to the precipitation of solution at 20 C, the
solution was first diluted with water (1:1 by vol%) and measurements were done at
constant temperature for more than five times to increase accuracy. With a simple
calculation the desired solution density was obtained and given in Table 4.1. Only
one sample was chosen for each set of experimental design from Tables 2.2, 2.3,
and 2.4, simply to see the difference in solution density due to the crystallization of
NaCl. The results clearly showed the loss of solution weight after the crystallization.
This result is verified with other analyses, e.g., AAS, EDS, and XRD. For the
properties of stock solution, i.e., element concentration, pH, and density, as well as,
viscosity, refer to Section C.2 in Appendix C. Although, the density measurement
here was to show the difference in solution before and after crystallization, the
density of feed solution requires special attention since it will be used in Section 5.2
for the interpretation of crystallization process.
96

Table 4.1. Solution density and specific


gravity of liquid measured at 20 C with a
pycnometer, 10 ml volume vessela .

Crystallization
Test No.b
Before After
T 14 1.479 1.474
Densityc
TS 14 1.475 1.470
(g/cm3 )
IS 14 1.476 1.471
T 14 1.480 1.475
Specific Gravity
TS 14 1.476 1.471
of Liquid
IS 14 1.477 1.472

a
As described in Section 3.1.1
b
For detailed experimental conditions of the test
number, see Table 2.2, 2.3, and 2.4
c
The uncertainties of all density measurements
were within 0.003 with confidence level of 95%.

4.2.2 Viscosity measurement


Viscosity of feed solution (C01 ) was measured with the BrookfieldT M DV-II +
programmable viscometer. Although the total allowable error in full scale viscosity
(500 cP) with the SC21 spindle was 5%, the calibration with S60 solution at 40 C
showed the measurement accuracy within the range of 0.45 cP (1%). The solution
sample was taken the same way as the density measurement for the feed solution as
described in previous section. Figure 4.1 shows the viscosity of solution with varying
temperature. It is believed that the difference in the viscosity with temperature
increase and decrease is due to supersaturation limit meaning that supersaturating
a solution some amount will not necessarily result in crystallization, as a result
causing the difference in viscosity between temperature increase and decrease. To
have same viscosity with increase and decrease in temperature, it might be necessary
to measure the viscosity in long period of time which might cause another problem
such as evaporation. This viscosity measurement is to show how the viscosity
will have an effect on the fluid dynamics in the reactor and resultant in CSD
measurement.
97

Figure 4.1. Viscosity of the solution with varying temperature, allowable error =
5%. Analysis of solution concentration (C01 ) is given in Table C.1, see Appendix C.

4.2.3 Element Analysis of Solution


The approximate concentration of the solution, provided by Dow chemical, is
given in Table C.1 for three different solution concentrations. The analysis for
CaCl2 was obtained from a Dow correlation of CaCl2 wt% and density. The solution
density was measured with an Anton Paar DMA 4500 density meter. The other
salts were measured by Perkin ElmerT M atomic absorption spectroscopy (AAS).
Due to aforementioned effects (in Section 2.2), the concentration of solution
analysis was necessary. The solution composition analysis was done with Perkin-
ElmerT M Atomic absorption spectroscopy (AAS) and inductively coupled plasma
atomic emission spectroscopy (ICP-AES). The result in Table 4.2 shows the dif-
ference before and after crystallization with the experimental condition for TS
13 sample stated in Table 2.3, see Section 2.7.2. The experimental result shows
reduction of NaCl only comparable to OLI analysis showing a reasonable prediction
of reduction in NaCl after the crystallization with about 18% difference. The
98

Table 4.2. Comparison of solution concentration by AAS and


OLI stream analyzer for TS 13 samplea,b .

CaCl2 (wt%) KCl (wt%) NaCl (wt%)


AAS OLI AAS OLI AAS OLI
Before Crystallization 46.88 46.88 0.93 0.93 0.50 0.50
After Crystallization 46.97 46.92 0.93 0.93 0.29 0.42

a
The uncertainties in concentration of CaCl2 , KCl, and NaCl were
1.73, 0.15, and 0.05, respectively, with confidence level of 95% for
the AAS analysis.
b
The crystallization condition for TS 13 is given in Table 4.5, see
Section 4.3.3.

discrepancy between experiment and OLI analysis is the indication of difference in


kinetics and thermodynamics between a real system and an ideal model. The OLI
calculation is only based on the thermodynamic model and the solubility database
which were discussed briefly in Appendix G.
With a consideration of analytical errors during sample preparation and analy-
sis, it clearly shows the reduction of NaCl only which means increasing amount
of NaCl particles during the crystallization experiments. An overall mass balance
for each crystallization process can be performed using the information obtained
with these analytical data, i.e., where the difference in liquid input and output
concentrations yields the solid phase concentration. This analysis is verified with
the calculated product yield to assure the solution concentration before and after
crystallization.

4.3 Crystallization Experiment


A study of the crystallization experiment to purify CaCl2 solution which con-
tains KCl and NaCl as impurities was performed by following predesigned experi-
mental plans which addressed 5 processing variables, resulting in 81 experiments.
The experimental designs with each variables were discussed in Section 2.7 and
summarized in Tables 2.2, 2.3, and 2.4 in Section 2.7.2. This section represents
99

the results of that study and discusses the implications of these results on the
crystallization process based on the proof from the analytical experiments and later
in Chapter 5, more details with the theoretical application and model validation
will be discussed.

4.3.1 Precipitation Point Analysis


Using the OLI Stream Analyzer the various salts are predicted as a function
of temperature as shown in Figure 3.3 and Table 3.1. The predictions have been
confirmed by the experimental results from solution analysis with ICP, and AAS as
shown in Table 4.2, and from solid element analysis with EDS shown in Figure 4.4.
The confirmation of solids precipitated demonstrates the accurate prediction of the
solution complexes in a very concentrated CaCl2 solution, which demonstrates the
power of the OLI solution thermodynamics modeling. This accurate prediction
is used to determine the supersaturation ratio in the crystallizing solution that
is the driving force for nucleation and crystal growth. This coupling of solution
conditions including solution concentration and equilibrium concentration with
solids generation is critical for accurate modeling of crystallization.

4.3.2 Crystallization of NaCl from CaCl2 , KCl, and NaCl solution


To begin, the chemical and physical properties of feed solution were monitored
at the start of each experiment to ensure proper handling of experiment and
data analysis. The measured parameters for feed stock solution including pH,
density, and viscosity, as well as elements concentration were chosen to represent
characteristics of the feed solution.
The solution which contains large amount of crystals at room temperature can
be solubilized by heating to 70 C except for some of KCl particles remaining at the
bottom of container. From this system, the CaCl2 is soluble and NaCl crystallizes
out of solution at temperature about 30 C. A cooling crystallization to purify the
CaCl2 is required of this crystallizer. The precipitation point analysis of the system
100

in Figure 3.3 (see Section 3.3.1) showed that as the temperature is cooled from
70 C to 30 C increasing amount of NaCl crystallizes out of solution.
The results of experiments show that indeed there was increasing amount of
crystals as the solution is cooled as shown in Figure 4.2. These particles are

Figure 4.2. NaCl particles after the crystallization at 31.00.18 C (SEM).


Particle sample was taken and filtrated at 10 .

produced by crystallization in a CMSMPR crystallizer (1.4 `) with four baffles and a


6-bladed flat disc turbine (Rushton type) in the middle of reactor operating at 384.9
rpm (T 14 sample, see Table 4.4 in Section 4.3.3). A solution of CaCl2 (47.631.76
wt%), KCl (1.070.16 wt%), and NaCl (0.670.08 wt%) at 72.440.26 C was
feed to the crystallizer and cooled to 31.00.18 C in the crystallizer to induce
crystallization. The uncertainties for all the measurement were obtained with
confidence level of 95%. The feed solution flow is initiated at a given flow rate
equal to the outlet flow to give a certain mean residence time ( ) in the crystallizer
between 20 to 70 min. Then the coolant temperature is reduced to the operating
temperature typically near 30 C. As the crystallizer contents cools crystallization
begins. When the temperature of coolant is set to crystallization temperature, time
is set to zero and on-line data acquisition is started from the time (t = 0) to the
101

end of experiment for the run, i.e., typically 10 times of mean residence time (10 ).
After the completion of trials for crystallization obtaining necessary conditions
and modifications of experimental set-up, all the crystallization experiments were
performed following statistically designed experimental conditions in Tables 2.2, 2.2,
and 2.2 (see Section 2.7).

4.3.3 Crystallization Condition Monitoring


The actual crystallization conditions were monitored during each experiment.
Such parameters measured were volumetric overflow rate (ml/min), feed solution
temperature ( C), feed solution temperature ( C) at the inlet end-tip, crystal-
lization temperature ( C), mixing speed (rpm) and mixer power (Watts), reaction
pH, solution level (cm), overflow slurry temperature ( C), coolant in and outlet
temperature ( C) to/from the tank jacket. All the equipments measuring above
parameter were calibrated following each of the calibration procedures described
in Appendix C.4. Figure 4.3 illustrates one of such parameter monitoring results
with on-line data acquisition measuring all the parameters in every second till the
completion of crystallization experiment, i.e., typically 10 times of mean residence
time (10 ). The measured parameters during the crystallization experiment of
IS 18 sample are tabulated in Table 4.3. These parameters were averaged from
the values measured from the time when the crystallization temperature reached
steady state to the end of crystallization experiment. All the measured parameters
are within reasonable values with uncertainties for the confidence level of 95%
when considering the precision of crystallization process for the control purpose
and the reproducibility except for the solution level. The high fluctuation of
solution level during the crystallization experiment of IS 18 sample was caused
mostly by the solution turbulence due to the high speed of mixing and somewhat
by the slurry sampling for PSD measurements in every mean residence time ( ).
In addition, the value for solution level is the measure of distance between solution
level indicator and level of solution, thus increase in solution level in Figure 4.3
means decrease in solution level in crystallizer. The pH measurement was made
102

Figure 4.3. An example of monitoring crystallization condition for IS 18 sample.


Note: Mixing speed (rpm) was divided by 10 for better illustration.

with a pH probe temperature compensation after the calibration at 25 C. The


mixer power (Watts) was measured during crystallization with OPTO 22 which
was connected to mixer controller. Yield (g/`) of NaCl particles was measured after
the completion of each crystallization experiment. This section gives the measured
results for crystallization temperature (TC ), feed solution flow rate (Q), mixing
speed (M S), and feed solution temperature (TS ), etc.. The remaining parameters
are presented throughout this chapter and other locations.
Table 4.4 is the result of monitored experimental conditions from the CMSMPR
crystallization which was designed from Table 2.2 in Section 2.7.2. Some of the
experiments were done for two or three times to make sure that the PSD mea-
surements are accurately reproducible. However, the crystallization experiments
at reaction temperature of 25 C and flow rate of 20 ml/min could not be done for
all experimental designs due to the feed tube plugging which was the indication of
crystallization before the feed solution entering the crystallizer. Such crystallization
103

Table 4.3. Measured valuesa for a CMSMPR


crystallization of IS 18 sample.

Description Measured Value


Source Solution Temperature
73.10.31
TS ( C)
Feed Solution Temperature
62.830.27
at inlet end-tip, TS1 ( C)
Crystallization Temperature
29.970.16
TC ( C)
Coolant Inlet Temperature
7.060.14
to Tank Jacket, Tcoolin ( C)
Coolant Outlet Temperature
9.500.18
from Tank Jacket, Tcoolout ( C)
Flow Rate
601.06
(ml/min)
Mixing Speed
875.4810.01
(rpm)
Mixer Power
10.400.15
(Watts)
Reaction pHb 6.990.02
Solution Levelc
36.561.43
(cm)

a
All measured values were taken and averaged from the
time when the crystallization temperature reached steady
state to the end of crystallization experiment. The uncer-
tainty was obtained with confidence level of 95%.
b
Reaction pH was measured during the crystallization
at the crystallization temperature (TC ) with a pH probe
temperature compensation.
c
The value for the solution level is the distance between
solution level indicator and level of solution.

inside feed tube should be avoided since the difference in crystallization environ-
ment, e.g., different solution mixing pattern (laminar and turbulent) and different
solution temperature (level of supersaturation), will make enough difference in crys-
tallization phenomena between the feed tube and the crystallizer. Tables 4.5 and 4.6
are the results of experimental conditions from CMSMPR crystallization which was
designed from Tables 2.3 and 2.4, respectively. The experimental condition for IS
104

Table 4.4. Monitored experimental conditions for CMSMPR crystal-


lization experiment (T samples)a .

Monitored Experimental Condition


Test No. TS b Tcoolin c Tcoolout d TC e Qf MS g
pH
( C) ( C) ( C) ( C) (ml/min) (rpm)
T 04 72.41 7.44 10.58 25.47 40 283.8 7.29
T 05 72.75 7.34 10.74 25.10 40 382.4 7.30
T 06 72.69 7.34 10.55 24.99 40 480.9 7.29
T 07 73.14 -2.81 0.10 25.00 60 281.0 7.06
T 08 72.70 -1.78 0.92 24.90 60 379.6 7.08
T 09 73.24 -1.35 1.17 25.10 60 478.9 7.01
T 10 72.16 22.43 24.27 30.10 20 284.6 7.33
T 11 72.79 22.41 24.19 30.00 20 383.5 7.33
T 12 72.52 22.64 24.28 30.20 20 482.5 7.32
T 13 72.79 15.40 17.84 30.10 40 287.7 7.35
T 14 72.44 16.49 18.86 30.60 40 387.1 7.33
T 15 72.17 16.37 18.91 30.40 40 482.2 7.25
T 16 72.15 6.84 10.85 30.50 60 282.2 7.15
T 17 72.99 8.63 11.16 29.94 60 379.3 7.09
T 18 73.23 8.62 11.23 30.00 60 477.7 7.07
T 29 72.03 28.63 30.16 35.20 20 284.2 7.07
T 20 72.77 28.66 30.24 35.30 20 382.6 7.07
T 21 72.99 28.77 30.32 35.30 20 481.0 7.09
T 22 72.05 22.76 24.90 35.30 40 284.5 7.13
T 23 72.63 23.61 25.81 35.50 40 384.8 7.12
T 24 72.25 23.40 25.45 35.50 40 482.5 7.10
T 25 72.82 15.91 18.23 34.80 60 280.8 6.98
T 26 73.22 17.00 19.25 34.92 60 379.5 6.93
T 27 73.53 16.95 19.27 35.00 60 477.8 6.90

a
Constant experimental parameters: feed solution concentration (C01 ) = CaCl2
(47.631.76 wt%), KCl (1.070.16 wt%), NaCl (0.670.08 wt%) with 95% con-
fidence level, feed solution temperature (TS ) = 70 C or above, mixing design =
standard baffle + turbine.
b
TS : Source solution temperature.
c
Tcoolin : Coolant inlet temperature to the tank jacket.
d
Tcoolout : Coolant outlet temperature from the tank jacket.
e
TC : Crystallization temperature.
f
Q: Flow rate.
g
MS : Mixing speed.
105

Table 4.5. Monitored experimental conditions for CMSMPR crystallization


experiment (TS samples)a .

Monitored Experimental Condition


Test No. TS b TS1 c Tcoolin d Tcoolout e TC f Qg MS h
pH
( C)) ( C) ( C) ( C) ( C) (ml/min) (rpm)
TS 04 73.09 58.53 7.51 9.96 24.50 40 282.1 7.21
TS 05 73.50 60.56 7.63 10.11 25.00 40 380.1 7.17
TS 06 73.06 59.22 8.23 10.69 24.90 40 478.6 7.20
TS 07 72.89 57.49 -2.29 1.20 24.90 60 281.7 7.20
TS 08 73.32 57.10 -1.24 1.77 24.80 60 380.4 7.20
TS 09 73.14 57.09 -0.92 2.25 25.00 60 478.7 7.20
TS 10 73.00 57.44 22.14 23.76 29.80 20 281.9 7.12
TS 11 73.51 56.81 22.54 24.05 30.00 20 380.3 7.11
TS 12 73.59 56.67 22.78 24.48 30.00 20 478.8 7.10
TS 13 73.16 62.90 15.06 17.22 30.10 40 281.9 7.12
TS 14 73.14 59.87 15.12 17.36 29.95 40 380.1 7.11
TS 15 73.22 58.43 15.98 18.03 30.10 40 479.1 7.10
TS 16 72.60 59.64 7.60 10.30 30.50 60 281.9 7.06
TS 17 73.41 59.14 8.07 10.55 30.20 60 380.1 7.09
TS 18 73.10 58.37 7.72 10.52 30.00 60 478.7 7.09
TS 19 73.02 57.95 28.72 30.17 34.90 20 281.8 7.01
TS 20 72.69 57.86 28.92 30.42 35.00 20 380.5 6.99
TS 21 73.58 58.63 28.95 30.34 35.10 20 479 6.98
TS 22 73.34 62.93 21.99 24.01 35.00 40 281.6 7.00
TS 23 72.81 61.61 22.46 24.50 34.90 40 380.5 7.01
TS 24 73.48 60.87 23.00 25.00 35.00 40 478.8 7.01
TS 25 73.31 60.83 15.48 17.97 34.90 60 281.7 6.98
TS 26 73.05 60.06 15.69 17.88 35.00 60 379.3 6.97
TS 27 73.18 59.76 16.30 18.82 35.00 60 480 6.94

a
Constant experimental parameters: feed solution concentration (C02 ) = CaCl2
(46.881.73 wt%), KCl (0.930.15 wt%), NaCl (0.500.05 wt%), feed solution temperature
(TS ) = 70 C or above, mixing design = standard baffle + turbine.
b
TS : Source solution temperature.
c
TS1 : feed solution temperature at the inlet end.
d
Tcoolin : Coolant inlet temperature to the tank jacket.
e
Tcoolout : Coolant outlet temperature from the tank jacket.
f
TC : Crystallization temperature.
g
Q: Flow rate.
h
MS : Mixing speed.
106

20 sample is missing due to data mishandling after the crystallization experiment.

Table 4.6. Monitored experimental conditions for CMSMPR crystallization


experiment (IS samples)a .

Monitored Experimental Condition


Test No. TS b TS1 c Tcoolin d Tcoolout e TC f Qg MS h
pH
( C)) ( C)
( C)
( C)
( C) (ml/min) (rpm)
IS 04 72.87 60.84 8.01 10.52 25.00 40 479.3 7.11
IS 05 73.66 60.43 9.08 11.19 25.10 40 677.5 7.08
IS 06 73.47 61.63 7.47 9.65 24.90 40 875.7 7.08
IS 07 73.09 62.02 -2.70 0.36 24.90 60 481.4 7.12
IS 08 73.07 62.02 -2.36 0.96 25.00 60 678.3 7.12
IS 09 73.23 61.00 -2.29 1.52 25.20 60 876.1 7.11
IS 10 73.36 60.57 22.01 23.74 30.00 20 481.3 6.99
IS 11 73.52 60.11 22.03 23.78 29.90 20 678.4 6.99
IS 12 73.50 60.32 21.99 23.66 29.90 20 876.2 6.98
IS 13 73.09 63.62 14.89 17.13 30.10 40 479.6 7.00
IS 14 73.51 62.16 16.09 17.94 30.00 40 676.7 7.00
IS 15 72.95 62.34 16.48 18.34 30.10 40 874.3 6.98
IS 16 73.25 63.66 6.08 8.41 30.20 60 481.6 7.01
IS 17 73.07 64.05 6.50 8.90 30.10 60 678.6 7.01
IS 18 73.27 62.88 7.09 9.57 30.00 60 875.6 7.00
IS 19 73.63 61.26 28.25 29.68 34.95 20 481.3 6.88
IS 20 n/a n/a n/a n/a 35.00 20 677 n/a
IS 21 73.56 61.04 28.40 29.93 35.10 20 876.2 6.87
IS 22 73.55 63.55 22.47 24.33 34.90 40 478 6.89
IS 23 73.75 64.01 23.96 25.58 35.30 40 676.8 6.86
IS 24 73.27 63.14 22.73 24.61 35.00 40 874.9 6.86
IS 25 73.28 63.98 15.01 17.42 34.90 60 481.7 6.91
IS 26 73.27 63.87 15.50 17.94 35.00 60 678.3 6.90
IS 27 72.99 65.31 15.48 17.90 35.00 60 876.2 6.88

a
Constant experimental parameters: feed solution concentration (C03 ) = CaCl2
(47.090.725 wt%), KCl (1.030.04 wt%), NaCl (0.440.034 wt%), feed solution tempera-
ture (Ts ) = 70 C or above, mixing design = standard baffle + impeller.
b
TS : Source solution temperature.
c
TS1 : Feed solution temperature at the inlet end.
d
Tcoolin : Coolant inlet temperature to the tank jacket.
e
Tcoolout : Coolant outlet temperature from the tank jacket.
f
TC : Crystallization temperature.
g
Q: Flow rate.
h
MS : Mixing speed.
107

4.4 Characterization of Particle Properties


The analysis of the solid crystals after the crystallization is discussed in this
section. Particle properties measured include particle size distribution by LS230
particle size analyzer, particle size and morphology by microscopy and SEM, ele-
ment analysis of solid by EDS, and crystallographic structure of particle by XRD.

4.4.1 Elemental Analysis of Solid


The composition of crystals was analyzed by EDS and XRD. However, the XRD
analysis showed too much background noise due to powder or particles absorbing
moisture from the environment during the analysis even with the attempt covering
the powder with plastic film. Thus, only EDS analysis of solid will be provided and
verified with the elemental analysis of solution by AAS and ICP-AES.
The crystals produced by a CMSMPR crystallizer were NaCl only as shown in
Figure 4.4. The figure presents the comparison of NaCl particles with dried powder.

Figure 4.4. EDS result for the comparison of NaCl particles after the crystalliza-
tion to the powder produced by evaporation of solution. The crystallization was
done at 25 C, 40 ml/min, and 283 rpm.
108

The NaCl particles were produced by crystallization at 25 C, 40 ml/min, and 283


rpm in a CMSMPR crystallizer. The dried powder was prepared by evaporation of
solution mixture containing CaCl2 , KCl, and NaCl with water. No noticeable trace
elements were observed beside the four elements in dried powder. The EDS result
of NaCl particles after the crystallization showed a 22.1% decrease in the Na and
1.0% decrease in the Ca concentration without a trace of K in this analysis which
makes correlation with the precipitation analysis (Section 3.3.1) and the element
analysis of solution (Section 4.2.3). Table 4.7 are tabulated to show the comparison
of concentration in wt% and at% for dried powder (prepared by evaporation of
solution) and NaCl particles (produced by crystallization at 25 C, 40 ml/min,
and 283 rpm in a CMSMPR crystallizer). The Ca in EDS analysis is believed

Table 4.7. Comparison of concentration for


dried powder and NaCl particles.

Element analysis
Element Dried Powdera NaCl Particlesb
(wt%) (at%) (wt%) (at%)
Na 0.97 1.55 37.29 47.86
Cl 65.52 67.77 62.09 51.68
K 1.26 1.18 0 0
Ca 32.25 29.51 0.63 0.46

a
Dried powder was prepared by evaporation of
solution.
b
NaCl particles were produced by the crystalliza-
tion at 25 C, 40 ml/min, and 283 rpm in a CMSMPR
crystallizer.

to be residual from the filtration. These experimental results are in rather good
agreement with the AAS analysis of solution when the experimental error such
as loosing CaCl2 during filtration and analytical errors in AAS and ICP analysis
are considered. However, there was a discrepancy in concentration of CaCl2 when
compared to the precipitation analysis (shown in Table 3.1 in Section 3.3.1) which
109

showed there should be CaCl2 6H2 O particles crystalizing out starting from at 28 C
as the solution cools down to 25 C. This might be the reason arising from the
difference between an ideal thermodynamic system and our system that is close to
ideal mixing but not the same as that in an ideal system.

4.4.2 Crystal Size and Shape

4.4.2.1 SEM and Microscopy


The microscopic photos taken with wet particles a short time after sampling are
shown in Figure 4.5. These particles were taken directly from the crystallizer with
a 2 ml pipette as a slurry sample and placed on a microscopic examining glass slide
for the examination. These wet particles with small volume of slurry were found out
to be quite stable for a relatively long period of time. It was checked by recording
these particles for 10 min. with a camcorder attached to the microscope thus
making sure that these particles were stable enough for the time of examination.
In Figure 4.5 (a) and (b) at 2 , we see delicate star shaped large particles of
100 300 m in size. One of the particles has a broken branch in both figures (a)
and (b). In addition there are some small octahedral particles of various size. In
these early times (2 ), the supersaturation is higher as the cooling from 70 to 25 C
with cooling rate of 0.7 C/min has just been completed giving rise to a unique
set of growth conditions at the beginning of the run. These delicate particles are
more easily broken. In Figure 4.5 (c) and (d) taken at 5 , we see the particles have
filled in to give bi-pyramidal octahedral particles of various size from 5 80 m
in size. The difference in particle morphology in these later particles indicates that
a different crystal growth rate is applicable at intermediate times. In some cases
these particles are aggregated together with aggregate size up to 150 m. And there
is less evidence of breakage in these more robust particles. Finally in Figure 4.5 (e)
and (f) at 10 we see the same bi-pyramidal particles but fewer small particles and
more aggregated particles.
The crystals produced at steady state are single crystals for the most part as
shown in Figure 4.6. In Figures 4.6 and 4.7, one can observe there are indeed some
110

(a) (b)

(c) (d)

(e) (f)

Figure 4.5. Microscopic photos taken (a) and (b) at 2 , (c) and (d) at 5 , and
(e) and (f) at 10 for T 05 sample. Note: The scale of 10 in figure equals 130 m.

agglomerated/composite and broken particles are also observed. Relatively little is


known about the growth of these irregular crystals, so called twin or macle shown
in Figure 4.7, but among the factors that generally cause their formation might be
poor agitation and the presence of certain impurities in the crystallizing solution [3].
Figure 4.8 also shows that there is a fraction of smaller spherical particles at the
early stage of crystallization process before the crystallizer has reached the steady
state. The structure of the assembly of molecules or ions which we call a critical
nucleus is not known. Figure 4.8 shows spherical shapes ranging from 1 to 5 m in
111

Figure 4.6. NaCl particles showing the crystal habit and size (SEM).

Figure 4.7. Agglomerated or composite NaCl crystals known as twin or macle


produced by a CMSMPR crystallization (SEM).

size. The spherical bodies with molecules or solvated ions are believed to be in a
state not too different from that in the bulk fluid, with no clearly defined surface.
It also shows chains formed initially and eventually, as the crystallization precedes,
112

Figure 4.8. NaCl particles produced at an early stage of the crystallization.


Crystals replacing the spherical shape of particles, size ranging from 2 m to 5
m (SEM).

a crystalline lattice which is an octahedral crystal built up as shown in Figure 4.8


replacing these spherical particles. This suggests that there is some special kinetic
phenomenon taking place before reaching the steady state during this crystallization
process and it may not be predicted by the equilibrium calculations.
Figure 4.9 illustrates a crystal structure of octahedron from the NaCl cubic
system. As shown in Figures 4.6 and 4.7, the shape of NaCl crystal shows that the
edges of octahedron are not straight lines. Although, Figure 4.9 shows a possible
way of octahedron crystal structure of NaCl as a unit cell, it has not been clearly
defined how octahedron crystals are forming in a solution containing urea.
As discussed in Section 1.5.1, the dendrite structure with primary arms in
direction of < 100 > is forming and growing until the heat of fusion removes
the constitutional supercooling. This is whey most of the big dendrite structure
was found at the beginning of crystallization. When the solution temperature is
not reached at the steady state there is temperature gradient between the tip of a
growing dendrite and solution due to the different heat conductivities between the
solid and the liquid. If the solid is isothermal (T 0 = 0) the growth rate of the tip
113

Figure 4.9. A regular octahedron inscribed in a cube.

will be given by a similar equation to Equation 1.108 provided TL0 is measured


in the direction of . A solution to the heat-flow equations for a hemispherical
tip shows that the (negative) temperature gradient TL0 is approximately given by
Tc /r where Tc is the difference between the interface temperature (Ti ) and
the temperature of the supercooled liquid far from the dendrite (T ) as shown in
Figure 4.10. Equation 1.108 therefore gives

KL TL0 KL T
= ' (4.1)
LV LV r

Thus for a given T , rapid growth will be favored by small values of r due to the
increasing effectiveness of heat conduction as r diminishes. However T is not
independent of r. As a result of the GibbsThomson effect equilibrium across a
curved interface occurs at an undercooling Tr below Tm given by

2Tm
Tr = (4.2)
LV r

The minimum possible radius of curvature of the tip is when Tr equals the total
undercooling T0 = Tm T . This is just the critical nucleus radius r given
114

Figure 4.10. Temperature distribution at the tip of a growing thermal den-


drite [66].

by (2Tm /LV T0 ). Therefore in general Tr is given by T0 r /r. Finally since


T0 = Tc + Tr Equation 4.1 becomes
r
 
KL 1
' 1 (4.3)
LV r r
It can thus be seen that the tip velocity tends to zero as r r due to Gibbs
Thomson effect and as r due to slower heat conduction. The maximum
velocity is obtained when r = 2r . However, as soon as the solution temperature
reaches the aimed crystallization temperature and stabilizes, the temperature dis-
tribution at the tip of a growing dendrite is so small due to well-mixing in solution
(Tr T0 ) that no such big dendritic particles with primary arms were found
afterwards.

4.4.3 Crystal Size Distribution (CSD)


The CSD is determined with a Beckman Coulter LS230T M , particle charac-
terization size analyzer (PCSA), with the aid of optical microscope and SEM
to verify the particle size and shape. It is needless to say that the accuracy of
CSD measurement is absolutely critical to this crystallization experiment since
the CSD measurement in every mean residence time ( ) will represent how these
115

particles are forming, growing, and evolving in a CMSMPR crystallizer with varying
experimental conditions with time. With many trials and with the application
of CdCl2 impurity, stabilizing NaCl particles in a saturated solution during CSD
measurement was successful within the equipments allowable error, i.e., 1% re-
producibility, see Appendix F for more details. Also, the sampling and the CSD
measurement in timely manner should be emphasized the same as the importance
in the crystallization experiment and the CSD measurement themselves. If any
of the processes during a crystallization experiment fails to meet the procedure
requirements (see Appendix C for the procedures of crystallization, sampling, and
CSD measurement), the whole experiment was discarded and reproduced again
following all the requirements.
Figures 4.11 and 4.12 give CSDs measured at 1.5 and 10.5 , respectively, by
using a LS230 PCSA. With these CSDs measured at every mean residence time ( ),

Figure 4.11. Differential and cumulative volume % of CSD at 1.5 for T 05 sample
crystallization temperature = 250.16 C, feed flow rate = 400.91 ml/min,
mixing speed = 3826.14 rpm, and crystallizer volume = 14009.4 ml. The mass
of particles was 2.6 g/` at 1 . The uncertainty was obtained with confidence level
of 95%.
116

the 3D plot of CSD was constructed to illustrate the evolution of CSDs with time
in a CMSMPR crystallizer. The particle size distributions (x-axis) at the outlet of

Figure 4.12. Differential and cumulative volume % of CSD at 10.5 for T 05


sample crystallization temperature = 250.16 C, feed flow rate = 400.91
ml/min, mixing speed = 3826.14 rpm, and crystallizer volume = 14009.4 ml. The
steady state was achieved after 9 . The uncertainty was obtained with confidence
level of 95%.

the crystallizer as a function of mean residence time (y-axis) and particle volume %
(z-axis) after the crystallizer contents reaches the operating temperature are shown
in Figures 4.13 and 4.14. The 3D plot of CSDs in Figure 4.14 is shown with the
view from z-direction. Initially the particles have dentritic growth arms, which with
time fill in to give bipyramidal particles, see Figure 4.5. The bipyramidal particles
aggregate as the time increases as well. The 3D plots of CSDs in Figures 4.13
and 4.14 show the particle diameter in m (x-axis) taken at different times measured
in units of the mean residence time, (y-axis), and differential volume % (z-axis).
The crystallization experiment for this particular plot was done at crystallization
temperature = 25 C, feed flow rate = 40 ml/min, and mixing speed = 382 rpm with
crystallizer volume = 1400 ml. The CSD measurements were performed every mean
117

Figure 4.13. A 3D plot of particle size distributions as a function of mean residence


time ( ) and differential volume (%) of particles for T 05 sample (see Table 4.4 for
experimental condition), x = particle diameter (m), y = number of mean residence
times ( = 35 min.), z = differential volume (%) crystallization temperature
= 250.16 C, flow rate = 400.91 ml/min, mixing speed = 3826.14 rpm, and
crystallizer volume = 14009.4 ml. The mass of particles was 2.6 g/` at 1 and the
steady state was achieved after 8 . The uncertainty was obtained with confidence
level of 95%.

residence time, i.e., (V /Q = 35 min.), for more than 10 times of mean residence time
( ). At times longer than 10 , there was no change in CSDs. Comparing the CSDs
with the microscopic photos in Figure 4.5 and the SEM in Figure 4.6 shows that
the CSDs accurately represent the particle size distribution for the crystallization
process in a CMSMPR crystallization. In Figure 4.13 the bi/tri modal (small
humps below 50 m in size) is an indication of the aggregation of small particles,
i.e., mostly from the secondary nucleation and the broken particles. No change in
CSDs from 8 to 10 is the indication that the steady state has been reached.
With Figure 4.13 and photos in Figure 4.5, it is clear that the crystal morpholo-
gies and size distributions are evolving with time therefore indicating the combined
effects of nucleation, growth, and breakage, as well as aggregation mechanisms
during the crystallization. Particles with size smaller than 5 m were not observed
118

Figure 4.14. A 3D plot of particle size distributions as a function of mean residence


time ( ) and differential volume (%) of particles with a view from z-direction for
T 05 sample (see Table 4.4), x = particle diameter (m), y = number of mean
residence time ( = 35 min.), z = differential volume (%).

with LS230 even though the instruments size resolution is 0.04 m. Particles
smaller than 5 m in size were also not observed with the microscope even though
it has a size resolution limited to 1.3 m at this magnification. However, the
SEM figure in Figure 4.8 shows particles in size ranging from 2 to 50 m of
spheres and octahedrons. These are dried particles showing the halted process
of crystallization. Some shows agglomerating to larger octahedral particles and
others show aggregating each other. By the comparison of this SEM figure with
microscopic figures and PSD measurements, it can be concluded that the wet
particles smaller than 5 m are not stable due to either redissolution of small
particles into the solution or agglomeration into the larger particles.
For details in the crystallization operation, sampling, and CSD measurement
procedures refer to Section C.6 in Appendix C.
119

4.5 Characterization of CSD


The CSDs measured and microscopic photos taken at every mean residence time
during the crystallization showed distinct characterizations, i.e., the CSD at the
beginning of crystallization and the evolution of CSDs with time, of NaCl crystals
depending on various experimental parameters, such as, solution concentration,
crystallization temperature (TC ), feed flow rate (Q), mixing speed (M S), and type
of mixer. In this section, the characterization of CSDs evolving with time will be
shown depending on such parameters.

4.5.1 Reproducibility of CSD


It is important in this crystallization experiments to show that the CSD mea-
surements should be representative for the condition of each crystallization run.
This was checked by running two or three crystallization experiments at same
condition. However, keep in mind that the crystallization will not produce the
same results since the running conditions cannot be exactly reproduced the same
all the time. With that consideration in mind, the CSDs produced with same
crystallization condition in Figure 4.15 show how closely they are reproducible.
The TS 10-01 sample is the replica of TS 10 sample with as closely controlled
as possible experimental condition. The experimental conditions and results are
also tabulated in Table 4.8 for the comparison purpose. The uncertainty of data
was obtained with confidence level of 95% for all measurements. The precision of
Beckman Coulter LS230T M gives an experimental reproducibility of 1%.

4.5.2 Effect of Solution Concentration


As designed in Table 2.2 and 2.3, the resultant CSDs were indeed affected by
the difference in feed solution concentration between C01 and C02 . The source
concentrations were given in Table C.1 (see Appendix C). The complete sets of
3D plot of CSDs for TS design are given in Appendix D. By comparison, each
of the CSD plots has shown some degree of distinctive characterization due to
the difference in feed solution concentration, i.e., C01 and C02 , also it was shown
120

(a) (b)

(c) (d)

Figure 4.15. Reproducibility of CSDs. The CSDs were measured and plotted as
a function of mean residence time ( = 70 min.) (a) and (b) for T 10 sample,
and (c) and (d) for T 10-01 sample (replica of T 10 sample) with a 6-bladed flat
disc turbine (Rushton type) at mixing speed = 284.6 and 284.8 rpm, crystallization
temperature = 30.1 and 30.3 C, respectively, with same feed flow rate = 20 ml/min
and crystallizer volume = 1400 ml. Steady state was achieved after 9 . The plots
in (b) and (d) are shown with a view from z-direction. See Table 4.8 for detailed
experimental conditions.

that the difference in CSD also depends on crystallization temperature and feed
flow rate in addition to the concentration difference. Figure 4.16 illustrates one
of the comparisons with same experimental conditions except for the solution
concentration. The figure shows two resultant CSDs measured and plotted as a
function of mean residence time; (a) and (b) for T 10 sample, and (c) and (d) for
TS 10 sample with a 6-bladed flat disc turbine (Rushton type) in the middle of
121

Table 4.8. Comparison of experimental conditions and results for CSD


reproducibility with confidence level of 95%a .

TC Q MS Yield Mean Size


Test No.
( C) (ml/min) (rpm) (g/`) Lmean , (m)
T 10 30.10.15 200.85 284.66.4 3.070.27 136.91.8
T 10-01 30.30.19 200.85 284.85.68 2.930.26 137.30.3
Mean 30.20.17 20.00.85 284.76.04 3.00.27 137.11.05
std 0.140.03 0 0.140.51 0.100.01 0.281.06

a
Constant experimental parameters: feed solution concentration (C01 ) = CaCl2
(47.631.76 wt%), KCl (1.070.16 wt%), NaCl (0.670.08 wt%) with 95% confidence
level, feed solution temperature (TS ) = 70 C, mixing design = standard baffle +
turbine, and crystallizer volume = 1.4 `.

crystallizer at mixing speed = 284.6 and 281.9 rpm, crystallization temperature =


30.1 and 29.8 C with same flow rate = 20 ml/min, and crystallizer volume = 1400
ml. The steady state was shown to be achieved after 9 . The plots in (b) and (d)
are shown with a view from z-direction. The experimental conditions are given in
Table 4.4 and 4.5 with details, also the resultant yield and mean size of particles
are given in Table 4.9 and 4.10 shown in Section 4.5.7.
Comparing these CSD plots, it is clear that the small difference in concentration
between C01 and C02 has a small but detectable difference in the resultant CSDs.
This comparison is also a proof of the reproducibility of crystallization process
and accuracy of CSD measurements. The TS 10 sample with low concentration
compared to T 10 sample shows slightly wider CSDs which is the result from lower
supersaturation. The higher supersaturation in T sample with C01 concentration
will have higher nucleation rate resulting in larger number of particles and relaxing
the supersaturation. As these large number of particles grow, however, they will
have more chances to be collided by particle-particle, particle-mixer, and particle-
baffle-wall interactions. This will lead to more particle breakage thus making the
CSDs narrower. Also keep in mind that the breakage of particles depends on not
only the number of particles but also the size and morphology.
122

(a) (b)

(c) (d)

Figure 4.16. Effect of feed solution concentration on CSDs. The CSDs were
measured and plotted as a function of mean residence time ( = 70 min.) (a)
and (b) for T 10 sample, and (c) and (d) for TS 10 sample with a 6-bladed flat
disc turbine (Rushton type) at mixing speed = 284.8 and 281.9 rpm, crystallization
temperature = 30.3 and 29.8 C with same flow rate = 20 ml/min and crystallizer
volume = 1400 ml. Steady state was achieved after 9 . The plots in (b) and (d) are
shown with a view from z-direction. See Table 4.4 and 4.5 for detailed experimental
conditions, also for resultant yield and mean size of particles in Table 4.9 and 4.10.

4.5.3 Effect of Crystallization Temperature


The 3D plots of CSD in Figure 4.17 show the effect of different crystallization
temperature on CSDs beginning with almost the same CSD at 1 2 . The detailed
experimental conditions for each sample are given in Table 4.4 and the yield and
mean size of particles are given in Table 4.9 (see Section 4.5.7). Figure 4.17 (a)
and (b) at 25.1 C crystallization temperature show a drastic change in CSDs with
123

time due to the breakage of large particles and the aggregation of small particles
resulting in distinct transition of CSDs between 3 6 . Figure 4.17 (c) and (d) at
30.6 C show a moderate change with time and Figure 4.17 (e) and (f) at 35.5 C
shows the least change in size with time. Consulting microscopic photos, some
broken particles were found in (a) and hardly any in (c). Steady state is clearly
evident from 8 to 10 for all Figure 4.17 cases.

4.5.4 Effect of Flow Rate (Mean Residence Time)


The effect of different feed flow rate on CSD is given in Figure 4.18. It shows
that an increase in flow rate results in more obvious transition of CSD with time
due to particles having less time to have particle-particle, particle-impeller, and
particle-baffle-wall interactions. The complication of CSD must be a result of such
interactions inside crystallizer with time. The effect of different flow rates on the
resultant mean size at 10 was small.

4.5.5 Effect of Mixing Speed


Figure 4.19 shows the effect of different mixing speed on CSD variation with
time in the crystallizer. The CSDs of (a), (b), and (c) at 1 2 start with different
CSDs indicating different effect of breakage on the initial CSD. At a mixing speed of
288 rpm, Figure 4.19 (a) shows a broad distribution of particles with less breakage
having the mean size of particles around 175 m at 2 , Figure 4.19 (b) at 387 rpm
gives a mean size of 145 m at 2 , and Figure 4.19 (c) at 482 rpm gives a mean size
of 110 m at 2 . These results at 2 show the prominent effect of different mixing
speed on the initial CSDs where the particles are either delicate and easily broken
or more secondary nucleation takes place. At long times in the crystallizer the
CSD is also drastically effected by mixing. The CSD at 10 narrows significantly
at higher mixer rpm. This indicates that either more breakage is occurring at higher
rpm than aggregation or more secondary nucleation takes place thus narrowing the
CSD.
124

4.5.6 Effect of Different Mixer


As depicted in experimental designs as in Tables 2.3 and 2.4, the different types
of mixer, turbine and impeller, were applied and attempted with feed solution
concentrations (C02 and C03 ) which will have as close as possible. As shown in
Section 4.5.2 (see Appendix B), these small differences in concentration should
have some effect on the resultant CSDs. However, as shown in previous section and
will be explained later in more detail, the mixing efficiency will be the major cause
on CSDs compared to other variables in crystallization design. Thus in this section,
an attempt to show the effect of different mixer on CSDs was made regardless of
the difference in concentration.
Compared to a turbine, an impeller needed higher mixing speed due to the
different pattern of mixing, i.e., single or bi-axial mixing, as mentioned in Sec-
tion B.1.2. As a result, the comparison was possible at only one mixing speed (e.g.,
480 rpm) with same crystallization temperature and feed flow rate. Figure 4.20
illustrates one of these examples. As indicated and explained in Section B.1.2,
the turbine is more efficiency of mixing than the impeller. The efficiency in mixing
with the turbine means much more breakage of particles due to the interactions, i.e.,
particle-particle, particle-mixer, and particle-baffle-wall. The CSDs in Figure 4.20
clearly show the effect of different types of mixer.
125

(a) (b)

(c) (d)

(e) (f)

Figure 4.17. Crystallization temperature effect on CSDs. The CSDs were


measured and plotted as a function of mean residence time ( = 35 min.)
Crystallization temperature (a) and (b) at 25.1 C (T 05 sample), (c) and (d) at
30.6 C (T 14 sample), and (e) and (f) at 35.5 C (T 23 sample) at same flow rate =
40 ml/min, mixing speed = 385 rpm, and crystallizer volume = 1400 ml. Steady
state was achieved after 9, 8, and 7 , respectively. The plots in righthand side
are shown with a view from z-direction. See Table 4.4 for detailed experimental
conditions.
126

(a) (b)

(c) (d)

(e) (f)

Figure 4.18. Flow rate effect on CSDs. The CSDs were measured and plotted
as a function of mean residence time ( ) Flow rate (a) and (b) 20 ml/min (
= 70 min.) (T 11 sample), (c) and (d) 40 ml/min ( = 35 min.) (T 14 sample),
and (e) and (f) 60 ml/min ( = 23.3 min.) (T 17 sample) at same crystallization
temperature = 30 C, mixing speed = 385 rpm, and crystallizer volume =
1400 ml. Steady state was achieved after 8, 9 and 9 , respectfully. The plots in
righthand side are shown with a view from z-direction. See Table 4.4 for detailed
experimental conditions.
127

(a) (b)

(c) (d)

(e) (f)

Figure 4.19. Mixing speed effect on CSDs. The CSDs were measured and plotted
as a function of mean residence time ( = 35 min.) Mixing speed (a) and (b) at
287.7 rpm (T 13), (c) and (d) at 387.1 rpm (T 14), and (e)and (f) at 482.2 rpm (T
15) at same crystallization temperature = 30 C, feed flow rate = 40 ml/min, and
crystallizer volume = 1400 ml. Steady state was achieved after 8 . The plots in
righthand side are shown with a view from z-direction. See Table 4.4 for detailed
experimental conditions.
128

(a) (b)

(c) (d)

Figure 4.20. Effect of different mixer on CSDs. The CSDs were measured and
plotted as a function of mean residence time ( = 23.3 min.) (a) and (b)
with a 6-bladed flat disc turbine (Rushton type) (TS 18) and (c) and (d) with
a 6-bladed pitch blade impeller (IS 16) at mixing speed = 479 and 482 rpm,
crystallization temperature = 30.0 and 30.2 C with same feed flow rate = 60
ml/min and crystallizer volume = 1400 ml. Steady state was achieved after 9 and
8 , respectively. The plots in righthand side are shown with a view from z-direction.
See Table 4.5 and 4.6 for detailed experimental conditions.
129

4.5.7 Overall Effect on Yield and Lmean


The yield (g/cm3 ) and mean particle diameter, Lmean (m) were measured at
the end of each experiment that is at 10 . These are tabulated in Tables 4.9, 4.10,
and 4.11. Three of yield data were missing due to mishandling of solution sample
preparation for the AAS analysis in Table 4.9. The uncertainties for the yield and
mean size of particles of each experiment were obtained with confidence level of
95%. Figure 4.21 shows mean size of particles (diameter) vs. sample number for T
samples. This figure shows clearly the major effect of mixing speed on the resultant

Figure 4.21. Effect of crystallization temperature (TC ), feed flow rate (Q), and
mixing speed (M S) on the mean size of particles (Lmean ) at steady state (10 ) for
T samples (see Table 4.4).

mean size of particles (Lmean ) with weak correlation of crystallization temperature


(TC ) and feed flow rate (Q) at 10 . This relationship is shown in Figure 4.22 along
with the linear regression statistics with confidence level of 95%. The spread of
data points with same mixing speed was due to mostly the effect of feed flow rate
and little with the effect of crystallization temperature. Also noticed the spread of
130

Table 4.9. Yield and mean size of


NaCl particles (T samples)a .

Yieldb Mean Sizec


Test No.
(g/cm3 ) Lmean , (m)
T 04 4.710.27 135.272.40
T 05 n/a 99.950.87
T 06 3.460.38 87.940.47
T 07 2.680.55 139.901.98
T 08 4.710.29 102.130.12
T 09 3.000.26 81.910.24
T 10 3.070.27 136.901.80
T 11 2.810.25 114.930.25
T 12 4.760.25 107.200.28
T 13 4.240.31 129.502.73
T 14 5.790.35 106.420.80
T 15 3.870.27 92.101.57
T 16 n/a 137.632.12
T 17 2.910.35 107.432.44
T 18 3.750.50 87.220.35
T 19 4.260.25 131.273.29
T 20 4.850.25 124.501.18
T 21 3.740.25 114.160.54
T 22 5.260.35 135.562.21
T 23 3.660.25 113.681.52
T 24 n/a 99.500.95
T 25 4.460.45 144.232.25
T 26 5.100.25 114.800.44
T 27 3.670.58 89.280.10

a
The uncertainties for yield and mean
size of particles were obtained with 95%
confidence level.
b
Yield of NaCl particles were obtained
from the results of AAS analysis.
c
Lmean : Mean Size of Particles (diame-
ter) obtained from the PSD measurements
by LS230.

data was increased with increase in mixing speed. With this regression, the mean
particle size at steady state (10 ) can be predicted within the standard error of
8.587.
131

Figure 4.22. Linear regression statics on the mean size of particles (Lmean ) vs.
mixing speed at steady state (10 ) for T samples (see Table 4.4 and 4.9). The
regression was performed with confidence level of 95%.

Table 4.10 shows the yield and mean particle size at steady state for TS samples.
The yield in TS samples also shows relatively large uncertainty comparable to yield
in T samples and has shown a weak relationship with crystallization temperature
(TC ), flow rate (Q), and mixing speed (M S) as shown in Figure 4.23. This was most
likely the result of difficulties inherent in obtaining accurate liquid compositions.
The sample solution containing solid particles at room temperature requires further
treatments such as making dilution with distilled water and adding 5% HCl. In
addition to the difficulties obtaining precise concentration of aqueous electrolyte so-
lutions, the accuracy of solution concentration was violated during such treatments
for T and TS samples.
The mean size of particles in TS samples shows a strong relationship with
mixing speed (MS) and weak relationships with flow rate (Q) and crystallization
temperature (TC ) as M S > Q > TC shown in Figure 4.24 same as in the case of
T samples which though showed less clear effect of crystallization temperature in
Figure 4.21. The detailed statistical analysis for the effects (main and interaction)
132

Table 4.10. Yield and mean size of


NaCl particles (TS samples)a .

Yieldb Mean Sizec


Test No.
(g/cm3 ) Lmean , (m)
TS 04 3.350.50 135.470.99
TS 05 3.380.50 110.382.29
TS 06 3.93 0.49 90.190.86
TS 07 1.540.50 124.872.16
TS 08 4.370.77 101.11 3.38
TS 09 2.860.51 82.360.42
TS 10 2.830.49 160.933.83
TS 11 2.670.49 133.573.75
TS 12 4.620.50 114.76 1.68
TS 13 4.100.61 144.133.01
TS 14 3.051.08 118.981.58
TS 15 3.630.53 100.873.59
TS 16 2.560.49 127.903.40
TS 17 3.571.18 100.651.34
TS 18 n/a 86.580.72
TS 19 4.120.49 173.081.11
TS 20 3.710.49 149.831.01
TS 21 3.600.50 136.520.32
TS 22 4.320.49 156.272.63
TS 23 3.520.49 125.831.28
TS 24 4.540.51 103.900.52
TS 25 4.120.49 131.002.38
TS 26 2.560.49 110.003.11
TS 27 4.230.50 92.950.73

a
The uncertainties for yield and mean
size of particles were obtained with 95%
confidence level.
b
Yield of NaCl particles were obtained
from the results of AAS analysis.
c
Lmean : Mean Size of Particles (diame-
ter) obtained from the PSD measurements
by LS230.

of independent variables to dependent product variables will be discussed further in


Chapter 5. The mean particle size in TS samples was larger than that in T samples.
This can be explained by the fact that the TS samples with lower concentration
133

Figure 4.23. Yield of NaCl product at 10 for TS samples.

Figure 4.24. Effect of temperature, flow rate, and mixing speed on the mean size
of particles (Lmean ) at steady state (10 ) for TS samples (see Table 4.5).
134

compared to T samples result in lower supersaturation, thus resulting in lower


nucleation rate. As a result, the particles having less interaction with others and
having less chance of collision or breakage resulted in relatively larger particles
for the same operating conditions. Figure 4.25 shows regression statics of mean
particle size on mixing speed for TS samples along with the indication of feed flow
rate and crystallization effects. In this particular regression, the value of R2 was

Figure 4.25. Linear regression statics on the mean size of particles (Lmean ) vs.
mixing speed at steady state (10 ) for TS samples (see Table 4.5 and 4.10). The
regression was performed with confidence level of 95%.

smaller than the T samples because of the stronger correlation of crystallization


temperature and feed flow rate to mean particle size in TS samples. As showed
in Figures 4.24 and 4.25, the mean particle size decreased with respect to increase
in feed flow rate and mixing speed and increased with temperature increase. This
clearly shows the mean particle size dependence as M S > Q > TC .
The product yield for IS samples also has large uncertainty as shown in Ta-
ble 4.11 but it was the smallest compared to other two data sets due to more careful
treatments for the sample preparations. Therefore, only in this IS sample case the
135

yield and mean size of particles data obtained for the IS samples will be attempted
for further analysis. Figure 4.26 shows the yield (g/`) of NaCl particles vs. sample

Table 4.11. Yield and mean size of


NaCl particles (IS samples)a .

Yieldb Mean Sizec


Test No.
(g/cm3 ) Lmean , (m)
IS 04 3.300.24 135.531.18
IS 05 3.190.27 100.145.37
IS 06 3.380.24 84.061.82
IS 07 2.720.22 114.470.99
IS 08 2.640.24 88.990.72
IS 09 2.890.24 74.342.19
IS 10 3.570.24 127.701.29
IS 11 3.490.24 109.403.82
IS 12 3.66 0.22 92.712.33
IS 13 3.220.35 149.003.93
IS 14 n/a 118.630.99
IS 15 3.160.27 97.182.37
IS 16 3.050.24 124.002.46
IS 17 2.640.27 96.700.72
IS 18 3.190.43 83.073.83
IS 19 3.190.24 141.803.88
IS 20 2.970.24 118.873.79
IS 21 3.220.22 102.341.12
IS 22 2.690.58 158.204.61
IS 23 2.830.27 122.30 1.53
IS 24 2.750.27 102.783.57
IS 25 2.940.22 135.102.85
IS 26 2.890.24 108.930.63
IS 27 2.860.27 76.373.92

a
The uncertainties for yield and mean
size of particles were obtained with 95%
confidence level.
b
Yield of NaCl particles were obtained
from the results of AAS analysis.
c
Lmean : Mean Size of Particles (diame-
ter) obtained from the PSD measurements
by LS230.
136

number. This figure shows the effects of crystallization temperature and feed flow

Figure 4.26. Yield of NaCl product at 10 for IS samples.

rate on the yield of NaCl at 10 . As the feed flow rate and the crystallization
temperature increase the yield of NaCl particles decreased. The effect of mixing
speed on yield was small. Figure 4.27 shows the yield of NaCl vs. feed flow rate
with temperature groupings. The linear regression was poor due to the effects of
crystallization temperature and mixing speed. As mentioned earlier, this analysis
along with the data shown in Figure 4.26 provides justification for the effect of feed
flow rate on the yield of NaCl particles, i.e., the particles having long residence
time (space time in crystallizer) will grow more as a result of particles spending
more time in crystallizer. Figure 4.28 shows the overall effect of crystallization
temperature, feed flow rate, and mixing speed on the mean size of particle (Lmean )
at 10 for IS samples. At 25 C, the experiment with feed flow rate of 20 ml/min
was not successful due to crystallization in the feed before entering the crystallizer.
Figure 4.28 together with Figure 4.29 indicates the effects of all three variables, i.e.,
137

Figure 4.27. Linear regression statics on the product yield (g/cm3 ) vs. flow
rate (ml/min) at steady state (10 ) for IS samples (see Table 4.6 and 4.11). The
regression was performed with confidence level of 95%.

Figure 4.28. Effect of temperature, flow rate, and mixing speed on the mean size
of particles (Lmean ) at steady state (10 ) for IS samples (see Table 4.6).
138

crystallization temperature, flow rate, and mixing speed, on resultant mean size of
particles. Figure 4.29 shows the regression analysis of mean particle size on mixing

Figure 4.29. Linear regression statics on the product mean size of particles (Lmean )
vs. mixing speed (rpm) at steady state (10 ) for IS samples (see Table 4.6 and 4.11).
The regression was performed with confidence level of 95%.

speed for IS samples. As shown in these two Figures 4.28 and 4.29, the particle
size decreases with increase in mixing speed which showed the most profound effect
on resultant particles, and increases with increase in crystallization temperature
within the experimental conditions, i.e., 25 35 C. The effect of feed flow rate
shows somehow interesting results in this system; the particle size increases with
increase in flow rate from 20 to 40 ml/min and decreases with further increase to
60 ml/min. This feed flow rate effect was large at lower mixing speed but small at
higher mixing. This was the indication showing the dominant effect of mixing on
resultant particles at higher mixing reducing other effects.
A correlation analysis was performed to measure the relationship between two
data sets for all three experimental designs. The two data sets are independent
of the unit of measurement in this analysis; a variable and a product or between
139

two products. The population correlation calculation returns the covariance of two
data sets divided by the product of their standard deviations based on the following
formulas.
cov(X, Y )
X,Y = (4.4)
X Y
where
1
x2 = (Xi x )2
n
1
y2 = (Yi y )2
n
This analysis is to determine whether two ranges of data move together, i.e.,
whether large values of one set are associated with large values of the other (positive
correlation), whether small values of one set are associated with large values of the
other (negative correlation), or whether values in both sets are unrelated (correla-
tion near zero). The results are shown in Table 4.12. The analysis indicates a strong

Table 4.12. Correlation analysis of inde-


pendent variables to yield and particle mean
size.

Sample TC a Qb MSc
Yieldd 0.3289 -0.0872 -0.1563
T
Lmean 0.2246 -0.1881 -0.8989
Yield 0.3320 -0.2081 0.2870
TS
Lmean 0.3732 -0.6084 -0.7292
Yield -0.1869 -0.6500 0.0747
IS
Lmean 0.3217 -0.2947 -0.8516

a
TC : Crystallization temperature ( C).
b
Q: Flow rate (ml/min).
c
MS: Mixing speed (rpm).
d
Yield of NaCl particles (g/cm3 ).

correlation between mixing speed and mean size of particles. The yield dependence
on feed flow rate and crystallization temperature should have strong function as
discussed and showed previously in precipitation analysis in Section 3.3.1. The
140

explanation for the weak correlation in T and TS samples might be the large
uncertainty in yield which was obtained by AAS analysis. As a result, the yield
data in Table 4.9 and 4.10 are used to show merely the reduction of NaCl and
cannot be used in further analysis.

4.6 Batch Breakage and Aggregation Experiments


Additional experiments have been performed that focus on aggregation and
attrition without the influence of nucleation and crystal growth being active. In
this case the crystallizer, at the end of the crystallization experiment, is kept
at a constant temperature and the feed and output flows are stopped. For this
experiment the crystallizer operated as a batch reactor and in a very short period
of time the supersaturation present at the end of crystallization is relieved so that
nucleation and growth are eliminated. Under these conditions only aggregation
and attrition are active mechanisms. Tables 4.13 and 4.14 give the monitored
experimental conditions for the BT and BI samples which were designed from
Table 2.5 with various conditions in Section 2.8. All the measured values and
uncertainties in these two tables are given with confidence level of 95%. In the
experiment shown in Figure 4.30, plotted as x = particle diameter (m), y = elapsed
time (hours), and z = differential volume % (the figures on righthand are shown
with a view from z-axis), the particle size distribution increases in mean size and
broadens with time in the batch reactor indicating that aggregation dominates over
breakage for these crystals in their mother liquor. For the detailed experimental
conditions for each sample, refer to Table 4.13. For all experiments the starting
volume of slurry was fixed to 1,400 ml.
The mean size of particles (m) as a function of time (hour) is tabulated in
Tables 4.15 and 4.16 for BT and BI samples. In both tables, some of data is
missing due to the difficulty in CSD measurement which was caused by aggregation
of small particles. Also, the data without uncertainty was obtained from only single
measurement due to the same problem.
141

Table 4.13. Monitored experimental conditions


for batch breakage and aggregation experiment (BT
samples)a .

Monitored Experimental Condition


Test No. TR b MS c
pH
( C) (rpm)
BT 01 25.470.138 283.346.103 7.280.009
BT 02 25.52 0.142 382.196.828 7.280.009
BT 03 25.610.138 480.627.439 7.270.008
BT 04 29.840.240 287.345.957 7.350.008
BT 05 30.590.131 386.437.251 7.340.008
BT 06 31.010.381 482.116.004 7.230.007
BT 07 35.230.148 284.115.435 7.130.007
BT 08 35.360.153 384.557.851 7.130.007
BT 09 35.380.133 482.346.248 7.110.007

a
Constant experimental parameters: Feed solution concen-
tration (C01 ) = CaCl2 (47.631.76 wt%), KCl (1.070.16
wt%), NaCl (0.670.08 wt%) with 95% confidence level, mix-
ing design = standard baffle + turbine. All the measured
values are given with confidence level of 95%.
b
TR : Reaction temperature.
c
MS : Mixing speed.

4.6.1 Temperature Effect on Breakage and Aggregation


Unfortunately, it was not easy to see the temperature effect on batch breakage
and aggregation as shown in Figure 4.30 due to the difference in the starting CSD.
Figure 4.31 is obtained by plotting the data from Tables 4.15 and 4.16. Interestingly,
both figures (a) and (b) show the same trends; the slope of regression lines are in
the same order as 30 > 25 > 35 C of reaction temperature. This phenomena can
be explained with a breakage/aggregation mechanism. At 25 C, the particles are
exposed to more breakage and resulted in more aggregation due to large number
of starting particles compared to 30 and 35 C. The particles at 35 C starting
with larger mean size but small number of particles will experience less breakage,
however, due to relatively high temperature some of smaller broken particles can be
redissolved into solution, therefore reducing the aggregation and inducing particle
142

Table 4.14. Monitored experimental conditions for batch break-


age and aggregation experiment (BI samples)a .

Monitored Experimental Condition


Test No. TR b MS c MP d
pH
( C) (rpm) (Watts)
BI 01 25.180.217 479.479.257 4.560.100 7.110.010
BI 02 25.000.163 677.128.006 7.200.089 7.080.007
BI 03 25.150.131 875.937.629 10.230.232 7.070.008
BI 04 29.910.227 479.698.487 4.520.098 7.000.011
BI 05 30.040.188 677.117.797 7.190.102 6.980.055
BI 06 30.190.223 873.786.544 10.180.103 6.980.012
BI 07 34.840.184 479.956.312 4.670.069 6.880.026
BI 08 34.810.147 676.976.471 7.210.082 6.880.008
BI 09 34.940.163 874.845.627 10.270.094 6.870.007

a
Constant experimental parameters: Feed solution concentration (C03 ) =
CaCl2 (47.091.76 wt%), KCl (1.030.16 wt%), NaCl (0.440.08 wt%) with
95% confidence level, mixing design = standard baffle + impeller. All the
measured values are given with confidence level of 95%.
b
TR : Reaction temperature.
c
MS : Mixing speed.
d
MP : Mixing power.

growth due to increased supersaturation from such redissolution of particles in this


experiment. The monitored pH values in Tables 4.13 and 4.14 can be also one of
the explanation for such a phenomenon, i.e., redissolution of particles resulting in
overall particle growth therefore reducing pH values. Also, notice that the small
increase in mean size of particles in Figure 4.31 (b) which shows overall less breakage
and aggregation compared to (a) showed the larger particle size at the end of
experiment.

4.6.2 Mixing Speed Effect on Breakage and Aggregation


Figure 4.32 shows the effect of mixing speed on resultant mean size of particles as
a function of time (hour). This figure clearly indicates that the growth of particles
in this experiment is mostly from the breakage and aggregation of particles. In
other word, higher the mixing speed, higher the breakage of particles which will
143

Table 4.15. Mean size of particles measured as a function of time (hour) for
BT samplesa .

Mean Size of Particles (m)


Test No.
t=0 0.5 hour 1 hour 2 hour 3 hour
BT 01 133.403.03 134.902.60 138.002.96 138.003.87 138.404.26
BT 02 99.960.88 100.101.01 100.501.87 105.202.57 108.202.35
BT 03 n/a n/a n/a n/a n/a
BT 04 129.502.75 139.20 138.50 n/a 156.90
BT 05 106.400.82 109.900.97 110.601.28 116.801.50 122.601.91
BT 06 90.390.02 92.060.22 94.420.76 100.400.81 103.200.87
BT 07 135.602.19 137.002.42 136.003.54 140.803.77 140.702.60
BT 08 113.701.52 113.401.75 114.302.41 116.701.66 120.002.80
BT 09 99.490.93 98.371.21 102.201.63 104.902.71 n/a

a
Constant experimental parameters: mixing design = standard baffle + turbine. All the
measured values are given with confidence level of 95%.

Table 4.16. Mean size of particles measured as a function of time (hour) for BI
samplesa .

Mean Size of Particles (m)


Test No.
t=0 0.5 hour 1 hour 2 hour 3 hour
BI 01 135.500.62 135.400.87 136.501.14 138.50 2.24 137.001.69
BI 02 98.60 99.95 96.19 105.70 n/a
BI 03 n/a n/a n/a n/a n/a
BI 04 149.002.01 150.401.57 149.001.91 153.702.41 153.001.77
BI 05 118.600.50 120.30 120.700.89 122.400.78 122.400.09
BI 06 96.941.86 100.900.17 n/a n/a n/a
BI 07 158.202.36 n/a 157.901.78 157.702.73 158.502.22
BI 08 122.300.78 n/a n/a n/a n/a
BI 09 101.301.17 101.501.99 105.100.45 107.30 107.400.36

a
Constant experimental parameters: mixing design = standard baffle + impeller. All the
measured values are given with confidence level of 95%.

result in prevention of particles from aggregation. Even with high speed of mixing
for BI sample, the effect of mixing with an impeller was negligible compared to a
turbine.
144

4.6.3 Effect of Different Mixer on Breakage and Aggregation


As shown in Figure 4.33, the effect of mixing with a 6-bladed flat disc turbine
(Rushton type) noticeably larger than that with a 6-bladed pitch blade impeller.
Figure 4.34 is to show the effect of different type of mixer, i.e., turbine and impeller,
with almost same condition of experiment beside the mixer on resultant breakage
and aggregation. This figure presents almost 3-fold higher growth rate with a
turbine than an impeller. The growth mechanism in this experiment was mostly
attributed to the aggregation of particles.

4.7 Summary
The crystallization experiment at 30 C showed the reduction of NaCl only which
was comparable with a OLI analysis showing a reasonable prediction of reduction in
NaCl after the crystallization. However, the precipitation analysis with C01 solution
concentration showed CaCl2 6H2 O crystallizing out from 28 C. The EDS analysis
of solids which were produced by the crystallization at 25 C showed no CaCl2 . The
discrepancy between experiment and OLI analysis was due to the difference in real
and ideal system of kinetics and thermodynamics.
The realtime crystallization conditions were monitored for each experiment with
on-line data acquisition measuring all parameters in every second till the completion
of crystallization experiment typically 10 times of mean residence time (10 ). The
SEM figures with the aid of microscopic figures demonstrate the detailed particle
shape and size to explain the possible phenomena of how these particles are evolving
in a CMSMPR crystallizer. Initially the mean particle size was 180 m and
with time the mean size shifts to 120 m. In other experiments, this shift of
mean size with operating time was always observed and it was affected mostly by
crystallization temperature, feed flow rate, mixer rpm, and different type of mixer.
In addition, a small satellite peak at 20 m in Figure 4.13 developed with time in
all of our experiments. These particle size distribution results were the indicative
of the shape change from dendritic to bipyramidal particles and more breakage of
145

the aggregates with time. The particle size distribution stabilized at long operating
times as would be expected for a CMSMPR crystallizer operating at steady state.
The experimental results with the 3D plots of CSDs clearly showed the proof
of CSD reproducibility which demonstrated the precision of crystallization process
and CSD measurement. With the accuracy in CSDs measured and microscopic
photos taken at every mean residence time during the crystallization showed distinct
characterizations, i.e., the CSD at the beginning of crystallization and the evolution
of CSDs with time, of NaCl crystals depending on various experimental parameters,
such as, feed solution concentration, crystallization temperature (TC ), feed flow rate
(Q), mixing speed (M S), and different type of mixer on resultant CSD with vivid
graphical representations.
Overall effects of crystallization temperature, feed flow rate, and mixing speed
on products, such as mean size of particles and yield, were discussed in Section 4.5.7.
However, due to the result of difficulties inherent in obtaining accurate liquid
compositions the effects on yield were not clear. The sample solution at room
temperature containing solid particles required further treatments such as making
dilution with distilled water and adding 5% HCl. In addition to the difficulties
obtaining precise concentration of aqueous electrolyte solutions, the accuracy of
solution concentration was violated during such treatments for T and TS samples.
However, with more careful treatments for IS samples it was possible to show the
effects of crystallization temperature and feed flow rate on the yield of NaCl at
10 . It was found out that as the feed flow rate and the crystallization temperature
increase the yield of NaCl particles decreased. The effect of mixing speed on yield
was small. The mean particle size decreased with respect to increase in feed flow
rate and mixing speed and increased with temperature increase. This clearly showed
the mean particle size dependence on M S > Q > TC . The analysis along with the
data shown in Figure 4.26 provided justification for the effect of feed flow rate on
the yield of NaCl particles, i.e., the particles having long residence time (space
time in crystallizer) will grow more as a result of particles spending more time in
crystallizer. For the crystallization temperature at 25 C, the experiment with feed
146

flow rate of 20 ml/min was not successful due to crystallization in the feed before
entering the crystallizer.
Additional experiments performed focusing on aggregation and attrition without
the influence of nucleation and crystal growth being active. In this experiment, the
effect of reaction temperature on mean size of particles as a function of time showed
in the order of 30 > 25 > 35 C for both of experimental designs (BT and BI) in
which it clearly showed the competing mechanism of breakage and aggregation.
With lower speed of mixing the lager particle was obtained as the high speed of
mixing resulted in more breakage and less aggregation of particles. The mixing
speed effect with an impeller was negligible. As expected, the turbine creating
biaxial mixing of solution below and above the mixer was resulted in almost 2-fold
higher particle growth by aggregation than the impeller.
As described in Section 1.2, all of the collected conditions, parameters, and
product data will be analyzed with the application of known model in next Chapter.
Purification of CaCl2 from a solution, CaCl2 + KCl + NaCl, involved nucleation
and growth of NaCl crystals from the solution without much aggregation. As a
result, we should be able to run a validation of the OLI model with this system.
147

(a) (b)

(c) (d)

(e) (f)

Figure 4.30. CSDs in differential volume % as a function of time (hours) for


reaction temperature (a) and (b) at 25.520.142 C for BT 02, (c) and (d) at
30.590.131 C for BT 05, and (e) and (f) at 35.360.153 C for BT 08 at nearly
same mixing speed = 382.96.8, 386.47.3, and 384.67.9 rpm, respectively, and
starting reactor volume = 1400 ml. x = particle diameter (m), y = time (hours),
and z = differential volume (%). The plots in righthand side are shown with a view
from z-direction. See Table 4.13 for detailed experimental conditions.
148

(a)

(b)

Figure 4.31. Effect of reaction temperature on mean size of particles as a function


of time (hours) (a) with a 6-bladed flat disc turbine (Rushton type) for BT 02, 05,
and 08 samples at mixing speed = 3846.0 rpm, and (b) with a 6-bladed pitch blade
impeller for BI 01, 04, and 07 samples at mixing speed = 4796.0 rpm, reaction
temperature near at 25, 30, and 35 C starting with same reactor volume = 1400
ml. See Table 4.13 and 4.14 for detailed experimental conditions.
149

Figure 4.32. Effect of mixing speed on mean size of particles as a function of


time (hours) with a 6-bladed flat disc turbine (Rushton type) for BT 04, 05, and
06 samples at mixing speed = 287.345.96, 386.437.25, and 482.116.0 rpm with
reaction temperature near at 30 C starting with same reactor volume = 1400 ml.
See Table 4.13 for detailed experimental conditions.
150

(a) (b)

(c) (d)

Figure 4.33. Effect of different mixer on CSDs as a function of time (hours) and
differential volume (%) (a) and (b) with a 6-bladed flat disc turbine (Rushton
type) for BT 06, and (c) and (d) with a 6-bladed pitch blade impeller for BI 04
at mixing speed = 482.16.0 and 479.78.49 rpm, crystallization temperature =
31.00.38 and 29.910.23 C starting with same reactor volume = 1400 ml. The
plots in righthand side are shown with a view from z-direction. See Table 4.13
and 4.14 for detailed experimental conditions.
151

Figure 4.34. Effect of different type of mixer on mean size of particles as a


function of time (hours) with a 6-bladed flat disc turbine (Rushton type) for
BT 06 and with a 6-bladed pitch blade impeller for BI 04. Mixing speed =
482.116.0 and 479.698.487 rpm, respectively, with reaction temperature near
at 30 C starting with same reactor volume = 1400 ml. See Table 4.13 and 4.14 for
detailed experimental conditions.
CHAPTER 5

ANALYSIS AND MODEL VALIDATION

The CMSMPR models described in Section 1.2.1 to evaluate the results from
CMSMPR experiments are only valid if the assumptions defined in that section
hold. Deviations from this idealized CMSMPR model may be caused by the
following effects:

1. kinetics of crystal growth (e.g., size dependent growth, growth rate disper-
sion) [63, 96].

2. attrition and breakage caused by mechanical stress of the crystals.

3. agglomeration and destruction of agglomerates or dissolution of fines.

4. classification due to fluid dynamics and product discharge [97].

5. insufficient mixing leading to differences of system properties.

6. insufficient mixed product removal.

7. non-steady state operating conditions.

Usually any operating condition will produce some of these effects. For instance
increase of stirrer speed causes better mixing but more attrition. Figure 5.21 in
Section 5.3.3 illustrates qualitatively how some of these deviations from the idealized
CMSMPR model can influence the steady state crystal size distribution plotted in
a semi-logarithmic population density plot. The straight solid line in Figure 5.21
represents the population density expected according to the idealized CMSMPR
model. In most cases it is difficult to evaluate the specific reason for the deviation
from the linear plot because the same type of deviation can result from several
153

causes. Therefore the shape of a population density plot does not explain the
particular effect that has caused the deviation in a particular experiment. The
purpose of this chapter is to illustrate the particular effects to the resultant CSD
focusing on the effect of independent variables, i.e., feed solution concentration,
crystallization temperature, feed flow rate, mixing speed, and different type of
mixer, to the dependent variables, i.e., number of nuclei, nucleation rate, growth
rate and aggregation rate.

5.1 Evaluation of Nucleation and Growth Rates


For well-mixed reactor, it is usually assumed that (1) particles are formed
only by nucleation and increase in size only through growth, (2) the processes
of breakage, attrition, and agglomeration are negligible, and (3) the shape factors
for particles are not a function of size so that the use of a single characteristic
size can be used as the dimension to describe the entire size range. Using the
same approach as was described previously in Section 1.2.1, it was found out that
n0 = 2.34 1013 1.986 1014 /m4 , G = 4.40 103 1.01 102 m/s, therefore
the nucleation rate to be B0 = n0 G = 1.03 105 2.01 106 /m3 s within our
experimental conditions. To have a general idea of growth rate, it was found out
from the literatures [3] that depending on the experimental conditions the growth
rates exhibited by inorganic salts in aqueous solution generally lie well within the
range from 101 to 103 m/s which is comparable with our experimental results.
From the number density plot (Figure 5.1), it is clear that particle size smaller
than 5 m was either not stable or agglomerate to the bigger particles during the
CSD measurement. This was verified with SEM, microscopy, and CSD measure-
ment in Section 4.4.3. Also the difference in population density in Figure 5.1 at
1.7 and 10.3 clearly shows different nucleation and growth rate which are the
indication of different particles growth mechanism in the crystallizer with time as
discussed in Section 4.4.2.1. Figures D.1 and D.3 in Appendix D.1 show plots of
differential number % vs. particle size, L (m), with microscopic photos inside to
aid the clear understanding of the particle shapes and distributions.
154

Figure 5.1. A plot of population number density distribution at 1.7 and 10.3
from the starting of a CMSMPR crystallizer.

Crystals larger than approximately 200 m were reduced in size by attrition


induced by particle collisions. Only small attrition fragments act as effective nuclei.
It has previously been shown that in industrial crystallizers attrition fragments are
generated in the size range between a few micrometers up to approximately 100
m and that only a small percentage (roughly 1%) of them (fast growers) grow
into the product size range. The effective rate of secondary nucleation (Na,ef f ) is
proportional to the collision frequency, the total number of the attrition fragments
(Na,tot ), and the ratio (Na,ef f /Na,tot ) which depends on supersaturation.

5.2 Application of Hounslows Equation


For the combined aggregation, growth, and nucleation, the PBE was obtained
by setting the breakage function and specific rate of breakage to zero, a feed
distribution consisting of nuclei of zero size, additionally, the aggregation kernel
and growth function to be constant. The analytical solution to the resulting PBE
has been derived by [98, 99]
155

I1 (x)
n(v) = 2n0 exp(px) (5.1)
x
where
r
1
p = 1+
20 n0 G0 2
v p
x = 20 n0 G0
G0
n(0) = n0 is the number density of nuclei in the product stream. The constants
and n0 were set to unity and three simulations were performed with the 0
(size-independent aggregation kernel) and G0 (growth rate constant) values given
in Figure 5.2. This figure clearly shows the effects of aggregation and growth to

Figure 5.2. FEM predictions of number density function (n(v)) for aggregation
and growth [99].

the number density of particles.


The experimental results of CSD at steady state were analyzed using Hounslows
constant aggregation rate equation (Equation 5.1) developed to solve the population
balance equation that accounts for nucleation, growth, and aggregation with a
156

constant aggregation rate kennel and a constant crystal growth rate [99]. With
measured CSD in number fraction (nL ), total mass of particles (MT ) was obtained
as Z
nL
MT = L3 dL (5.2)
0 100
where = 0.471 is shape factor for octahedron, = 2.170 g/cm3 is the density of
NaCl. Then the mass of particles in size L is given as

mvL
ML = (5.3)
MT

where vL is total volume of particles in size L and m is the mass of particles per
unit suspension volume. The number density, n, is given as

ML
n= (5.4)
(L3 ) L

To use Hounslows equation to our experimental results, a nonlinear regression


model is applied as

     
n L L
n(L) = 0  I1 2 B exp (A 1) (5.5)
L G G
B
G

With the nonlinear regression data fit adjusting initial variables, i.e., A, B, and
G (approximate G was obtained with the application of a MSMPR model), to the
experimental result of particle number density (N ), n0 , G, A and B are known, so
does B0 (= n0 G). Therefore, the aggregation rate constant is obtained.

B
Ka = (5.6)
G 2 n0

A typical nonlinear regression data fit is shown in Figures 5.3 and the resulting
coefficients for all the experimental runs for T samples are shown in Table 5.1. In
Figure 5.3, the little humps below 50 m in particle size were not used in the data fit.
Only the larger particle sizes where there is a clear indication of aggregation were
used. This analysis also does not account for breakage which is evident in Figure 5.3
due to lower trend in the data than the fit line for large particles. While this fitting
methodology does not account for all phenomena, i.e., breakage is missing, it will
157

Figure 5.3. A typical plot of nonlinear regression data fit to the particle number
density (N ) using the Hounslows constant aggregation rate equation at steady
state. The data points (2s) filled with s were used for the fit which is the solid
line. The particle population at zero size (n0 ) according to the fit is shown as a .

be used for further analysis since it is the only method available in the literature to
account for all possible phenomena in CMSMPR crystallizer. There is no analytical
solution to the population balance in a CMSMPR crystallizer to deal with the
nucleation, growth, aggregation, and breakage, altogether. Hounslows solution is
only for nucleation, growth, and aggregation, and does not account for breakage.
As a result, the results for the aggregation rate constant must be considered an
effective aggregation rate one that combines the dynamic balance of aggregation
with breakage.
With the application of Hounslows solution into the measured CSD at 10 for
all the experimental runs, the number of nuclei (n0 ), growth rate (G), nucleation
rate (B0 ), and aggregation rate (Ka ) were obtained and tabulated for each set of
designs in Tables 5.1, 5.2, and 5.3 which also give the standard error of estimate
indicated as a fitting error. The experimental conditions, such as crystallization
temperature (TC ), feed flow rate (Q), and mixing speed (M S), for each set of data
158

Table 5.1. Analytical data from CMSMPR crystallization


experiment (T samples)a,b .

Analytical Data
Test No. n0 G B0 Ka Fitting
(#/m4 ) (m/s) (#/m3 s) (m3 /s) error c
T 04 4.02E+13 8.21E-03 3.30E+05 1.03E-15 0.023
T 05 1.21E+14 6.43E-03 7.77E+05 4.38E-16 0.022
T 06 3.52E+14 5.84E-03 2.06E+06 1.65E-16 0.063
T 07 4.53E+13 1.10E-02 5.19E+05 1.47E-15 0.02
T 08 1.03E+14 9.02E-03 9.29E+05 7.69E-16 0.046
T 09 2.63E+14 5.11E-03 1.35E+06 4.93E-16 0.055
T 10 3.11E+13 4.47E-03 1.39E+05 6.12E-16 0.033
T 11 7.22E+13 3.60E-03 2.60E+05 3.27E-16 0.041
T 12 1.17E+14 3.19E-03 3.73E+05 2.28E-16 0.048
T 13 4.25E+13 8.30E-03 3.53E+05 9.64E-16 0.003
T 14 9.52E+13 6.68E-03 6.36E+05 5.35E-16 0.022
T 15 2.44E+14 5.63E-03 1.37E+06 2.48E-16 0.045
T 16 3.99E+13 1.30E-02 4.99E+05 1.53E-15 0.026
T 17 9.71E+13 9.93E-03 9.63E+05 7.94E-16 0.033
T 18 2.39E+14 8.68E-03 2.07E+06 3.69E-16 0.037
T 19 3.09E+13 4.59E-03 1.42E+05 5.99E-16 0.058
T 20 4.72E+13 4.04E-03 1.91E+05 4.46E-16 0.07
T 21 8.72E+13 3.42E-03 2.98E+05 2.85E-16 0.053
T 22 3.54E+13 8.49E-03 3.00E+05 1.13E-15 0.038
T 23 7.67E+13 7.07E-03 5.42E+05 6.28E-16 0.045
T 24 1.45E+14 6.07E-03 8.81E+05 3.86E-16 0.054
T 25 3.60E+13 1.30E-02 4.52E+05 1.69E-15 0.036
T 26 9.34E+13 1.00E-02 9.38E+05 8.16E-16 0.019
T 27 2.04E+14 8.09E-03 1.65E+06 4.63E-16 0.023

a
Constant experimental parameters: feed solution concentration
(C01 ) = CaCl2 (47.631.76 wt%), KCl (1.070.16 wt%), NaCl (0.670.08
wt%) with 95% confidence level, feed solution temperature (TS ) = 70 C
or above, mixing design = standard baffle + turbine.
b
For experimental condition of each test no., see Table 4.4 in Sec-
tion 4.3.3.
c
The analytical data with fitting error bigger than 0.07 was discarded
in this set of experiments.

were tabulated previously in Tables 4.4, 4.5, and 4.6 in Section 4.3.3. As shown in
Figure 5.3, the aggregation of particles was noticed in size range less than 50 m
159

Table 5.2. Analytical data from CMSMPR crystallization


experiment (TS samples)a,b .

Analytical Data
Test No. n0 G B0 Ka Fitting
(#/m4 ) (m/s) (#/m3 s) (m3 /s) error c
TS 04 4.87E+13 7.24E-03 3.53E+05 9.64E-16 0.021
TS 05 1.18E+14 6.05E-03 7.16E+05 4.75E-16 0.039
TS 06 2.95E+14 4.87E-03 1.43E+06 2.37E-16 0.071
TS 07 5.56E+13 1.10E-02 6.04E+05 1.27E-15 0.022
TS 08 1.21E+14 9.08E-03 1.10E+06 6.99E-16 0.006
TS 09 3.38E+14 7.08E-03 2.39E+06 3.20E-16 0.085
TS 10 2.18E+13 4.51E-03 9.81E+04 8.67E-16 0.025
TS 11 4.56E+13 3.75E-03 1.71E+05 4.98E-16 0.051
TS 12 9.32E+13 3.19E-03 2.97E+05 2.86E-16 0.048
TS 13 3.72E+13 7.74E-03 2.88E+05 1.18E-15 0.026
TS 14 6.20E+13 7.04E-03 4.36E+05 7.80E-16 0.023
TS 15 1.34E+14 5.88E-03 7.87E+05 4.32E-16 0.041
TS 16 4.75E+13 1.10E-02 5.39E+05 1.42E-15 0.02
TS 17 1.12E+14 9.26E-03 1.04E+06 7.39E-16 0.033
TS 18 2.29E+14 7.73E-03 1.77E+06 4.32E-16 0.059
TS 19 1.59E+13 4.70E-03 7.46E+04 1.14E-15 0.036
TS 20 2.92E+13 4.21E-03 1.23E+05 6.92E-16 0.051
TS 21 5.79E+13 3.60E-03 2.09E+05 4.08E-16 0.067
TS 22 2.43E+13 8.64E-03 2.10E+05 1.62E-15 0.034
TS 23 5.27E+13 7.35E-03 3.88E+05 8.77E-16 0.019
TS 24 1.09E+14 6.17E-03 6.74E+05 5.05E-16 0.038
TS 25 4.78E+13 1.10E-02 5.40E+05 1.42E-15 0.018
TS 26 9.03E+13 9.68E-03 8.75E+05 8.75E-16 0.034
TS 27 1.58E+14 8.44E-03 1.34E+06 5.73E-16 0.033

a
Constant experimental parameters: feed solution concentration
(C02 ) = CaCl2 (46.881.73 wt%), KCl (0.930.15 wt%), NaCl (0.500.05
wt%) with 95% confidence level, feed solution temperature (TS ) = 70 C
or above, mixing design = standard baffle + turbine.
b
For experimental condition of each test no., see Table 4.5 in Sec-
tion 4.3.3.
c
The analytical data with fitting error bigger than 0.085 was discarded
in this set of experiments.

in most of crystallization experiments, thus making the line fitting unsuccessful in


the size range less than 50 m.
160

Table 5.3. Analytical data from CMSMPR crystallization


experiment (IS samples)a,b .

Analytical Data
Test No. n0 G B0 Ka Fitting
(#/m4 ) (m/s) (#/m3 s) (m3 /s) error c
IS 04 5.36E+13 6.80E-03 3.64E+05 9.34E-16 0.018
IS 05 1.48E+14 5.56E-03 8.22E+05 4.14E-16 0.032
IS 06 4.55E+14 4.21E-03 1.92E+06 1.77E-16 0.05
IS 07 5.96E+13 9.90E-03 5.90E+05 1.30E-15 0.016
IS 08 1.24E+14 8.28E-03 1.03E+06 7.44E-16 0.025
IS 09 2.60E+14 6.79E-03 1.76E+06 4.34E-16 0.041
IS 10 4.38E+13 3.81E-03 1.67E+05 5.09E-16 0.025
IS 11 9.63E+13 3.11E-03 2.99E+05 2.84E-16 0.019
IS 12 1.78E+14 2.72E-03 4.86E+05 1.75E-16 0.019
IS 13 3.23E+13 8.05E-03 2.60E+05 1.31E-15 0.002
IS 14 6.53E+13 6.85E-03 4.47E+05 7.61E-16 0.017
IS 15 1.45E+14 5.56E-03 8.08E+05 4.21E-16 0.023
IS 16 5.05E+13 1.10E-02 5.44E+05 1.41E-15 0.024
IS 17 1.01E+14 8.75E-03 8.83E+05 8.67E-16 0.02
IS 18 2.69E+14 7.12E-03 1.92E+06 3.99E-16 0.025
IS 19 2.53E+13 4.26E-03 1.08E+05 7.91E-16 0.047
IS 20 5.89E+13 3.44E-03 2.02E+05 4.20E-16 0.056
IS 21 1.19E+14 2.93E-03 3.48E+05 2.44E-16 0.033
IS 22 2.13E+13 8.54E-03 1.82E+05 1.87E-15 0.018
IS 23 5.24E+13 7.02E-03 3.68E+05 9.24E-16 0.021
IS 24 1.05E+14 5.91E-03 6.18E+05 5.50E-16 0.021
IS 25 3.98E+13 1.10E-02 4.52E+05 1.69E-15 0.023
IS 26 8.65E+13 9.31E-03 8.05E+05 9.51E-16 0.023
IS 27 3.31E+14 6.70E-03 2.22E+06 3.45E-16 0.061

a
Constant experimental parameters: feed solution concentration
(C03 ) = CaCl2 (47.091.74 wt%), KCl (1.030.15 wt%), NaCl (0.440.04
wt%), feed solution temperature (Ts ) = 70 C or above, mixing design =
standard baffle + impeller.
b
For experimental condition of each test no., see Table 4.6 in Sec-
tion 4.3.3.
c
The analytical data with fitting error bigger than 0.061 was discarded
in this set of experiments.

5.3 Statistical Experimental Analysis


The purpose of this section is to show the relative importance of indepen-
dent variables, i.e., crystallization temperature (TC ), feed flow rate (Q), and mix-
161

ing speed (M S), affecting a crystallization process, as well as the importance


of their interactions to the research experiments according to statistical design
principles [100]. The analytical data obtained with the application of Hounslows
constant rate equation, which was discussed previously, will be analyzed following
the statistical design of experiments in Section 2.7.2.
The linear regression analysis with the least square method to fit a line through
a set of observations was used to analyze how a single dependent variable was
affected by the values of each independent variable. The result for T samples is
tabulated in Table 5.4. The linear regression was chosen because all independent
variables show high degree of linearity to dependent variables for all experiments.
T statistics provides some clue regarding the relative importance of each variable in
the model. For linear regression models, R is the correlation between the observed
2
and predicted values of the dependent variable. Rad is the square of this correlation
and adjusted by the number of variables and the sample size in the model to
compensate for the optimistic bias of R2 . The F statistics is used to test the
hypothesis that the slope (1 ) is 0 (or for multiple linear regression, that 1 , . . .
P = 0). F is large when the independent variable helps to explain the variation in
the dependent variable. Here, the linear relations of TC to the dependent variables
are not so significant (i.e., the Fcrit values for the F are bigger than 0.0005) shown
in Table 5.4. The linear regression analysis showing the relative linear fit to data
points means that the relationship between independent and dependent variables,
i.e., mixing speed (M S) on n0 , feed flow rate (Q) on G, feed flow rate (Q) and
mixing speed (M S) on B0 , and mixing speed (M S) on Ka for all experimental
designs, can be properly explained by this model. Other independent variables
2
with low Rad values showing broad distribution or poor relationship to dependent
variables might not be the best choice for describing or controlling the dependent
variables.
Table 5.5 shows the multiple linear regression analysis of two factor interaction
of independent variables to dependent variables for T samples. The interaction of
feed flow rate (Q) and mixing speed (M S) plays major role mostly on dependent
162

Table 5.4. Linear regression statistics of independent variables (TC , Q, and


M S) to dependent variables (n0 , G, B0 , and Ka ) for T samples (main effects)a .

Dependent Variable
Independent
n0
Variable
Reg. Coef.b tstat c 2 d
Rad F Fcrit
0 3.23E+141.36E+14 2.3674
TC 0.0597 2.4605 1.31E-01
1 -6.89E+124.39E+12 -1.5686
0 5.22E+135.13E+13 1.0164
Q 0.0203 1.4767 2.37E-01
1 1.38E+121.13E+12 1.2152
0 -2.15E+145.34E+13 -4.0286
MS 0.6223 38.8904 2.82E-06
1 8.53E+111.37E+11 6.2362
G
0 8.60E-034.65E-03 1.8469
TC -0.0414 0.0856 7.73E-01
1 -4.38E-051.50E-04 -0.2926
0 1.03E-039.80E-04 1.0507
Q 0.6598 45.6131 8.69E-07
1 1.46E-042.17E-05 6.7537
0 1.35E-022.53E-03 5.3310
MS 0.1897 6.3831 1.92E-02
1 -1.64E-056.49E-06 -2.5265
B0
0 1.98E+069.05E+05 2.1889
TC 0.0368 1.8783 1.84E-01
1 -4.00E+042.92E+04 -1.3705
0 -7.95E+042.92E+05 -0.2721
Q 0.2623 9.1774 6.16E-03
1 1.95E+046.45E+03 3.0294
0 -9.97E+054.41E+05 -2.2624
MS 0.4016 16.4331 5.29E-04
1 4.57E+031.13E+03 4.0538
Ka
0 6.96E-167.00E-16 0.9938
TC -0.0454 0.0003 9.87E-01
1 -3.68E-192.25E-17 -0.0163
0 1.25E-162.25E-16 0.5547
Q 0.2085 7.0583 1.44E-02
1 1.32E-174.96E-18 2.6567
0 2.26E-152.64E-16 8.5575
MS 0.6109 37.1131 3.93E-06
1 -4.11E-186.75E-19 -6.0921

a
Regression statistics was performed with confidence level of 95%.
b
Regression model with one independent variable: y = 0 + 1 x1 .
c
T statistics provide the relative importance of each variable in the model.
d 2
Rad : R is the correlation between the observed and predicted values of the dependent
2
variable. Rad is the square of this correlation and adjusted by the number of variables in the
model and the sample size to compensate for the optimistic bias of R2 .
163

variables (i.e., G, B0 , and Ka ) followed by the interaction of M S TC , except


for the growth rate (G) in which the interaction of TC Q is the second. The

Table 5.5. Multiple linear regression statistics of independent variables (TC , Q,


and M S) to dependent variables (n0 , G, B0 , and Ka ) for T samples (two-factor
interactions)a .

Dependent Variable
Independent
n0
Variable b
Reg. Coeff. tstat c 2 d
Rad F Fcrit
0 2.50E+141.61E+14 1.5534
TC Q 1 -5.90E+124.56E+12 -1.2945 0.0495 1.5989 2.26E-01
2 1.01E+121.15E+12 0.8739
0 -2.79E+145.66E+13 -4.9289
Q MS 1 1.46E+126.47E+11 2.2582 0.6816 25.6185 2.32E-06
2 8.58E+111.26E+11 6.8332
0 -1.17E+128.74E+13 -0.0134
M S TC 1 8.54E+111.18E+11 7.2122 0.7167 30.0910 6.82E-07
2 -6.96E+122.41E+12 -2.8864
G
0 -2.53E-033.08E-03 -0.8187
TC Q 1 1.06E-048.75E-05 1.2143 0.6670 24.0356 3.72E-06
2 1.53E-042.21E-05 6.9140
0 7.16E-031.15E-03 6.2123
Q MS 1 1.45E-041.32E-05 10.9983 0.8744 81.0756 1.33E-10
2 -1.59E-052.56E-06 -6.2124
0 1.48E-024.89E-03 3.0307
M S TC 1 -1.64E-056.63E-06 -2.4729 0.1551 3.1105 6.56E-02
2 -4.25E-051.35E-04 -0.3149
Table continues on next page.

a
Regression statistics was performed with confidence level of 95%.
b
Regression model with two independent variables: y = 0 + 1 x1 + 2 x2 + .
c
T statistics provide the relative importance of each variable in the model.
d 2
Rad : R is the correlation between the observed and predicted values of the dependent
2
variable. Rad is the square of this correlation and adjusted by the number of variables in the
model and the sample size to compensate for the optimistic bias of R2 .

interaction of M S TC was the first on n0 followed by Q M S. The three factor


interaction of independent variables to dependent variables was nicely described by
the multiple linear regression model which predicts the interactions with over 70 to
164

Continued from Table 5.5a .


Dependent Variable
Independent
B0
Variable b
Reg. Coeff. tstat c 2 d
Rad F Fcrit
0 6.62E+059.35E+05 0.7075
TC Q 1 -2.21E+042.65E+04 -0.8348 0.2520 4.8740 1.82E-02
2 1.82E+046.71E+03 2.7073
0 -1.87E+063.54E+05 -5.2892
Q MS 1 2.00E+044.04E+03 4.9423 0.7102 29.1790 8.66E-07
2 4.65E+037.85E+02 5.9159
0 2.43E+057.90E+05 0.3070
M S TC 1 4.58E+031.07E+03 4.2781 0.4608 10.8289 5.87E-04
2 -4.03E+042.18E+04 -1.8489
Ka
0 -3.23E-167.23E-16 -0.4472
TC Q 1 1.34E-172.05E-17 0.6524 0.1873 3.6498 4.36E-02
2 1.40E-175.19E-18 2.7017
0 1.70E-151.94E-16 8.7567
Q MS 1 1.28E-172.21E-18 5.7853 0.8429 62.6786 1.40E-09
2 -4.07E-184.29E-19 -9.4766
0 2.26E-155.10E-16 4.4276
M S TC 1 -4.11E-186.91E-19 -5.9519 0.5924 17.7131 3.11E-05
2 -3.34E-201.41E-17 -0.0024

a
Regression statistics was performed with confidence level of 95%.
b
Regression model with two independent variables: y = 0 + 1 x1 + 2 x2 + .
c
T statistics provide the relative importance of each variable in the model.
d 2
Rad : R is the correlation between the observed and predicted values of the dependent
2
variable. Rad is the square of this correlation and adjusted by the number of variables in the
model and the sample size to compensate for the optimistic bias of R2 .

90% accuracies in the case of T samples shown in Table 5.6. The linear regression
analyses of single, two, and three factor interaction were performed for all samples,
i.e., T, TS, and IS samples. However, for the illustration purpose only the result
of T samples is given.
The correlation analysis for each one of independent variables such as crystalliza-
tion temperature (TC ), feed flow rate (Q), and mixing speed (M S), to each depen-
dent variable was performed for T, TS, and IS samples as shown in Table 5.7. The
negative sign indicates that as an independent variable increases the corresponding
dependent variable tends to decrease. As shown in Table 5.7, this analysis presents
165

Table 5.6. Multiple linear regression statistics of independent variables (TC , Q,


and M S) to dependent variables (n0 , G, B0 , and Ka ) for T samples (three-factor
interactions)a .

Dependent Variable
Independent
n0
Variable b
Reg. Coeff. tstat c 2 d
Rad F Fcrit
0 -8.19E+139.39E+13 -0.8718
1 -5.89E+122.36E+12 -2.4945
TC Q M S 0.7450 2.34E+01 9.49E-07
2 1.09E+125.97E+11 1.8257
3 8.58E+111.12E+11 7.6342
G
0 3.61E-031.98E-03 1.8281
1 1.06E-044.97E-05 2.1337
TC Q M S 0.8926 6.47E+01 1.79E-10
2 1.51E-041.26E-05 12.0499
3 -1.59E-052.36E-06 -6.7160
B0
0 -1.13E+066.43E+05 -1.7631
1 -2.21E+041.62E+04 -1.3646
TC Q M S 0.7216 2.09E+01 2.26E-06
2 1.86E+044.09E+03 4.5465
3 4.65E+037.70E+02 6.0353
Ka
0 1.25E-153.48E-16 3.5888
1 1.33E-178.75E-18 1.5234
TC Q M S 0.8521 4.52E+01 4.29E-09
2 1.36E-172.21E-18 6.1559
3 -4.07E-184.16E-19 -9.7691

a
Regression statistics was performed with confidence level of 95%.
b
Regression model with three independent variables: y = 0 + 1 x1 + 2 x2 + 3 x3 + .
c
T statistics provide the relative importance of each variable in the model.
d 2
Rad : R is the correlation between the observed and predicted values of the dependent variable.
2
Rad is the square of this correlation and adjusted by the number of variables in the model and
the sample size to compensate for the optimistic bias of R2 .

the dominant interactions (the number within a box) of feed flow rate (Q) and
mixing speed (M S) on resultant product data in all designs, also shows the weak
interactions of crystallization temperature (TC ). The strong interaction of mixing
speed (M S) and weak interaction of crystallization temperature (TC ) on number of
nuclei (n0 ), nucleation rate (B0 ), and aggregation rage (Ka ) could be explained by
secondary nucleation which occurs in the presence of a solute-particle interface or
166

Table 5.7. Correlation analysis of independent variables to dependent variables


for T, TS, and IS samples.

Independent Variables Dependent Variables


Sample
TC a Qb MSc n0 G B0 Ka
n0 -0.2900 0.3237 0.8179 1
G -0.1913 0.8030 -0.3631 -0.1581 1
T
B0 -0.2710 0.5769 0.7212 0.9228 0.1541 1
Ka -0.1396 0.4049 -0.7782 -0.5966 0.7837 -0.4059 1
n0 -0.3355 0.4970 0.7278 1
G -0.0464 0.8848 -0.3187 0.0622 1
TS
B0 -0.3064 0.6987 0.5994 0.9592 0.3090 1
Ka 0.2429 0.1532 -0.8721 -0.6284 0.5753 -0.4562 1
n0 -0.1691 0.3191 0.7939 1
G -0.1113 0.8623 -0.4180 -0.1667 1
IS
B0 -0.2058 0.6401 0.6017 0.9158 0.1928 1
Ka 0.1066 0.4011 -0.7554 -0.6104 0.7919 -0.3353 1

a
TC : Crystallization temperature ( C).
b
Q: Flow rate (ml/min).
c
MS: Mixing speed (rpm).

causes mostly by collisions of crystals with other crystals, crystallizer, and mixer.
Once supersaturation creates solid particles by primary nucleation in a CMSMPR
crystallizer, the nucleation rate (B0 ) mostly depends on stirrer rotation rate (rpm)
in the presence of such solid particles. This is called secondary or contact nucle-
ation as discussed in Section 1.3.3. Compared to the strong contribution of mixer
rpm, the effect of supersaturation on secondary nucleation was weak as shown in
Equation 1.36, thus explaining the weak interaction of crystallization temperature
(TC ) and the strong interaction of mixer rpm on products (dependent variables).
The large correlation of dependent variables with mixing speed indicates that the
secondary nucleation and growth play the dominant crystallization mechanism for
this system, as most industrial crystallizer does. In addition to the large number of
small particles created by secondary nucleation and growth, particle breakage and
aggregation were certainly resulted in deviation of CSD from linearity as indicated
167

in Figure 5.3 in previous section. The humps in small particle size range (below 50
m) were the indication of aggregation among these small particles and relatively
large particles. The deviation of CSD from the linear line in large particle size
region (above 230 m) was the result of large particle breakage. The growth rate
(G) shows relatively weak correlation with mixing speed due to such dominant
particle interaction. The flow rate effect, i.e., Q = V / where = mean residence
time, shows strong correlation with growth rate (G) and nucleation rate (B0 ) which
induced mostly by supersaturation. Also the bold numbers in Table 5.7 indicate
rather strong correlations between dependent variables, i.e., B0 and n0 in B0 = n0 G.
The aggregation rate (Ka ) shows relatively strong correlation with the number of
nuclei (n0 ), growth rate (G), and nucleation rate (n0 ) indicating the importance of
secondary nucleation and the crystallization mechanism involving growth, breakage,
and aggregation together.
The individual and interaction effects of independent variables, i.e., crystalliza-
tion temperature (TC ), feed flow rate (Q), and mixing speed (M S), on each one
of dependent variables, i.e., number of nuclei (n0 ), growth rate (G), nucleation
rate (B0 ), and aggregation rate (Ka ), were attempted to compare T, TS, and IS
sample designs to show the degree of importance among processing variables to the
differences in product variables due to the difference in designs which were shown
previously in Tables 2.2, 2.3, and 2.4 in Section 2.7.2. Tables 5.8, 5.9, and 5.10
present the results obtained with the application of factorial design analysis. In
this analysis however, all the observations should be used to supply information on
each of the individual and interaction effects. However, as explained in Section 4.3.3,
the crystallization experiments at 25 C with 20 ml/min were not carried out for all
experimental designs. Therefore, the measures of individual and interaction effects
were performed disregarding all of experimental results at 25 C. The regression and
correlation analyses showed previously that the crystallization temperature effect
on the dependent variables was small. This leads the choice made for this factorial
design analysis was reasonable to show the individual and interaction effects of
independent variables focusing the effects of feed flow rate and mixing speed on
168

product variables. In Tables 5.8, 5.9, and 5.10, the main effects of crystallization
temperature (TC ), feed flow rate (Q), and mixing speed (M S) are denoted by A,
B, and C, respectively. The two or three-factor interactions are indicated as AB,
BC, and CA, or ABC. Figures 5.4, 5.5, and 5.6 illustrate the effects of individuals
and interactions on growth rate (G), nucleation rate (B0 ), and aggregation rate
(Ka ), respectively for T, TS, and IS samples. The effect on number of nuclei (n0 )
is omitted here since the relation, i.e., B0 = n0 G, is known. Figure 5.4 clearly

Figure 5.4. The order of importance in main and interaction effects on growth
rate (G). Crystallization temperature (TC ), feed flow rate (Q), and mixing speed
(M S) are denoted by A, B, and C, respectively. Dotted lines indicate the standard
error of estimate.

shows the importance of main and interaction effects on growth rate (G). The
feed flow rate (Q) provides the largest positive effect on growth rate. The main
effect of mixing speed (M S) and the two-factor interaction effect (BC) of feed
flow rate and mixing speed show the second and the third effects as negative to
growth rate. All other effects are within the standard error of estimation. The
difference in feed solution concentration between T and TS samples (C01 and C02 )
169

clearly stands out in all three figures with lower effects of flow rate and mixing
speed on growth, nucleation, and aggregation rates for TS samples as a result of
lower NaCl concentration in C02 compared to C01 . Figures 5.5 and 5.6 present the
effects on nucleation rate (B0 ) and aggregation (Ka ), respectively. The main and

Figure 5.5. The order of importance in main and interaction effects on nucleation
rate (B0 ). Crystallization temperature (TC ), feed flow rate (Q), and mixing speed
(M S) are denoted by A, B, and C, respectively. Dotted lines indicate the standard
error of estimate.

interaction effects of feed flow rate and mixing speed on nucleation rate explain
the importance of secondary nucleation caused by high supersaturation and high
turbulence which easily washes nuclei away from the existing crystal surface in
solution. By comparing three figures, the effects on IS samples show almost same
trends as on T and TS samples. The effects on growth and aggregation rate were
almost comparable to T sample. The effects on nucleation rate was even higher than
the T and TS samples even though the concentration (C03 ) of IS sample was the
lowest among the designs and the speed of mixing was higher compare to T and TS
samples. The mixing pattern due to the different type of mixer, i.e., an impeller,
170

was single axial mixing rather than bi-axial mixing. In Figure 5.6, the effect of

Figure 5.6. The order of importance in main and interaction effects on aggregation
rate (Ka ). Crystallization temperature (TC ), feed flow rate (Q), and mixing speed
(M S) are denoted by A, B, and C, respectively. Dotted lines indicate the standard
error of estimate.

mixing shows negative effect on aggregation rate meaning that the mixing plays an
important role on secondary nucleation which is clearly indicated in Figure 5.5.
Table 5.8. Factorial design analysis of independent variables (TC , Q, and M S) to dependent
variables (n0 , G, B0 , and Ka ) for T samplesa,b .

n0 G
Interaction
Reg. Coef. F Fcrit S.E.E.c Reg. Coef. F Fcrit S.E.E.
A -9.21E+12 15.16 12.25 -3.03E-05 0 12.25
B 3.42E+13 208.52 12.25 3.52E-03 3784.61 12.25
C 6.88E+13 845.91 12.25 -1.52E-03 702.61 12.25
AB -1.12E+12 0 12.25 4.73E+12 -1.23E-04 4.58 12.25 1.15E-04
BC 3.04E+13 165.16 12.25 -8.80E-04 236.24 12.25
CA -8.12E+12 11.77 12.25 -6.25E-05 1.19 12.25
ABC -1.42E+11 0.00 12.25 -9.03E-05 2.49 12.25
B0 Ka
Interaction
Reg. Coef. F Fcrit S.E.E. Reg. Coef. F Fcrit S.E.E.
A -7.23E+04 5.09 10.04 3.93E-17 3.59 10.04
B 5.02E+05 245.56 10.04 3.10E-16 222.70 10.04
C 4.26E+05 177.20 10.04 -4.11E-16 391.13 10.04
AB -5.30E+04 0 10.04 6.41E+04 2.75E-17 0 10.04 4.16E-17
BC 3.21E+05 100.56 10.04 -2.25E-16 117.57 10.04
CA -6.05E+04 3.57 10.04 6.57E-19 0 10.04
ABC -3.96E+04 0 10.04 -1.79E-17 0 10.04

a
Regression statistics was performed with confidence level of 95%.
b
Crystallization temperature (TC ), feed flow rate (Q), and mixing speed (M S) are denoted by A, B, and
C, respectively.
c
S.E.E.: Standard error of estimation.
171
Table 5.9. Factorial design analysis of independent variables (TC , Q, and M S) to dependent
variables (n0 , G, B0 , and Ka ) for TS samplesa,b .

n0 G
Interaction
Reg. Coef. F Fcrit S.E.E.c Reg. Coef. F Fcrit S.E.E.
A -1.52E+13 43.51 12.25 1.71E-04 18.70 12.25
B 3.99E+13 299.80 12.25 2.87E-03 5239.55 12.25
C 5.51E+13 571.58 12.25 -1.07E-03 726.56 12.25
AB -4.01E+12 3.03 12.25 4.61E+12 1.42E-05 0 12.25 7.92E-05
BC 2.43E+13 110.75 12.25 -4.42E-04 124.36 12.25
CA -1.37E+13 35.24 12.25 1.21E-04 9.31 12.25
ABC -5.70E+12 6.11 12.25 6.47E-05 2.67 12.25
B0 Ka
Interaction
Reg. Coef. F Fcrit S.E.E. Reg. Coef. F Fcrit S.E.E.
A -7.42E+04 15.32 12.25 7.00E-17 69.55 12.25
B 4.76E+05 630.69 12.25 1.50E-16 320.71 12.25
C 3.20E+05 285.91 12.25 -4.13E-16 2424.23 12.25
AB -4.37E+04 5.31 12.25 3.79E+04 -3.37E-17 16.10 12.25 1.68E-17
BC 2.30E+05 147.20 12.25 -6.83E-17 66.16 12.25
CA -6.81E+04 12.91 12.25 -9.79E-19 0 12.25
ABC -5.04E+04 7.06 12.25 3.88E-17 21.35 12.25

a
Regression statistics was performed with confidence level of 95%.
b
Crystallization temperature (TC ), feed flow rate (Q), and mixing speed (M S) are denoted by A, B, and
C, respectively.
c
S.E.E.: Standard error of estimation.
172
Table 5.10. Factorial design analysis of independent variables (TC , Q, and M S) to dependent
variables (n0 , G, B0 , and Ka ) for IS samplesa,b .

n0 G
Interaction
Reg. Coef. F Fcrit S.E.E.c Reg. Coef. F Fcrit S.E.E.
A -3.69E+12 1.22 12.25 3.09E-05 0 12.25
B 4.39E+13 172.68 12.25 2.87E-03 3816.85 12.25
C 9.97E+13 892.45 12.25 -1.38E-03 877.88 12.25
AB 1.74E+13 27.30 12.25 6.68E+12 -1.39E-04 9.02 12.25 9.29E-05
BC 3.81E+13 130.46 12.25 -7.49E-04 260.41 12.25
CA 4.23E+12 1.60 12.25 -8.49E-05 3.35 12.25
ABC 1.53E+13 21.04 12.25 -2.36E-05 0 12.25
B0 Ka
Interaction
Reg. Coef. F Fcrit S.E.E. Reg. Coef. F Fcrit S.E.E.
A 1.39E+03 0 12.25 7.75E-17 36.82 12.25
B 5.46E+05 477.79 12.25 2.83E-16 489.79 12.25
C 5.03E+05 404.08 12.25 -4.31E-16 1135.56 12.25
AB 5.49E+04 4.83 12.25 5.0E+04 -1.58E-17 1.53 12.25 2.56E-17
BC 3.51E+05 196.66 12.25 -1.96E-16 236.18 12.25
CA 4.25E+04 2.89 12.25 -7.35E-17 33.10 12.25
ABC 6.37E+04 6.49 12.25 -1.69E-17 1.75 12.25

a
Regression statistics was performed with confidence level of 95%.
b
Crystallization temperature (TC ), feed flow rate (Q), and mixing speed (M S) are denoted by A, B, and
C, respectively.
c
S.E.E.: Standard error of estimation.
173
174

5.3.1 Analysis of T samples


In connection to Figures 5.4 through 5.6 which showed previously the effects
of main, two-factor, and three factor interactions on analytical data obtained with
the application of Hounslows equation to T, TS, and IS samples, this section
discusses more details on T samples. The analytical data for T samples was shown
in Table 5.1 and the experimental conditions for each test number could be found
in Table 4.4 (see Section 4.3.3. Although the effect of crystallization temperature
(TC ) was small on analytical data as shown in previous analyses, the effect of
crystallization temperature was evident at high feed flow rate (Q) and high mixing
speed (M S). Figure 5.7 shows the trend that as the crystallization temperature
increases the number of nuclei (n0 ) and nucleation rate (B0 ) were decreased linearly.
Growth rate (G) and aggregation rate (Ka ) were increased slightly. However,
as shown in the Figure 5.7 the temperature effect was more evident at higher
mixing speed (Figure 5.7 (b)) and at higher flow rate. Although not shown here,
as the feed flow rate increases (i.e., short mean residence time), the number of
nuclei, growth rate, nucleation rate, and aggregation rate all together increase with
different magnitude. The main effect of feed flow rate was more evident at low
crystallization temperature and high mixing speed. As shown in Figure 5.7 when
the mixing speed increased from 285 to 481 rpm the number of nuclei (9-fold) and
the nucleation rate (5-fold) are increased, and the growth and aggregation rates are
decreased at 25 C with respect to mixer rpm. The main effect of mixing speed was
evident at low crystallization temperature and low feed flow rate.
Among the three variables, i.e., crystallization temperature, flow rate, and
mixing speed, the speed of mixing has clearly demonstrated the most dominant
effect on n0 and Ka , and the feed flow rate on G and B0 . In Figure 5.8, the
relationship between n0 and B0 as indicated in B0 = n0 G is plotted to show the
effect of feed flow rate and mixing speed. The effect of crystallization temperature
was insignificant compare to other two variables as shown in Figure 5.8. Previously,
the fact that the correlation and factorial analyses showed the weak correlation
of crystallization temperature to the nucleation rate and the number of nuclei is
175

(a)

(b)

Figure 5.7. Influence of crystallization temperature and mixing speed only (main
effect) on number of nuclei(n0 ), growth rate (G), and nucleation rate (B0 ), as well
as aggregation rate (Ka ) at steady state for T samples. (a) T 04, T 13, and T 22
samples at 285 rpm and (b) T 06, T 15, and T 24 samples at 481 rpm (see Tables 4.4
and 5.1 for experimental conditions and analytical data). The feed flow rate was
40 ml/min for both (a) and (b).
176

proved here again in Figure 5.8. As the strong correlation (0.9228) between n0 and

Figure 5.8. Nucleation rate (B0 ) vs. number of nuclei(n0 ) at steady state for T
samples (see Tables 4.4 and 5.1 for experimental conditions and analytical data).

B0 was indicated in Table 5.7, the correlation along with the interaction of mixing
speed, in addition the main interaction of flow rate on n0 and B0 are shown in
Figure 5.8. This plot shows as mixing speed increases from 283 to 385 rpm and to
483 rpm the effect of mixing on n0 and B0 was increased linearly as 2 and 4-fold
magnitude, respectively. With respect to increase in feed flow rate, the effect on
B0 tends to increase more while the effect on n0 increases less. This is showing the
relation of mean residence time on both of n0 and B0 meaning that the short mean
residence time creates high supersaturation resulting in high nucleation and growth
rate meanwhile particles spending less time in crystallizer reducing the effect of feed
flow rate on n0 meaning less secondary nucleation. In other words, lower feed flow
rate creates easy of relaxation for supersaturation resulting in lower nucleation rate,
but particles spending more time in crystallizer will experience more interactions,
i.e., particle-particle, particle-mixer, and particle-baffle-crystallizer wall. As a result
177

of those interactions more nuclei were created at low feed flow rate. The feed flow
rate along with mixing speed having major effect on the growth rate shows increase
in slope of line in the order of 3, 5, and 8 as feed flow rate increases in the order of
20, 40, and 60 ml/min.
The power law curve fit of particle mean size (Lmean ) to nucleation rate (B0 )
shown in Figure 5.9 presents the relation as higher nucleation rate mostly caused
by high supersaturation creating large number of nuclei results in smaller size of
particles or vice versa.

Figure 5.9. Nucleation rate (B0 ) vs. particle mean size (Lmean ) at steady state
for T samples (see Tables 4.4 and 5.1 for experimental conditions and analytical
data).

5.3.2 Analysis of TS samples


The only difference in design of T and TS samples was the feed solution con-
centration as indicated in Section 2.7.2. The experimental and analytical data for
TS samples were shown previously as in Tables 4.5 and 5.2, respectively. As shown
previously in Figure 5.7 for T samples, the effect of crystallization temperature
178

(TC ) showed the linear decrease in number of nuclei (n0 ) and nucleation rate
(B0 ) while growth rate (G) and aggregation rate (Ka ) showed linear increase.
TS samples showed the same trends as T samples for the case of crystallization
temperature effect on dependent variables. The linearity showed almost same
order of magnitude in both samples. In both sample designs the interaction of
crystallization temperature to the product data was weak. In T samples, mixing
speed effect was significant on n0 , B0 , and Ka while feed flow rate was interacting
strongly with growth rate (G). However in TS samples, the nucleation rate (B0 )
interacted more with feed flow rate instead of interacting with mixing speed. Also
in both designs, the correlation of product analytical data, i.e., between dependent
variables, showed somewhat different results. The interaction between n0 and B0
was significant for both designs. Meanwhile the interaction between Ka and G was
significant in T samples, Ka correlated more with n0 in TS samples.
When comparing the main effects of feed flow rate and mixing speed only to
the analytical product data shown in Figure 5.10 plotted as a function of feed flow
rate, all dependent variables tend to show increase in values as the feed flow rate
increases. However, the difference was minor when compared to T samples. In both
T and TS samples, the increase in growth and aggregation rate was small compared
to the number of nuclei and nucleation rate as the flow rate increases.
For the comparison of three-factor interaction on the number of nuclei (n0 ) and
nucleation rate (B0 ), the effects of crystallization temperature (TC ), feed flow rate
(Q), and mixing speed (M S) all together were plotted in Figure 5.11. Figure 5.11
shows the same pattern that the growth rate was almost constantly decreased with
respect to increase in mixing speed and decrease in crystallization temperature at
same feed flow rate when compared to Figure 5.8. However, there are two little
differences that can be pointed out between these two figures. The experimental
conditions for T and TS samples were almost same except for only a small difference
in feed solution concentration (C01 and C02 , see Appendix C). First, the slope of
linear line to 60 ml/min flow rate was smaller by 2 orders of magnitude when
compared to Figure 5.9 meaning that the growth rate was smaller in TS samples
179

(a)

(b)

Figure 5.10. Influence of feed flow rate and mixing speed only (main effect)
on number of nuclei(n0 ), growth rate (G), and nucleation rate (B0 ), as well as
aggregation rate (Ka ) at steady state for TS samples. (a) TS 10, TS 13, and TS
16 samples at 282 rpm and (b) TS 12, TS 15, and TS 18 samples at 479 rpm (see
Tables 4.5 and 5.2 for experimental conditions and results). The crystallization
temperature was 30 C for both (a) and (b).

than T samples at feed flow rate of 60 ml/min. Second, the mixing effect at feed
flow rate of 40 ml/min in TS samples was small compared to T samples. The growth
180

Figure 5.11. Nucleation rate (B0 ) vs. number of nuclei(n0 ) at steady state for TS
samples (see Tables 4.5 and 5.2 for experimental conditions and analytical data).

rate (G) at feed flow rate of 20 ml/min was almost same for both T and TS samples
due to the major effect of mixing at low feed flow rate. T samples with C01 having
slightly higher concentration than TS samples in all three components, i.e., CaCl2 ,
KCl, and NaCl, produced more particles which in turn created a suitable condition
for the particles to be exposed more to the effects of mixing speed (M S) and feed
flow rate (Q). As was the case with the T samples, the crystallization temperature
effect to the products in TS sample was small when compared to the effects of mixing
speed and feed flow rate. To see overall effect on growth rate, Figure 5.12 is plotted
B0 vs. 1/n0 . The growth rate within the same feed flow rate follows power law
relationship due to the effect from mixing speed and crystallization temperature.
This called growth rate dispersion which is more evident at low feed flow rate and
high mixing speed.
The correlation analysis at each of constant crystallization temperature, con-
stant feed flow rate, and constant mixing speed was performed to explore the
details of main effects to the product data as shown in Table 5.11. This is to
181

Figure 5.12. Nucleation rate (B0 ) vs. reciprocal of number of nuclei(n0 ) at


steady state for T samples (see Tables 4.5 and 5.2 for experimental conditions
and analytical data).

show the relationship between two variables while one is set constant. The crys-
tallization temperature interaction was small when compared with mixing speed,
i.e., the interactions of mixing speed with dependent variables were dominant over
the crystallization temperature in all cases. When compared the correlations of
crystallization temperature (TC ) and feed flow rate (Q) to the dependent variables,
Q was mostly dominant over TC at constant mixing speed, however, as mixing
speed increases from 281.8 to 479.0 rpm the interaction of TC to the products
is increased. At 479.0 rpm, the interactions of TC to number of nuclei (n0 ) and
aggregation rate (Ka ) are dominant over Q. This could be attributed to more
particle breakage and less aggregation as mixer rpm increases. Also, high rpm will
induce uniformity of the temperature distribution in the crystallizer which in turn
will increase temperature gradient between feed and slurry in crystallizer, therefore
resulting in high nucleation rate (B0 ) and aggregation rate (Ka ) due to the effect
of both temperature and mixing. The correlations of feed flow rate (Q) and mixing
182

speed (M S) at constant crystallization temperature (TC ) show rather complicate


interaction on products. At 24.85 C the interaction of mixing speed is dominant
over feed flow rate on n0 , B0 , and Ka except for the feed flow rate dominating
G. As temperature increases the interaction of feed flow rate dominates G and
B0 more. The effect of feed flow rate on dependent variables are increasing as
the crystallization temperature increases except for the interaction of mixing speed
dominating Ka in all cases.
By the analysis shown in Table 5.11, it was found out that the effect of feed flow
rate (Q) dominates over crystallization temperature (TC ) at low mixing speed and
that mixing speed (M S) dominates over feed flow rate at low crystallization tem-
perature. However, as mixing speed increases the effect of temperature dominates
on number of nuclei and aggregation. Also, as crystallization temperature increases
the dominance of mixing speed on nucleation rate was replaced by the feed flow
rate. Overall, the effect of crystallization temperature was evident at 40 ml/min
and higher mixing speed. The effect of feed flow rate at higher crystallization
temperature and lower mixing speed was evident. The mixing speed effect was
dominant at low crystallization temperature and low feed flow rate.
As discussed previously and in this section it is worth knowing how the inde-
pendent variables are interacting together and alone onto the dependent variables
to find out the controlling variables for the purpose of crystallization at certain
conditions. For instance, if someone wants to control nucleation and growth rate
the key controlling variable should be the feed flow rate with some help from either
temperature or mixing speed. Or if someone wants to control the aggregation
rate the mixing speed would be the major controlling variable with crystallization
temperature and feed flow rate as supporting variables within the experimental
conditions presented.
To show the influence of crystallization temperature only (main effect) on num-
ber of nuclei(n0 ), growth rate (G), and nucleation rate (B0 ), as well as aggregation
rate (Ka ) at steady state, the test conditions in test pairs TS 04 TS 13 TS 22 are
similar with respect to feed flow rate and mixing speed, but different with respect
183

to crystallization temperature. Therefore, the difference in the results within each


of these pairs reflect the effect of temperature alone. Similarly, the test conditions
for test pairs TS 09 TS 18 TS 27 or so on can be tried to see the main effect of
crystallization temperature (see Tables 4.5 and 5.2 for experimental conditions and
analytical data). As a result of this analysis, the effect of crystallization temperature
(TC ) on analytical data shows most evident at low feed flow rate and at high
mixing speed in both T and TS samples. Although not shown the crystallization
temperature effect for TS samples, the same trend was observed as in Figure 5.7
for T samples. The main effect of crystallization temperature (TC ) in TS samples
showed little increase in growth rate (G) and aggregation rate (Ka ) while showing
rapid decrease in number of nuclei (n0 ) and nucleation rate (B0 ) with respect to
increase in crystallization temperature at high mixing speed. This is true for both
B0 and n0 since the nucleation, most probably in this case heterogeneous nucleation
and secondary nucleation, depends on temperature as B0Hetero exp (1/T ) and
B0secondary (S 1) or ln S (1/T ).
The effect of feed flow rate only on dependent variables at steady state showed
same trends with different order of magnitude for all of T and TS samples as shown
in Figure 5.10. The feed flow rate only on growth and aggregation rate was most
significant at low crystallization temperature and low speed of mixing while the
number of nuclei and nucleation rate showed significant effect of feed flow rate at
low crystallization temperature and higher mixing. Particles spending less time
in a crystallizer at higher flow rate (i.e., shorter residence time distribution) and
at higher level of solution supersaturation explains the increase in nucleation rate
(B0 ) and little increase in growth rate (G), in turn resulting in increased aggregation
rate.
It is interesting to point out that the growth and aggregation rate increased with
increasing both crystallization temperature and feed flow rate and that the number
of nuclei and nucleation rate decreased as crystallization temperature increased,
but increased with feed flow rate as shown in Figures 5.7 and 5.10, respectively.
The relationship between measured mean size of particles (Lmean ) and nucleation
184

rate (B0 ) also verifies further the interaction of variables on products as shown in
Figure 5.13.

Figure 5.13. Particle mean size (Lmean ) vs. nucleation rate (B0 ) at steady state
for TS samples (see Tables 4.5 and 5.2 for experimental conditions and results).

For the TS samples, the predictions of B0 , G, and Ka were obtained with


respect to independent variables by the multiple linear regression model; y = 0 +
1 x1 + 2 x2 + 3 x3 + where x1 , x2 , and x3 are crystallization temperature, feed
flow rate, and mixing speed, respectively. Figures 5.14, 5.15 and 5.16 present the
comparison between the predicted and measured values of dependent variables. The
multiple linear regression model of independent variables to dependent variables
was made with 95% confidence level for all cases. In Figure 5.14 the prediction of
model shows 91% accuracy. At low nucleation rate (B0 ), the predicted B0 tends
to show overestimates while at high nucleation rate shows underestimations than
the measured ones. This is showing the difficulty obtaining accuracy from the CSD
measurement and its analysis at such high rate of nucleation. Other reasons for
the discrepancy might be from the data fitting that as shown in Figure 5.3 particle
185

Figure 5.14. Predicted B0 vs. measured B0 at steady state for TS samples (see
Tables 4.5 and 5.2 for experimental conditions and results). Multiple regression
model of independent variables to B0 (i.e., y = 0 + 1 x1 + 2 x2 + 3 x3 + where
x1 , x2 , and x3 are crystallization temperature, feed flow rate, and mixing speed,
respectively.) was performed with 95% confidence level and showed 91% prediction
accuracy.

size smaller than 50 m cannot be used in the data fit and from the fitting errors
shown in Tables 5.1, 5.2, and 5.3 for T, TS, and IS samples. The predictions of G
and Ka to the measured show 95 and 96% accuracies, respectively, with confidence
level of 95% in Figures 5.15 and 5.16.
186

Figure 5.15. Predicted G vs. measured G at steady state for TS samples (see
Tables 4.5 and 5.2 for experimental conditions and results). Multiple regression of
independent variables to G (i.e., y = 0 + 1 x1 + 2 x2 + 3 x3 + where x1 , x2 , and
x3 are crystallization temperature, feed flow rate, and mixing speed, respectively.)
was performed with 95% confidence level and showed 99% prediction accuracy.
Table 5.11. Correlation analysis of independent variables to the dependent variables at each
of constant crystallization temperature (TC ), constant feed flow rate (Q), and constant mixing
speed (M S) for TS samples.

Variables Variables Variables


TC a Qb MSc
Q MS TC MS TC Q
n0 0.0776 0.9563 n0 -0.3391 0.8835 n0 -0.6023 0.8733
G 0.7503 -0.6421 G 0.3056 -0.9347 G -0.1028 0.9897
24.85 20 281.8
B0 0.3923 0.8680 B0 -0.3277 0.9056 B0 -0.3635 0.9725
Ka 0.2799 -0.9410 Ka 0.3075 -0.9256 Ka 0.5142 0.5930
n0 0.5069 0.7777 n0 -0.4667 0.7410 n0 -0.6872 0.8048
G 0.9205 -0.3598 G 0.4966 -0.8502 G -0.0706 0.9823
30.01 40 380.2
B0 0.7563 0.5247 B0 -0.4608 0.7843 B0 -0.4614 0.9506
Ka 0.3616 -0.8903 Ka 0.4325 -0.8658 Ka 0.6324 0.4716
n0 0.6030 0.7411 n0 -0.3312 0.8563 n0 -0.8290 0.6780
G 0.9287 -0.3433 G 0.2035 -0.9658 G 0.0331 0.9601
34.98 60 479.0
B0 0.8162 0.4857 B0 -0.3182 0.8844 B0 -0.6182 0.8781
Ka 0.2170 -0.9308 Ka 0.2124 -0.9623 Ka 0.8032 0.3535

a
TC : Crystallization temperature ( C).
b
Q: Feed flow rate (ml/min).
c
MS: Mixing speed (rpm)
187
188

Figure 5.16. Predicted Ka vs. measured Ka at steady state for TS samples (see
Tables 4.5 and 5.2 for experimental conditions and results). Multiple regression of
independent variables to Ka (i.e., y = 0 + 1 x1 + 2 x2 + 3 x3 + where x1 , x2 , and
x3 are crystallization temperature, feed flow rate, and mixing speed, respectively.)
was performed with 95% confidence level and showed 96% prediction accuracy.
189

5.3.3 Analysis of IS sample


The correlation analysis of crystallization temperature (TC ), flow rate (Q), and
mixing speed (MS) with dependent variables such as number of nuclei(n0 ), growth
rate (G), and nucleation rate (B0 ), as well as aggregation rate (Ka ) at steady state
for IS samples in Table 5.7 showed almost the same trends as for TS samples. The
three-factor interactions of independent variables on B0 and n0 as products of IS
samples are shown in Figure 5.17. When compared to Figure 5.11, the IS samples

Figure 5.17. Number of nuclei(n0 ) vs. nucleation rate (B0 ) at steady state for IS
samples (see Table 5.3).

with an impeller mixer shows almost same trends even with higher mixing speed
than the TS samples with a turbine mixer. At low feed flow rate, i.e., 20 and
40 ml/min, the slope of regression line fit in IS samples shows lower magnitude
compared to TS samples. As in all cases (T, TS, and IS samples), the growth rate
190

shows almost constant behavior with constant feed flow rate meaning that the feed
flow rate is the major controlling variable for the growth rate.
Figure 5.18 illustrates the power law relation of nucleation rate (B0 ) to growth
rate (G). The analysis of the unseeded CMSMPR crystallizer with size-independent

Figure 5.18. Nucleation rate (B0 ) vs. growth rate (G) at steady state for IS
samples (see Table 5.3).

growth leads to an exponential CSD [2]. In terms of population density n(L) and
particle size L, this is expressed as:
 
L
n(L) = n0 exp (5.7)
L
= G is the product of the growth rate G and the mean
where the mean size L
residence time . The nuclei population density (n0 ) equals B0 /G, where B0 is the
nucleation rate. Alternatively, the distribution can be expressed in terms of the
cumulative number oversize distribution, N (L), given by:
191
 
L
N (L) = N0 exp (5.8)
L
where N0 = B0 is the total number of crystals. The material balance relation
links the parameters of the above distribution to the suspension density, MT , by
MT
= 3 = n0 (G )4 = N0 (G )3 (5.9)
6c kv
where c is the crystal density, kv the volumetric shape factor, and 3 the unnor-
malized third moment of the distribution [101]. Strictly, the term material balance
relation applies when the suspension density is related to the change between inlet
and outlet solution concentrations. The above relation is really a self-consistency
check; it tests that the measured quantities satisfy the third moment for the
expected exponential distribution. The last two equalities in Equation 5.9 result
from the direct integration of the distribution of Equation 5.7. The estimate of
B0 is correlated with the estimate of G, according to B0 G3 . Figure 5.18
shows the results of several independent estimates. The 3 power dependence is
obvious therefore proving the accuracies in CSD measurements and in analytical
data obtained from those CSD measurements.
The correlation analysis with one independent variable constant and the other
two are changing was performed to explore the details of main effects to the
dependent variables, i.e., n0 , B0 , G, and Ka , shown in Table 5.12. When looking
at individual effects of independent variables to products, IS samples follow almost
the same behavior as explained previously for TS samples except for the constant
mixing speed case which shows dominant effects of crystallization temperature on n0
in all cases. However, as mixing speed increases the effects of TC on n0 , B0 , and Ka
tend to decrease contrary to the case in TS samples which showed increasing effect of
TC (except for G) with respect to mixing speed. Overall, the effect of crystallization
temperature was evident at 40 ml/min and at 677.8 rpm (479 rpm). The effects of
feed flow rate dominated on dependent variables at 30 C (35 C) and at lower
mixing speed. The mixing speed effect was dominating at lower crystallization
temperature and higher feed flow rate (lower). The indications in parentheses
are for TS samples. All of above differences between TS and IS samples can be
192

attributed to the different types of mixer, i.e., turbine and impeller. However, the
difference in speed of mixing and concentration cannot be ignored.
As an illustration for the effects discussed previously, the main effect of mixing
speed and feed flow rate in Figure 5.19 show an increase in number of nuclei and
nucleation rate, but decrease in growth and aggregation rate. The rapid increase
in number of nuclei (n0 ) and nucleation rate (B0 ) with respect to increase in mixer
rpm is mostly due to secondary nucleation. The linear decrease in aggregation rate
(Ka ) and growth rate (G) with increasing mixer rpm is expected due to turbulent
mixing of the crystallizers contents resulting in more breakage and less aggregation.
Figure 5.20 shows a plot of nucleation rate (B0 ) vs. mean size of particles (Lmean ).
The mean size of particles was smaller and the slope of power law curve fit showed
more gradual in IS samples when compared to TS samples mainly due to different
type of mixing pattern, i.e., uniaxial and biaxial mixing.

5.3.4 Influence of Different Mixer


The crystalline material is usually exposed to mechanical stress in any type of
crystallizer, caused by fluid shear or by crystal contact with other solid surfaces
such as other crystals, the stirrer, and the wall or other parts of the crystallizer.
Possible effects of mechanical stress are:

1. attrition at the crystal surface.

2. breakage of crystals or agglomerates.

3. the possibility that preordered species or surface nuclei may be removed from
the crystal surface by shear forces.

Mechanical stress can contribute to secondary nucleation and may also influence the
growth rate of large crystals. In Figure 5.21 the experimental population density
distributions are compared to illustrate the effect of different type of mixer on
resultant population number density with indications of the possible explanation
for the deviation from idealized MSMPR model, i.e., a solid straight line on this
plot. The experimental conditions were same at 30 C, 40 ml/min, and 480 rpm
193

(a)

(b)

Figure 5.19. Influence of mixing speed only (main effect) on number of nuclei(n0 ),
growth rate (G), and nucleation rate (B0 ), as well as aggregation rate (Ka ) at steady
state for IS samples. (a) IS 04, IS 05, and IS 06 samples and (b) IS 07, IS 08, and
IS 09 samples (see Tables 4.6 and 5.3 for experimental conditions and analytical
data).

with C02 feed solution concentration and with different types of mixer, i.e., turbine
and impeller. Although possible explanations are indicated in Figure 5.21 for the
194

Figure 5.20. Particle mean size (Lmean ) vs. nucleation rate (B0 ) at steady state
for IS samples (see Tables 4.6 and 5.3 for experimental conditions and analytical
data).

purpose of illustration, in general it is difficult to evaluate the specific reasons for


the deviation from the linear plot because the same type of deviation can result
from several causes. However, it was found that an impeller stirrer reduces the
particle breakage and attrition compare to a turbine and that the nucleation rate
depends on the agitation thus influencing the resultant particle size distributions
as shown graphically with 3D CSDs as a function of mean residence time ( ) at
same crystallization conditions but with different type of mixer in Figure 4.20 in
Section 4.5.6.
195

Figure 5.21. Influence of different type of mixer, i.e., turbine and impeller,
on population number density distribution at steady state. The experimental
conditions were same at 30 C, 40 ml/min, and 480 rpm with C02 feed solution
concentration except for the different types of mixer. The straight solid line
indicates the population density expected according to the idealized MSMPR
model.
Table 5.12. Correlation analysis of independent variables to the dependent variables at one
independent variable constant (other two are changing) for IS samples.

Variables Variables Variables


TC a Qb MSc
Q MS TC MS TC Q
n0 -0.2546 0.8807 n0 -0.3730 0.9048 n0 -0.8439 0.4998
G 0.7680 -0.6348 G 0.3068 -0.9350 G -0.0472 0.9731
25.02 20 480.5
B0 0.0810 0.9671 B0 -0.3875 0.9079 B0 -0.5120 0.9191
Ka 0.4250 -0.8860 Ka 0.3886 -0.8748 Ka 0.3337 0.6996
n0 0.1917 0.8744 n0 -0.5219 0.6462 n0 -0.8583 0.3365
G 0.8944 -0.3877 G 0.5269 -0.8305 G -0.0364 0.9705
30.03 40 677.8
B0 0.6476 0.6072 B0 -0.5339 0.6946 B0 -0.5895 0.8830
Ka 0.5590 -0.7316 Ka 0.4975 -0.8210 Ka 0.3197 0.7818
n0 0.3849 0.7096 n0 0.0230 0.9348 n0 -0.5695 0.4701
G 0.8628 -0.4348 G 0.1703 -0.9681 G -0.0598 0.9522
35.02 60 875.6
B0 0.6261 0.5449 B0 0.0247 0.9448 B0 -0.3952 0.8532
Ka 0.3833 -0.8052 Ka 0.1484 -0.9667 Ka 0.2491 0.5308

a
TC : Crystallization temperature ( C).
b
Q: Feed flow rate (ml/min).
c
MS: Mixing speed (rpm)
196
197

5.4 Model Validation


5.4.1 Phenomenological Models for Crystallization Modeling
Several models are identified for nucleation, growth aggregation and breakage.
The models for nucleation included are homogeneous nucleation, heterogeneous
nucleation, and secondary nucleation. In industrial crystallizers homogeneous nu-
cleation is very rare but heterogeneous and secondary nucleation are important.
With mechanical stirring secondary nucleation is usually the most important. The
models for secondary nucleation are best characterized by power law fits of experi-
mental data. Unfortunately no better models are available for secondary nucleation
at this time.
For crystal growth, available models include mono and poly surface nucleation
limited crystal growth, screw dislocation growth, diffusion limited growth, surface
integration limited crystal growth, chemical reaction and heat transfer limited
crystal growth. In industrial crystallizers either screw dislocation growth or the
combination of bulk diffusion and surface integration are the most common crystal
growth rates observed.
For aggregation, different collision rate kernels based upon the turbulent sub-
ranges include macro scale, inertial, transition, viscous taking from large scale to
the Kmologrov scale, the size of the smallest eddy, and within the smallest eddy
Brownian aggregation is available. In industrial crystallizers all turbulent scales
exist within the crystallizer, and therefore it is necessary to use the appropriate
aggregation rate kernel for the appropriate degree of turbulence.
For crystal attrition or breakage, several daughter distribution functions, also
called the primary progeny function, exist. They are characterized as power law
and exponential functions. All of the functions selected obey the constraint that
all of the progeny add up to the volume of the parent. The verdict is still out as to
which of these approaches is correct but all of the daughter distribution functions
chosen have adjustable parameters that give them flexibility to fit a broad range of
experimental data.
198

As a result of the work done in this project, it is clear that the models for
secondary nucleation and breakage are lacking a basis on first principles that is
needed for accurate prediction based upon first principles. However, the models
incorporated into the new OLI and Fluent software products to date are found
out to be useful for developing empirically based crystallization models that are
significantly advanced compared to previous crystallization models.

So far the models that have been implemented in the products of this project:
OLI Crystallization Compartment Modeler (CCModeler)

1. Nucleation - Power Law (for secondary nucleation)

2. Growth - Power Law, Diffusion Limited

3. Aggregation - Size Independent, Brownian Motion, Gravitational, Shear,


Particle Inertia

4. Breakage - Probability Theory

In addition, User-Defined is available for all phenomena.

5.4.2 Crystallization Model Validation


The major emphasis of this project was to develop models of crystallization and
to validate them against experimental data. Model development was performed by
OLI Systems and Fluent in cooperation with this project team. Model validation
took place and discussed in this section.
Validation experiments take place at various levels as is the current practice in
the validation literature. There was an effort to validate the fluid flow in the reactor
using velocity profile measurements in the literature; combined fluid flow and mass
balance in the reactor using residence time measurements, combined fluid flow and
heat transfer in the reactor using transient heat up experiments of the reactor,
and finally particle modeling using transient and steady state population balance
modeling. The total validation effort is called here as crystallization model valida-
tion. Population balance modeling has various aspects that have been compared to
199

experiments that attempt to separate phenomena. Separate experiments have been


performed to isolate aggregation and attrition or breakage studying them separately.
In addition, model comparisons are made to the crystallization experiments where
the experimental systems focus on the crystallization of NaCl where crystal growth
is the dominant reason for the ultimate particles. This section focuses on the various
modeling validation work performed.
Fluid flow modeling was performed with FLUENT 6.1T M done separately by
another researcher in this project. The fluid flow velocity field was compared
to those measured with laser velocimetry available in the literature. Qualitative
comparisons showed that the velocity field were similar for the turbine. Since our
mixer (and crystallizer) was of a different geometry than the experimental results
in literature, quantitative comparison could not be made. For this reason, residence
time distributions were used for direct quantitative comparison. For the residence
time distribution to be correct the fluid flow velocity field must be accurately de-
scribed throughout the tank as well as the concentration profile of a tracer chemical
in the tank. A direct comparison of the residence time distribution is provided in
Appendix B. To get the residence time distribution both the momentum balance
and the mass balance equations are solved dynamically for the tank. The results
of the outlet concentration to a pulse input in the inlet is shown in Figure B.14.
Here we see that Fluent accurately predict the behavior in the tank for all rpms.
At 200 rpm the tank behaves ideally with a mean residence time equivalent to
the ratio of the tank volume to the feed flow rate and the standard deviation of
the residence time distribution divided by the mean residence time equal to 1.0
within experimental error. At all rpms less than 50 and specifically at 20 rpm
shown in Figure B.14 (b), the tank is not ideal due to the delay of the mixing
but Fluent predicts this reasonably well. Therefore the accuracy of the momentum
balance and the mass balance equations are validated with this work for a turbulent
stirred tank. Additional work with respect to energy balance equation calculation
validations were also performed with Fluent for the heat transfer within the tank.
These results were as accurate as the residence time distribution measurements.
200

For crystallization validation, Table 5.13 presents inlet and outlet conditions
and compositions for T, TS, and IS samples involving three salt solution of CaCl2 ,
KCl, and NaCl. The precipitation analyses of inlet and outlet solutions are obtained
using OLI stream analyzer. The Figure 3.3 and Table 3.1 in Section 3.3.1 illustrate
one of solution concentration (C01 ). As the solution is cooled, NaCl is the first
to precipitate at a temperature of approximately 45, 35, and 31 C for C01 , C02 ,
and C03 , respectively. The next salt is a CaCl2 6H2 O staring to crystallize out at
approximately 28, 27, and 27 C for the same order of solution as before. The KCl
was not crystallized out at temperature above 20 C for all samples. The experi-
mental results confirmed that there was no CaCl2 crystallizing out at crystallization
temperature of 25 C or above. With the analysis being done, preliminary kinetic
model parameter estimates were determined for combined kinetic phenomena using
CCModeler by OLI systems. Parameter estimates were obtained by trial and error
while monitoring the following variables.

1. the first three moments.

2. the maximum weight fraction interval.

3. the lower and upper interval where the weight fraction was 1%.

4. suspension density.

A combined kinetic model of nucleation, growth, aggregation, and breakage was


tried with the default models for each of the kinetic phenomena as

1. nucleation power law with supersaturation ratio term exponent.

2. growth power law with supersaturation ratio term exponent.

3. aggregation size independent

4. breakage probability theory.

Figures 5.22 through 5.24 show the measured vs. predicted discrete volume fraction
of particle size distributions as an example for each of T, TS, and IS samples.
201

The experimental conditions and results are given previously for each of sample

Figure 5.22. Measured and predicted volume fraction of particle size distribution
for T 16 sample considering nucleation, growth, aggregation, breakage kinetics (see
Table 5.13).

designs. From these very preliminary results, it appears that nucleation and growth
were the dominant influences for the crystallization of NaCl. The high degree of
correspondence in the downward slope between 100 and 200 m particle size range
was noticed for T and TS samples. The IS sample with higher mixing speed (481
rpm) requires much more rigorous works should be done in this model. This model
has not been accommodated in the different type of mixer and speed of mixing,
yet. Table 5.14 shows the estimated values for the kinetic model parameters. These
results and the ease of performing this preliminary analysis demonstrate the promise
of this modeling technology.
202

Figure 5.23. Measured and predicted volume fraction of particle size distribution
for TS 16 sample considering nucleation, growth, aggregation, breakage kinetics
(see Table 5.13).

Figure 5.24. Measured and predicted volume fraction of particle size distribution
for IS 16 sample considering nucleation, growth, aggregation, breakage kinetics (see
Table 5.13).
Table 5.13. Inlet and outlet conditions and composition for T, TS, and IS samples.

T TS IS
Inlet Outlet Inlet Outlet Inlet Outlet
Pressure (kPa) 101.325 101.325 101.325 101.325 101.325 101.325
Temperature (K) 343 303 343 303 343 303
Stream
Mass flow (g/min) 87.9024 87.9024 87.9024 87.9024 87.9024 87.9024
Volume Flow (ml/min) 60 59.0963 60 59.0628 60 59.0564
Mass flow (g/min) 88.6884 88.545 87.9024 87.8313 88.0944 88.0373
Liquid Mass density (g/ml) 1.4781 1.5 1.4650 1.4879 1.4682 1.4914
Volume flow (ml/min) 60 59.03 60 59.03 60 59.03
Mass flow (g/min) 0 0.1434 0 0.0711 0 0.0571
Solid Mass density (g/ml) 0 2.1638 0 2.1638 0 2.1638
Volume flow (ml/min) 0 0.0663 0 0.0328 0 0.0264
Water 50.63 50.84 51.69 51.91 51.44 51.66
Composition CaCl2 47.63 47.82 46.88 47.08 47.09 47.29
(wt%) KCl 1.07 1.07 0.93 0.93 1.03 1.03
NaCl 0.67 0.41 0.5 0.42 0.44 0.42
203
Table 5.14. Kinetic model parameter estimation for T, TS, and IS samples.

Kinetics Model Parameter T sample TS sample IS sample


Model Option Power Law Power Law Power Law
Power Law Constant (#/m3 s) 2.00E+08 1.00E+15 5.00E+07
Nucleation
Supersaturation Exponent 0.5 2.77 0.9
Supersaturation Density Exponent 1 1 0.91
Model Option Power Law Power Law Power Law
Power Law Constant (m/s) 8.00E-05 1.00E-04 1.00E-06
Growth
Supersaturation Exponent 1.5 1 0.7
Length Ratio Exponent 0.5 0 0
Model Option Size Independent Size Independent Size Independent
Aggregation
Size Independent Constant (m3 /s) 1.00E-14 1.00E-15 1.00E-15
Model Option Probability Theory Probability Theory Probability Theory
Breakage Probability Theory Constant (s1 ) 5.00E+07 5.00E+07 2.00E+08
Length Ratio Exponent 3 3 3
204
205

5.5 Summary
With the application of Hounslows solution into the measured CSD at 10 for
all the experimental runs, the number of nuclei (n0 ), growth rate (G), nucleation
rate (B0 ), and aggregation rate (Ka ) were obtained for each set of designs. However,
the aggregation of particles was noticed in size range less than 50 m in most of
crystallization experiments, therefore, the line fitting was unsuccessful with particle
size less than 50 m. The deviation of CSD from the linear line in large particle
size region due to particle breakage was also not included in this analysis.
To show the relative importance of independent variables, i.e., crystallization
temperature (TC ), feed flow rate (Q), and mixing speed (M S), affecting a crys-
tallization process, as well as the importance of their interactions to the research
experiments, the analytical data obtained with the application of Hounslows con-
stant rate equation was analyzed following the statistical design of experiments
according to factorial design analysis and other various statistical analyses. With
these analyses it was possible to explore the detailed crystallization phenomena
and to find out the controlling variables for the purpose of crystallization at certain
conditions.
CHAPTER 6

GENERAL CONCLUSION

6.1 Conclusions
The following general conclusions are drawn from the results of this research
project.
Reactor characterization
A comprehensive experiment was performed to characterize a laboratory scale
reactor and ultimately to find out a suitable condition for the crystallization system
built in our Lab.

1. The mean and variance of the residence time distribution for a stirred tank
deviating from ideal values were observed at low mixer rpm. As the mixer
rpm increased, the mean and variance of the residence time distribution
were observed approaching the ideal values1 .

2. For the nonbaffled tank, ideal behavior was observed at 100 rpm or an
mixer Reynolds number of 3,877 and above, while for the baffled tank ideal
behavior was being observed at 60 rpm or an mixer Reynolds number of
2,326 and above using a turbine. The rpms where ideal mixing occurred
take place for impeller Reynolds numbers in the transition range.

3. Predictions of the RTD and its mean and variance using CFD with a k
model of turbulence showed good agreement with experiment for
transitional turbulent flow regimes in the tank2 .
1
Appendix B.
2
The prediction of RTD to compare with experimental results was provided by an affiliation
of this research project team and with their contributions.
207

Crystallization of NaCl

1. The system designed and constructed requires precise control of equipment


to acquire dependable data. This was accomplished by using automatic
system control and data acquisition3 . The realtime crystallization conditions
were monitored for each experiment with on-line data acquisition measuring
all parameters in every second till the completion of crystallization
experiment, typically 10 times of mean residence time (10 ) 4 .

2. A substantial amount of experiments involving crystallization of pure NaCl


from a brine solution was performed to show the effects of independent
variables, i.e., crystallization temperature, feed flow rate, and mixing speed,
to dependent variables, i.e., number of nuclei, nucleation rate, growth rate,
and aggregation rate for the crystallization processes following statistically
designed experiments. These data are available on the web to other groups
interested in crystallization modeling of a CMSMPR crystallizer5 .

3. The analysis by OLI stream analyzer showed as an helpful tool to predict


precipitation of complex ionic solution. However, there was a discrepancy in
predicting precipitation of CaCl2 6H2 O due to the difference in
thermodynamics and kinetics between ideal and real system6 .

4. The detailed guidelines for preparation, operating procedure, and sampling


of crystallization are established and given in Appendix C7 .

Characterization of NaCl Particles

1. The SEM figures with the aid of microscopic photos demonstrate the
detailed particle shape (octahedron) and size (1 to 300 m) to explain the
3
Section 2.5
4
Section 4.3.3.
5
URL: http://www.che.utah.edu/ring/CrystallizationWeb.
6
Section 3.3.1, 4.2.3, 4.4.1, and Appendix F.
7
Section 4.3.2.
208

possible phenomena of how these particles are evolving in a CMSMPR


crystallizer8 .

2. Even with the careful treatments of slurry sample to measure the CSD in
LS230, the change in temperature of 2 to 5 C during the measurement was
enough to alter the particle characteristics by either growth or dissolution.
Therefore, a special technique was required to make the particles stable and
the CSD measurements successful. The application of CdCl2 impurity to
limit zero-growth rate and to stabilize NaCl particles was discussed with
underlying theory and showed promising result of CSD measurement
without changing the characterization of particles9 .

3. The CSDs measured and microscopic photos taken at every mean residence
time during the crystallization showed distinct characterizations, i.e., the
CSD at the beginning of crystallization and the evolution of CSDs with
time, of NaCl crystals depending on various experimental parameters, such
as, solution concentration, crystallization temperature, feed flow rate,
mixing speed, and type of mixer10 .

Breakage and Aggregation Experiments

1. Additional experiments have been performed that focus on aggregation and


attrition without the influence of nucleation and crystal growth being
active11 . The independent variables for this experiment were reaction
temperature, mixing speed, and different type of mixer.

Analysis and Validation

1. Overall effects of crystallization temperature, feed flow rate, and mixing


speed, as well as different type of mixer on mean size of particles and yield
8
Section 4.4.
9
Appendix F.
10
Section 4.4.2 and 4.5.
11
Section 4.6.
209

were discussed in Section 4.5.7. However, due to the result of difficulties


inherent in obtaining accurate liquid compositions the effects on yield were
not clear for T and TS samples12 .

2. With the application of MSMPR model and Hounslows equation to


experimentally measured CSD at steady state (typically at 10 ), it was
found out that n0 = 2.34 1013 1.986 1014 /m4 ,
G = 4.40 103 1.01 102 m/s, therefore the nucleation rate to be
B0 = n0 G = 1.03 105 2.01 106 /m3 s were typical within our
experimental conditions13 .

3. With the application of Hounslows equation and nonlinear regression data


fit to experimental results of population number density, aggregation rate
constant, Ka = 2.58 1016 1.11 1015 m3 /s, was obtained without the
consideration of breakage in this analysis within our experimental
conditions14 .

4. Various statistical experimental analyses were performed to show the relative


importance of independent variables, i.e., crystallization temperature, feed
flow rate, mixing speed, and different type of mixer, affecting a
crystallization process as well as the importance of their interactions to the
research experiments following the statistical design of experiments15 .

5. The factorial design analysis to nucleation rate, growth rate and aggregation
rate without considering experimental results at 25 C showed the
importance of feed flow rate and mixing speed as main effects and their
interaction to the resultant analytical data16 .
12
Section 4.5.7.
13
Section 5.1.
14
Section 5.1.
15
Section 2.7.2 and 5.3.
16
Section 5.3.
210

6. More rigorous analyses for T, TS, and IS samples were performed to show
how the independent variables are interacting together or alone onto the
product variables to find out the controlling variables at certain condition of
crystallization and also, to show how the role of independent variables on
products are changing with change in crystallization conditions17 .

7. Predictions of multiple linear regression on nucleation rate, growth rate, and


aggregation rate of TS samples showed 91, 99, and 96% accuracies,
respectively, with confidence level of 95%18 .

8. The experimental population density distributions were compared to


illustrate the effect of different type of mixer on resultant population
number density with indications of the possible explanation for the
deviation from idealized MSMPR model. It was found that an impeller
stirrer reduces the particle breakage and attrition compare to a turbine and
that the nucleation rate depends on the type of agitation thus influencing
the resultant mean size of particles as showed graphically with 3D CSDs as
a function of mean residence time ( ) at same crystallization condition but
with different type of mixer19 .

9. As a result of the work done in this project, it is clear that the models for
secondary nucleation and breakage are lacking a basis on first principles that
is needed for accurate prediction based upon first principles. However, the
models incorporated into the new OLI and Fluent software products to date
are found out to be useful for developing empirically based crystallization
models that are significantly advanced compared to previous crystallization
models20 .
17
Section 5.3.1, refsec:aofts, and 5.3.3.
18
Section 5.3.2.
19
Section 4.5.6 and 5.3.4.
20
Section 5.4.
211

6.2 Recommendations
1. The analytical values, i.e., number of nuclei (n0 ), growth rate (G),
nucleation rate (B0 ), and aggregation rate (Ka ), obtained with the
application of Hounslows model are not the absolute ones, but they tend to
show accurately enough the possible phenomena for a crystallization in a
CMSMPR crystallizer. However, much more rigorous work should be done
to account for the breakage and secondary nucleation into the model.

2. The ex situ CSD measurements performed in this project account for the
NaCl particles with size ranging from 1 to 300 m with currently available
particle size analyzer. The treatment for NaCl particles to stabilize with the
application of Cd2+ impurity was required for this ex situ CSD measurement
due to the different environments. For the study of crystallization however,
it is important to be able to monitor the reactions and kinetics of
crystallization with high accuracy without any alterations and modifications
to the crystallization phenomena in a CMSMPR crystallizer. With current
advance in extracting large amounts of information from a single
multiwavelength measurement, the in situ application of multiwavelength
ultraviolet/visible (UV-VIS) spectroscopy could be one of possible
candidates with added benefits of being rapid, inexpensive, and relatively
simple to operate; the distribution of intensities as a function of wavelength
depends on the size, shape, and chemical composition of the sample.
APPENDIX A

CRYSTALLIZER GEOMETRY AND


CONFIGURATION

The geometry of the 1.4 ` laboratory stirred crystallizer is shown in Fig-


ure A.1. It has four baffles of the conventional thickness in contact with the

Figure A.1. Geometry of the laboratory scale crystallizer (1.4 ` working volume)
with baffles and a turbine.

wall of the crystallizer, a 6-bladed flat disc turbine (Rushton type) or 6-bladed
213

pitch blade impeller and three internal pipes, two feed ports and a thermal well
protruding into the reactor from its top. One of the feed tubes dead ends at the
approximate height of the impeller and the other is bent to feed directly below
the impeller. Also protruding ever so slightly into the top of the liquid are a
pH probe, conductivity probe, thermocouple, potassium specific ion electrode and
level indicator. All of these instruments as well as the feed pump, product pump,
the jacket feed temperature, thermal well temperature and the rpm and torque on
the stirrer are connected to an OPTO 22T M data acquisition and control system
collecting data at time intervals of 1 s. The crystallizer control system has the
capability of controlling the feed source and crystallizer temperature to 0.25 C,
the feed and product flow rate to 0.5 ml/min, liquid level to 3.9 mm and mixing
speed to 3 rpm.
Figures A.1 through A.3 show crystallizer geometry with baffles and a turbine
in place along with the location of inlet and outlet. The scales in figure are all in
mm. Also the dimension of standard baffle width 1/12 of the vessel diameter is
given in Table A.1 along with the dimension of mixer. Figure A.4 and A.5 show
stirrer geometry of a 6-bladed flat disc turbine (Rushton type) and a 6-bladed pitch
blade impeller.
214

Figure A.2. Dimension and configuration of a 1.4 ` crystallizer (sideview).


215

Figure A.3. Dimension and configuration of a 1.4 ` crystallizer (topview).


216

Table A.1. Dimension of vessel, standard baffle, and mixer, as well as in


and outlet tube.

Description Dimension (mm)


Vessel Diameter, D 114
Vessel
Vessel Height, H 145
Baffle Width, W 9.5 (= D/12)
Baffle Thickness, T 2.5
Baffle
Gap between Wall and Baffle, Gw 1.5 (' D/72)
Gap between Vessel Base and Baffle, Gv 2.4 (' W/4)
Blade Height 11.5
6-bladed flat Blade Width 11.5
disc turbine Blade Thickness 1.5
(Rushton type) Diameter 46
Blade Angle 90
Blade Height 11.5
Blade Width 14.5
6-bladed pitch
Blade Thickness 1.5
blade impeller
Diameter 47
Blade Angle 45
In and Outlet Inner Diameter 1.6
217

Figure A.4. Dimension and configuration of a 6-bladed flat disc turbine (Rushton
type).
218

Figure A.5. Dimension and configuration of a 6-bladed pitch blade impeller.


APPENDIX B

REACTOR CHARACTERIZATION

Residence Time Distributions in a Stirred Tank - Comparison of CFD predictions with


Experiment was published in Journal of Industrial and Engineering Chemistry [92].
The materials provided here are mostly from the publication with more updates.

B.1 Residence Time Distribution in a CSTR


Continuous stirred tanks are used ubiquitously in the chemical process industry
for mixing, reaction, and crystallization. The mixing in a continuous stirred tank
is often not ideal. The residence time distribution (RTD) is one of the ways to
characterize the nonideal mixing in the tank. Comparison of the measured RTD
with that of an ideal reactor allows the process engineer to diagnose the ills of the
tank and mixer design. The engineer can then use an appropriate mixing model
for the tank in combination with the kinetics of the reaction to be performed in the
tank to develop an appropriate model for the reactor [91].
Constant stirred tank mixers and reactors have been the focus of various RTD
studies over the years. Experiments carried out by Khang and Levinspiel [102],
Zaloudik [103], and Gianetto and Cazzulo [104] for constant stirred tanks of stan-
dard size (diameter equals height) with standard baffle designs and various impeller
designs have been done for a broad range of tank sizes, feed flow rates, and impeller
rpms. These studies show a transition from nonideal to ideal RTD behavior
as rpm is increased. Idealized RTDs are always observed for an mixer Reynolds
number in the transition flow regime, with the critical Reynolds number depending
upon the type of impeller being used and the baffle structure. Various theories
have been presented to predict the RTD for nonideal RTDs including one and
two parameter models and three and higher parameter models [93, 104, 105, 106].
220

These higher parameter models consider by-pass, dead zones, recycle, ideal back
mixed regions, and plug flow regions. These models have been used to predict
the nonideal RTD [87, 93, 102, 103, 104, 105, 106, 107] for the performance of
reactors operating with various chemical reactions [108, 109, 110] and nonideal
reactor scale-up [104, 111].
This work measures the RTD of a laboratory reactor both with and without
baffles and makes comparisons of these measurements with predictions for the RTD
using computational fluid dynamics (CFD) with a k turbulence model coupled
with a dynamic two-species, fluid and tracer, mass balance operating with a step
change of the concentration of tracer in the feed.

B.1.1 Experimental Section


The geometries of the 1.4 ` laboratory stirred tank reactor are shown in
Figure A.1 with all internal dimensions given in Table A.1. For the unbaffled
experiments the baffles were removed and the tank was filled to 1.3 `. The
baffled tank has four baffles of conventional thickness, a turbine or an impeller,
and three internal pipes protruding into the reactor from its top, two of which are
feed ports and one of which is a thermal well. One of the feed tubes dead-ends
at the approximate height of the mixer and was not used in these experiments;
the other tube is bent to feed directly below the mixer. Also protruding ever
so slightly into the top of the liquid are the outlet tube, pH probe, conductivity
probe, thermocouple, potassium-specific ion electrode, and level indicator (not
shown in Figure A.1). The reactor output flows directly into a flow cell of a
UV-Vis spectrometer operated at a fixed wavelength. All of these instruments,
as well as the feed pump, product pump, the jacket feed temperature, thermal well
temperature, and the rpm and torque on the stirrer, were connected to an OPTO
22T M data acquisition and control system collecting data at time intervals of 1 s.
The reactor control system has the capability of controlling the feed source and
reactor temperature to 0.25 C, the feed and product flow rate to 0.5 ml/min,
the liquid level to 3.9 mm, and the mixing speed to 3 rpm.
221

The ratio of liquid level to the tank diameter (H/D) of the vessel in Figure A.1
is 1.27 that is within the general range of 1.0 to 1.5 for most industrial stirred
tank precipitator [20, 86]. To reduce the energy input to the system while main-
taining mixing uniformity, a standard baffle design [87] was used consisting of four
flat vertical plates, radially-directed (i.e., normal to the vessel wall), spaced at
90 around the vessel periphery running the length of the vessels straight side.
Standard baffle widths are between 1/10 and 1/12 of the vessel diameter (D/10 or
D/12), see Table A.1 in Appendix A) for details. The gaps with the vessel wall and
base are left to allow the flow to clear the baffles. Recommended gaps are equal to
1/72 of the vessel diameter (D/72) between the baffles and the vessel wall, and 1/4
to one full baffle width between the bottom of the baffles and the vessel base. More
of the detailed baffling information could be found in Myers, et al. [88]. Stirring
was achieved by a 6-bladed flat disc turbine (Rushton type) or a 6-bladed pitch
blade impeller. The dimensions and configurations of a turbine and an impeller are
shown in Figures A.4, A.5, and Table A.1 in Appendix A.
The rpm of peristaltic pumps was controlled by an OPTO 22T M control system
establishing the feeding and removal of product solution from the tank. Flow rate
was obtained by the careful and repeated calibration of the speed of pump and tube
size to ensure the desired feed and withdrawal rates as well as a constant residence
time. The OPTO 22 system also measured the on-line sensor quantities; tempera-
ture, pH, absorbance, and conductivity, as well as potassium ion concentration. All
of these measurements were used to monitor the concentration of the tracer inside
and exit of the reactor. The calibration for all of these systems is done before each
run to assure measurement accuracy.
The RTD measured was the combination of process dynamics and sensor dynam-
ics [112]. The sensor dynamics are much faster than that of the process. The sensor
time constants are the following: conductivity, which has a measurement time of 1
s; K + and pH electrodes, which have measurement times of 4 8 s; absorbance,
which has a measurement time related to the flow rate through the absorbance
222

sample cell, or at worst 10 s; and temperature, which has a measurement time of


0.5 s.

B.1.2 Measurement of RTD


RTD is determined experimentally by injecting a 10 ml volume of inert multi-
component chemical tracer (a hot solution of 0.1 mg/l methyl blue dye, 50 mg/l
NaCl, 20 mg/l KCl and 3.6 mg/l HCl with water) into the tank at time zero. In
all experiments, the feed tube is the bent one that feeds just below the impeller.
On-line analyzers detected each of the species in the tracer; blue dye with the aid of
video camera1 and uv-vis spectrometry, temperature with thermocouple, salt with
conductivity, and H+ concentration with pH meter were measured simultaneously
with various conditions such as different flow rate and mixer rpm. Great care was
used in synchronizing the pulse input with the initiation of data accumulation for
RTD analysis and to calibrate the pumps to assure that the reactor space-time
[= Vtank /Q] was accurately measured. All the experimental conditions used to
measure RTDs are given in Table B.1 for the without-baffle experiments and
Table B.2 and B.3 for the with-baffles experiments. Table B.1 gives the Reynolds
number for the mixer and the feed tube.
Table B.2 for a turbine and Table B.3 for an impeller also give the mixer
Reynolds number and the mixer power number, calculated from the measurement
of the mixer torque. For all of these experiments, data was taken at 1 s intervals.
Figure B.1 and B.2 illustrate bad mixing at 0 rpm and good mixing at 287 rpm
for the tank without baffles and with a turbine as measured by conductivity, pH,
and temperature. The uv-vis spectrometry data and K + data are not shown in
these figures, however, they are similar to those for the other curves presented in
these figures. All of these curves show a fast increase at an early time followed by
an exponential decrease as expected for a continuous stirred tank. The delay of the
increase from time zero and the roughness of the increases clearly indicate parasitic
mixing behavior as compared to the ideal mixing curve since it should increase
1
Video clips of these experiments are available at http://www.che.utah.edu/
ring/CrystallizationWeb/RTD%20Videos/dads%20ex%203.mov
223

Table B.1. Experimental conditions and results of RTD mea-


surements without baffles and with a turbine.

Qa MS b tm c Vd V /Q Re e
/tm f
(ml/min) (rpm) (min.) (ml) (min.) FT Mg
0 18.90 0.636 1,290 11.73 0
10.2 15.94 0.665 1,290 11.73 395
30 12.28 0.964 1,290 11.73 1,163
49.7 11.75 0.978 1,290 11.73 1,927
110 1,676
69.5 11.66 1.011 1,290 11.73 2,695
90.6 11.61 1.021 1,292 11.75 3,513
190 11.65 1.003 1,294 11.76 7,367
286.7 11.76 1.026 1,298 11.80 11,120
0 8.09 0.746 1,290 5.73 0
9.8 7.26 0.819 1,290 5.73 380
29.7 6.26 0.915 1,290 5.73 1,152
50.2 5.93 0.965 1,290 5.73 1,946
225 3,428
70 5.96 0.988 1,290 5.73 2,714
91.6 5.67 1.011 1,292 5.74 3,551
190.6 5.78 1.015 1,294 5.75 7,390
288 6.10 1.007 1,298 5.77 11,170
0 5.32 0.863 1,290 4.37 0
11 4.70 0.939 1,290 4.37 426
30.2 4.24 1.017 1,290 4.37 1,171
50.5 4.41 0.974 1,290 4.37 1,958
295 4495
69.9 4.42 1.016 1,290 4.37 2,710
90 4.32 1.047 1,292 4.38 3,480
188.5 4.27 1.024 1,294 4.39 7,308
290 4.42 1.041 1,298 4.40 11,240
Table continues on next page.

a
Q: flow rate.
b
MS : mixer speed.
c
tm : mean residence time.
d
V : reactor working volume.
e
Re: Reynolds number.
f
F T : feed tube.
g
M : mixer.

abruptly in an instantaneous rise at time zero and decay exponentially with time
thereafter. The RTD is obtained from these experimental data by normalization in
typical manner [93].
224

Continued from Table B.1.


Qa MS b tm c Vd V /Q Re e
/tm
(ml/min) (rpm) (min.) (ml) (min.) FT f Mg
0 4.15 0.866 1,290 3.23 0
10.5 3.78 0.890 1,290 3.23 407
30.5 2.96 0.920 1,290 3.23 1,183
51 3.21 1.009 1,290 3.23 1,977
400 6,095
70.7 3.24 0.978 1,290 3.23 2,741
89.8 3.19 1.004 1,292 3.23 3,482
190 3.17 1.043 1,294 3.24 7,367
287 3.38 1.060 1,298 3.25 11,130
0 3.53 0.743 1,290 2.43 0
9.7 3.28 0.746 1,290 2.43 376
30 2.67 0.879 1,290 2.43 1,163
50 2.72 0.899 1,290 2.43 1,939
530 8,075
70 2.53 0.883 1,290 2.43 2,714
90 2.43 0.981 1,292 2.44 3,489
189.5 2.39 1.043 1,294 2.44 7,347
288 2.45 1.061 1,298 2.45 11,170

a
Q: flow rate.
b
MS : mixer speed.
c
tm : mean residence time.
d
V : reactor working volume.
e
Re: Reynolds number.
f
F T : feed tube.
g
M : mixer.

C(t) C(t = 0)
E(t) = R (B.1)
0
[C(t) C(t = 0)]dt
where C(t) represents the concentration (or temperature) of tracer, e.g., Figures B.1
and B.2. An example of the RTD is plotted in Figure B.3. Even for this good mixing
case there are deviations from ideal mixing at short residence times where the initial
rise is not immediate and is not a smooth curve. This figure also contains a plot of
the ideal RTD for the stirred tank for comparison purposes. The mean residence
time (tm ) was calculated by integrating the RTD as follows [91],
Z
tm = tE(t)dt (B.2)
0

The variance, or square of the standard deviation of the RTD is calculated using:
225

Table B.2. Experimental conditions and results of RTD measurements


with baffles and a turbine.

Qa MS b tm c V V /Q Re d
/tm MPN g
(ml/min) (rpm) (min.) (ml) (min.) FT e Mf
0.0 106.66 0.741 1400 82.35 0
8.6 111.73 0.819 1,400 82.35 376 1,194,000
28.5 87.51 0.928 1,400 82.35 1,148 48,290
47.8 86.40 0.939 1,400 82.35 1,904 10,990
17 307
67.1 86.84 0.936 1,400 82.35 2,714 3,649
87.5 81.77 0.937 1,400 82.35 3,416 1,668
185.1 82.87 0.938 1,400 82.35 7,289 249
283.8 81.98 0.916 1,400 82.35 11,040 108
0.0 50.32 0.766 1,400 40.58 0
8.8 54.08 0.833 1,400 40.58 446 729,500
28.5 44.97 0.985 1,400 40.58 1,221 40,370
48.2 44.08 1.036 1,400 40.58 1,973 9,866
34.5 614
41.1 37.71 0.930 1,400 40.58 2,710 3,666
87.4 40.79 0.931 1,400 40.58 3,551 1,524
185.3 40.77 0.906 1,400 40.58 7,367 243
283.8 40.29 0.885 1,400 40.58 11,090 107
0.0 31.37 0.866 1,400 27.18 0
9.5 30.46 0.838 1,400 27.18 446 729,500
29.5 27.68 0.896 1,400 27.18 1,221 40,370
48.6 27.26 0.912 1,400 27.18 1,973 9,866
51.5 614
67.7 28.38 0.952 1,400 27.18 2,710 3,666
88.0 28.06 0.948 1,400 27.18 3,551 1,524
186.2 28.05 0.940 1,400 27.18 7,367 243
284.9 27.07 0.924 1,400 27.18 11,090 107

a
Q: Flow rate
b
MS : Mixer speed
c
tm : Mean residence time
d
Re: Reynolds number
e
FT : Feed tube
f
M : Mixer
g
MPN : Mixer power number

Z
2
= (t tm )2 E(t)dt (B.3)
0

The magnitude of this 2nd moment is an indication of the spread of the RTD.
226

Table B.3. Experimental conditions and results of RTD measurements


with baffles and an impeller.

Qa MS b tm c V V /Q Re d
/tm MPN g
(ml/min) (rpm) (min.) (ml) (min.) FT e Mf
0 111.82 0.752 1,400 78.65 0
9.1 101.73 0.767 1,400 78.65 376 1,194,000
28.4 85.18 0.962 1,400 78.65 1,148 48,290
17.8 47.9 80.44 0.987 1,400 78.65 307 1,904 10,990
67.1 81.86 1.021 1,400 78.65 2,714 3,649
86.8 78.94 0.990 1,400 78.65 3,416 1,668
185.0 78.39 0.990 1,400 78.65 7,289 249
0 50.32 0.766 1,400 37.33 0
9.0 48.05 0.835 1,400 37.33 446 729,500
29.1 41.91 0.907 1,400 37.33 1,221 40,370
37.5 47.7 39.10 0.940 1,400 37.33 614 1,973 9,866
67.6 37.71 0.969 1,400 37.33 2,710 3,666
86.8 37.87 0.992 1,400 37.33 3,551 1,524
185.1 37.44 0.982 1,400 37.33 7,367 243
0 29.29 0.871 1,400 25.09 0
9.4 28.87 0.834 1,400 25.09 446 729,500
28.6 29.06 0.897 1,400 25.09 1,221 40,370
55.8 47.4 27.31 0.947 1,400 25.09 614 1,973 9,866
67.1 25.73 0.960 1,400 25.09 2,710 3,666
87.0 25.30 0.978 1,400 25.09 3,551 1,524
185.1 25.09 0.985 1,400 25.09 7,367 243

a
Q: Flow rate
b
MS : Mixer speed
c
tm : Mean residence time
d
Re: Reynolds number
e
FT : Feed tube
f
M : Mixer
g
MPN : Mixer power number

In Figure B.4, the mean resident time obtained from the experimental data
is plotted as a function of mixer rpm for various feed flow rates (high flow rate
case) for the tank without baffles and a turbine. Figures B.5 and B.6 show the
experimental data plotted as a function of mixer rpm for various feed flow rates
(low flow rate case) for the tank with baffles and a turbine and for the tank with
227

Figure B.1. Concentration profile measured with conductivity, H+ concentration,


and temperature. Operating condition: flow rate = 110 ml/min, mixing speed = 0
rpm, and reactor volume = 1,300 ml for a tank without baffles and with a turbine.

baffles and an impeller, respectively. Here we see that the mean residence time
decreases with increasing rpm until a constant value (solid line) is reached. This
constant value is approximately Vtank /Q, the space-time for the reactor, where Vtank
is the reactor volume and Q is the volumetric flow rate. The error in the calculation
of the mean residence time, tm , is 8% of Vtank /Q for all experiments and was
determined by experiments done in duplicate and some in triplicate. These errors
are due to the variations in tank volume caused by the liquid level control (3.9
mm corresponding to 41 ml) and to the variations in the flow rate caused by
a small calibration drift and pulsating peristaltic pump flow. By comparison of
these three figures (Figures B.4, B.5, and B.6), one can easily see the difference in
RTD as a function of mixer rpm, flow rate, and type of mixer. In high flow rate
case (Figure B.4), the constant value (Vrank /Q) has reached at below 100 rpm of
mixing speed compared to above 100 rpm with low flow rate case (Figure B.5). Also
228

Figure B.2. Concentration profile measured with conductivity, H+ concentration,


and temperature. Operating condition: flow rate = 110 ml/min, mixing speed =
287 rpm, and reactor volume = 1,300 ml for a tank without baffles and with a
turbine.

due to the different mixing pattern between a turbine and an impeller, Figure B.6
with axial mixing shows smooth decrease in tm with little longer space time while
Figure B.5 shows fluctuation of tm due to bi-axial mixing.
In Figures B.7 and B.8, the variance of the RTD divided by the space-time is
plotted as a function of mixer rpm for various feed flow rates for the tank without
baffles and with a turbine and for the tank with baffles and with an impeller,
respectively. The error in the variance of the residence time divided by the mean
residence time is 12% at all rpms and is larger than that of the mean residence
time because two integrals with their inherent errors are involved in the variance
divided by mean residence time calculation. For an ideal reactor the value of /tm
should be 1.0, which is observed, within the experimental error, at higher impeller
rpms. With the Figures B.7 and B.9 one can see that mixer speed of 100 rpm was
enough to approach perfect mixing for the unbaffled tank. With baffles, the results
229

Figure B.3. RTD measured with H+ concentration. Operating condition: flow


rate = 110 ml/min, mixing speed = 287 rpm, and reactor volume = 1,300 ml for a
tank without baffles and with a turbine.

Figure B.4. Comparison of tm vs. mixer rpm at different flow rate for 1,300 ml
laboratory reactor without baffles and with a turbine.
230

Figure B.5. Comparison of tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor with baffles and a turbine.

Figure B.6. Comparison of tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor with baffles and an impeller.

are shown in Figures B.8 and B.10 (low flow rate case). With baffles (low flow rate
case), the results are similar to those without baffles (high flow rate case), however,
231

Figure B.7. Comparison of /tm vs. mixer rpm at different flow rates for 1.3 `
laboratory reactor without baffles and with a turbine.

Figure B.8. Comparison of /tm vs. mixer rpm at different flow rate for 1.4 `
laboratory reactor with baffles and an impeller.

the approach to ideal reactor behavior takes place at 80 rpm as measured by


/tm and tm /(V /Q) which is lower than the tank without baffles at high flow rates.
232

Figure B.9. Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for 1.3
` laboratory reactor without baffles and with a turbine.

The high flow rate case with baffles and the low flow rate case without baffles could

Figure B.10. Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for
1.4 ` laboratory reactor with baffles and a turbine.

not be done due to time constrictions.


233

Figure B.11. Comparison of tm /(V /Q) vs. mixer rpm at different flow rate for
1.4 ` laboratory reactor with baffles and an impeller.

B.1.3 CFD Predictions


A model of the stirred tank shown in Figure A.1 was constructed in Fluents
geometry and grid generation tool using a rotating mesh in the region of the impeller
and a fixed mesh elsewhere. The mesh generated contains 626,512 elements. This
grid was then loaded into FluentT M 6.1 for resolution of the fluid flow within the
tank. The standard k turbulent model with standard wall functions was chosen
to predict the flow profile; even in the case of zero rpm, since the k model reduces
to laminar flow model when the energy dissipation rate, , and the turbulent kinetic
energy, k, are zero. The experimental conditions are for the most part in the
transition flow regime (i.e., mixer Reynolds number greater than 10, considered
to be laminar, and less than 10,000 [113], considered to be fully turbulent), see
Table B.2, with one case in the fully turbulent regime and one in the laminar
regime. The simulations were performed for laminar (0 rpm) and transition (20, 40,
80, and 200 rpm) flow regimes. The boundary condition assumed for the flow at the
inlet corresponds to plug flow. The flow rate in the inlet tube is laminar in all cases
for the baffled tank (see Table B.2) and has an average velocity of 0.33 m/s for a feed
234

flow rate of 40 ml/min, and 0.17 m/s for a feed flow rate of 20 ml/min. The 0.17 m/s
average feed velocity is larger than the impeller tip velocity in the 0, 20, 40, and 80
rpm simulations and the 0.33 m/s average feed velocity is larger than the impeller
tip velocity in the 0, 20, and 40 rpm simulations. The tank output was given a
pressure outlet boundary condition with no slip. The walls of the tank, baffles, and
the other tank internals were assigned standard wall function boundary conditions,
the top surface was assigned a symmetry boundary condition, the surface of the
moving zone was assigned an interface boundary condition, and the surface of the
impeller was assigned a (stationary) standard wall function inside the mesh that is
rotating. The model was allowed to run using double precision calculations until
all the scaled residuals reached a value of 104 . This level of convergence took
1, 100 iterations. The resulting velocity profile is given in Figure B.12 and shows
that the steady-state solution contains two major circulation cells - one above and
one below the impeller. The overall flow pattern in each circulation cell is that of

Figure B.12. Velocity vector profile for turbulent flow in the 1.4 ` laboratory
reactor operating at 40 ml/min feed flow rate and a mixer speed of 80 rpm.
235

a helical path on the surface of a torus, circulating from the impeller to the tank
wall, up or down the tank wall (up for the upper circulation cell and down for
the lower circulation cell), back into the center of the tank and into the impeller
again. This overall flow pattern is interrupted by the flow around and behind the
baffles. The flow behind the baffles plays an important role in passing fluid from
the top circulation cell to the bottom circulation cell as there is a minor circulation
cell of cylindrical form behind each baffle in which the material can enter from the
top circulation cell and exit into the bottom circulation cell or vice versa. There
is also some mixing of material between the two circulation cells at the plane of
the impeller. As the flow moves radially out some of the fluid is exchanged from
the upper circulation cell to the lower circulation cell and vice versa. This later
mechanism operates in the tanks both with and without baffles.
The behavior of the tracer is modelled by fixing the fluid flow field and adding a
user defined scalar to model the concentration of tracer with a diffusion coefficient of
105 cm2 /s. No source term was used for the user-defined scalar nor any turbulent
diffusivity. The boundary conditions for the tracer consist of a feed of mass fraction
of 1.0 and an output of whatever concentration is at the outlet tube located at
the top of the tank. All the tank internals use a zero flux boundary condition.
The initial condition for this unsteady simulation was to fill the tank with zero
concentration of tracer and feed the tank with a solution with a mass fraction of
1.0 from the feed tube located just below the impeller. The simulation was allowed
to precede using time steps of 20 s with the same convergence criterion. An overall
mass balance was verified to be accurate to within the convergence criterion. The
convective flux of the tracer at outlet is collected from this simulation and plotted
against time as shown in Figure B.13 where runs for several rpm values at a flow
rate of 40 ml/min are plotted. Here we see that initially the concentration of tracer
is zero and increases with time until the tracer concentration approaches that of the
feed, a value of 1.0. This type of plot is characteristic of the response of a stirred
tank reactor to a step input. The residence time distribution is obtained from this
plot of tracer concentration by differentiating the curve with respect to time, i.e.,
236

Figure B.13. Output concentration as a function of time for the Fluent prediction
of the residence time distribution for various mixer speeds (i.e., 0, 20, 40, 80 and
200 rpm) and a feed flow rate of 40 ml/min.

dC(t)
E(t) = (B.4)
dt

The RTD determined in this way is normalized since the feed tracer concentration
was 1.0. This step input is a different way to determine the RTD than that
used in the experiments, which was a pulse input. In attempting to use a pulse
input method in the simulations, the concentrations were very small and round
off errors were sufficiently large to invalidate the overall mass balance for the
reactor significantly beyond the convergence criterion. Mass balance errors have
been shown to cause drastic errors in RTD measurements [114]. For this reason,
the step method of determining the RTD is used in these simulations.
A comparison of the Fluent-predicted RTD and that measured experimentally
is shown in Figure B.14 for different impeller rpm values for 40 ml/min flow rate
with the baffled tank. The predictions show similar trends in that there is a delay
before the RTD increases after time zero and there is an initial roughness in the
curve. The RTD then decreases exponentially with time. The prediction typically
237

(a)

(b) (c)

Figure B.14. Comparison of experimental result of residence time distribution


function, E(t), experimental measurements, and calculated ideal E(t) for a perfectly
mixed CSTR. (a) 0 rpm, 40 ml/min flow rate, (b) 20 rpm, 40 ml/min and (c) 200 rpm,
40 ml/min. The red line indicates the Fluent simulation, the magenta dot-dash line
indicates the perfectly mixed CSTR and the green line indicates the conductivity
data and the blue dotted line indicates the H+ concentration obtained from the pH
data.

over predicts in the initial spike or two in the first tens of seconds, the experimental
data, and the perfectly mixed theoretical curve.
The predictions for various flow rates and impeller rpm values were analyzed
for the mean residence time and the variance of the RTD. These results (filled
238

data points) are also plotted in Figures B.15 and B.16 for comparison with the
experimental data. The error bars associated with the theoretical points were

Figure B.15. Comparison of experiment and Fluent simulation result for mean
residence time ( ) vs. mixing speed (rpm) at various flow rate for a 1.4 ` laboratory
reactor with baffles and a turbine. The larger filled symbols are from the Fluent
simulation results.

determined by running another simulation of the concentration breakthrough but


with time steps of 10 s. The predictions show good agreement with the experimental
data for all flow rates and rpm values except zero rpm. The error between the
experimental results and predictions is well within the error bars of the two methods
of determining the mean residence time, tm , and the variance () of the RTD.
As a result, the Fluent predictions of the RTD for a stirred tank reactor with a
complex internal geometry operating in transitional turbulent flow has been shown
to give an adequate approximation of experimental measurements when the stirrer
is operating. When the stirrer is operating, the fluid flow profile is caused by the
impeller as shown in Figure B.12. Without the stirrer operating the fluid flow is
driven by the jet of fluid entering the vessel from the feed tube, i.e., the curved tube
that ends just below the impeller that provides a complicated fluid flow profile. The
239

Figure B.16. Comparison of experiment and Fluent simulation result for standard
deviation (/tm ) of RTD vs. mixing speed (rpm) at various flow rate for a 1.4 `
laboratory reactor with baffles and a turbine. The larger filled symbols are for the
Fluent simulation results.

fluid flow profile will depend on the location of the vanes of the stopped impeller
relative to the other tubes inside the tank and this location may change until the
flow becomes stabilized. The location of the impeller vanes with respect to the
feed tube was not noted in the experiments and could not be used for a fixed solid
boundary condition in the CFD prediction. As a result the Fluent CFD prediction
is inaccurate when the impeller is not rotating.
No simulations of the unbaffled tank were performed. We were focusing our
work on the baffled tank since it is most commonly used in industry. Furthermore,
our nonbaffled tank has three inlet tubes in an odd geometry that act as crude and
poorly characterized baffles. In retrospect, due to more deviations in concentration
from an ideal stirred tank, the simulation of the nonbaffled tank may have been a
better choice to validate the CFD predictions of concentration. We did not predict
the pulse concentration with time leaving the reactor in our simulations but the
step concentration. So direct comparisons would not have been possible. In further
240

consideration, the nonbaffled experiments have a larger amount of error associated


with them than the baffled experiments, so that validation with respect to mean
and variance of the residence time would not have been as good as that with the
nonbaffled tank experiments. As a result, we think that, given these considerations,
our work does an adequate job in proving the value of Fluent simulations for the
prediction of residence time distributions in complicated tank geometries operating
in the transitional flow regime.

B.1.4 Conclusion
The mean and variance of the residence time distribution for a stirred tank
deviate from ideal values at low impeller rpm. As the impeller rpm is increased, the
mean and variance of the residence time distribution approach the ideal values. For
the nonbaffled tank ideal behavior is observed at 100 rpm or an mixer Reynolds
number of 3,877 and above, while for the baffled tank ideal behavior is observed
at 80 rpm or an mixer Reynolds number of 2,326 or above using a turbine and
an impeller. With an impeller, a smooth decrease in constant space time with
respect to mixer speed was observed while with a turbine showed fluctuation before
reaching the constant values due to different pattern of mixing. Predictions of the
RTD and its mean and variance using CFD with a k model of turbulence show
good agreement with experiment for transitional turbulent flow regimes in the tank.

B.1.5 Nomenclature
C(t): Concentration as a function of time
E(t): Residence time distribution function
t: time (sec)
Vtank : Volume of the reactor
Q: Flow rate into and out of the reactor (ml/min)
APPENDIX C

OPERATING PROCEDURE

C.1 Experimental Setup


A typical laboratory scale jacketed reactor, with working volume 1.4 ` in size,
equipped with automatic temperature control by circulating a solution of ethylene
glycol/water (1:1 vol%) through the outer jacket of the reactor vessel. PolyscienceT M
digital temperature controller and bath were used to cool or heat the solution,
making it possible to keep the temperature constant within 0.25 C. The feeding
and withdrawing of solution and product from the feed and withdrawal crystallizer,
respectively, are established by peristaltic pumps controlled by OPTO 22T M . The
computer-interfaced experimental set-up is shown in Figure 2.1, see Section 2.4.
Flow rate was obtained by the careful calibration of the speed of pump and
tube size and therefore ensuring the desired feed and withdrawal rates as well
as a constant residence time. To ensure constant solution level in the reactor,
level transmitter was used to monitor solution level and to control the outlet
pump speed while maintaining constant feed flow rate. Also, the OPTO 22 unit
acts as a data acquisition and monitors temperature, pH, conductivity, and ionic
concentration as well as flow rate and mixer speed on-line in every second. All
of these measurements were used to monitor the reaction and the steady state
condition inside the crystallizer. The calibration for all of these systems was
performed before each run to assure measurement accuracy.
The feed concentration of solution was measured beforehand for each of these
experiments since the resultant CSD will depend on the level of supersaturation
at certain temperature. Also, care must be taken so that no crystallization takes
place before the feed entering the crystallizer. This was checked with measuring
242

temperature at the end-tip of feed-point inside the crystallizer. Also, the crystal-
lizer control system has the capability of controlling the feed source and reactor
temperature to 0.25 C, the feed and product flow rate to 0.5 ml/min, liquid
level to 1.5 mm and mixing speed to 3 rpm.
The power input per unit volume, crystallizer geometry and layout should be
carefully considered to achieve the vessel to be well mixed so that powders are
suspended uniformly. An empirical correction for the additional power needed for
homogeneity, proposed by Yamasaki et al. [85], was incorporated to estimate the
minimum power requirement for our 1.4 ` working volume crystallizer vessel with
the configuration of baffles and a turbine or an impeller. The geometry of laboratory
scale crystallizer was shown in Figure A.1, see Appendix A. This figure presented
the strategic placement of all the flow lines and baffles, also the position of mixer.
The feed is injected 4/5 of the way down the crystallizer to aid well mixing before
going out to the outlet at the top surface.
The ratio of liquid level to the crystallizer diameter (H/D) of the vessel is 1.27
which is within the general range of 1.0 to 1.5 for most industrial precipitator [20,
86]. To reduce the energy input to the system while maintaining uniformity, a
standard baffle design [87] is used. The detailed baffling information could be
found in Myers et al. [88]. The dimension of standard baffle was given in Table A.1
along with the dimension of impeller.
The continuous stirred crystallizer is used ubiquitously in the chemical process
industry for mixing, reaction and crystallization. The mixing in a continuous stirred
crystallizer is often not ideal. The RTD is one of the ways to characterize the
non-ideal mixing in the crystallizer. Comparison of the measured RTD with that
of an ideal reactor allows the process engineer to diagnose the ills of the crystallizer
and mixer design. The engineer can then use an appropriate mixing model for
the crystallizer in combination with the kinetics of the reaction to be performed in
the crystallizer to develop an appropriate model for the reactor [91]. Experimental
procedure and result with modeling could be found elsewhere [92].
243

C.2 Preparation of starting materials


A stock solution of CaCl2 , KCl, and NaCl mixture was donated to this project by
the Dow Chemical (Ludington, Michigan). The mixture was at about 62 C when it
was drummed at the Dow Chemical. The sample was unfiltered, unsettled process
liquor from the evaporator train. Each drum containing large amount of crystals at
room temperature can be solubilized by heating to 70 C or above. However, most
of KCl particles cannot be driven back into solution simply by heating because the
process liquor is a slurry of sodium chloride in calcium chloride solution in both
the second and third effects. To ensure complete solubilization of CaCl2 and NaCl,
except for the KCl, the solution temperature was maintained at 70 C or above by
a thermostat controlled drum heater for more than 24 hr.
Element analysis, pH, density, and viscosity measurement of this stock solution
for each drum were performed and shown in Table C.1. The chemical reaction

Table C.1. Analysis of feed solution used for the crystallization of


NaCl.

Element analysis Densitya


Solution No. pHd
Element Dowb (wt%) Utahc (wt%) (g/cm3 )
CaCl2 47.00 47.63
C01 KCl 1.30 1.07 5.99 1.479
NaCl 0.80 0.67
CaCl2 45.23 46.88
C02 KCl 1.33 0.93 6.17 1.475
NaCl 0.63 0.50
CaCl2 45.28 47.09
C03 KCl 1.31 1.03 6.04 1.476
NaCl 0.61 0.44

a
As described in Section 3.1.1.
b
The analysis was done by the Dow Chemical, Ludington, Michigan. The
solution samples were taken at 50 C when the solution was drummed.
c
Solution samples were taken at 70 C after heating the solution more than
24 hrs. Analysis was done by AAS and ICP as described in Section 4.2.3
d
pH of solution was measured at 70 C with a pH probe using temperature
compensation.
244

chosen to study in this work is the purification involving precipitation of NaCl from
the aqueous solution mixture of CaCl2 , KCl and NaCl based upon its characteristic
nucleation and crystal growth mechanism with less aggregation as described in
Section 2.2.
Samples of these starting solutions for the precipitation of NaCl were extracted
in the following manner: First, the density and pH of the starting stock solution were
measured to verify that a representative sample was being taken from the storage
drum (D1). The solution density was measured as described in Section 3.1.1. A 100
ml sample was also taken from this stock solution and stored for future analysis.
These values were monitored in order to maintain daily consistency in the sampling
of the stock solution.

C.3 Operating Flow and Control Diagram


A schematic layout of the experimental system and accessories are shown in
Figure C.1. The solid lines with different arrows show solution and coolant flows.
The air and communication flows are indicated by the dashed lines with arrows
in different end styles to show each flow direction. The lines with arrows in both
directions indicate bidirectional communications for automatic equipment control
and data acquisition.
The major items of equipments in Table C.2 are described according to the flow
sheet in Figure C.1. The operating procedures are discussed in appropriate sections
throughout this chapter and other locations with more details.

C.4 Equipment Calibration Procedure


C.4.1 Thermocouple Calibration
The calibration of thermocouple was performed following the Manual on the
Use of Thermocouples in Temperature Measurement [115] and the Standard Test
Method for Calibration of Thermocouple by Comparison Techniques by by ASTM [116].
245

Table C.2. Equipment list and specification.

Equipment
Equipment Specifications
Symbol
B Thermo NESLABT M EX-510D Bath with Digital Controller
Cole ParmerT M Conductivity Meter with TopacT M 4 Electrodes
C
Epoxy body Platinum plates Conductivity Probe
C1 Preheating Solution Container (4 `)
C2 Preheating Solution Container (4 `)
Continuous Stirred Jacked Tank Reactor (1.4 `) with
CSTR
Turbine and Baffles
D1 Storage Drum (200 `)
D2 Overflow Collection Drum (30 `)
DH Thermostat Controlled Drum Heater (Off to 290 C)
ApplikonT M Mixer operating between 1 1250 rpm
M
with Digital Controller
OPTO 22T M Controller. Reads and Controls Temperature,
OPTO
pH, Conductivity, Level Indicator, Mixer Speed, Pump Speed
P1 Source Refill Pumpa
P2 Preheating Solution Refill Pump
P3 Inlet Solution Refill Pump
P4 Outlet Solution Refill Pump
P5 Rotary Vane Vacuum Pump with 1/2 HP
PC Personal Computer used to commmunicate with and control OPTO 22
OaktonT M pH/mV/Rel mV/ C Benchtop Meter
PH
with Cole-ParmerT M precision pH electrode, epoxy body
PH1 Thermo Heating or Cooling Bath for Crytallizerb
PH2 Thermo Heating or Cooling Bath for PH1
Table continues on next page.

a
P1, P2, P3, and P4 use MasterflexT M Console Analog L/S Pump Drive operating between
6 600 rpm using a Masterflex L/S Easy-LoadT M II Pump Head with NorpreneT M tubing.
b
PH1 and PH2 use PolyScienceT M Digital Temperature Controller with a solution of water
and ethylene glycol mixture controlling temperature between 30 and 130 C.

C.4.2 pH Calibration
pH meter and probe with automatic temperature compensation were calibrated
at 25 C using OrionT M buffer solution pH of 4.00.01, 7.010.01, and 10.000.02,
then incorporation with OPTO measuring DC mV from pH meter once again they
246

Continued from Table C.2.


Equipment
Equipment Specifications
Symbol
S1 Surry Sampling Device
S2 Filtering Device
FlowLine RicochetT M Small Tank Alphasonic Level Transmitter
LT Model LA15 with Signal Output 4 20 mA and 12 36 VDC
Resolution 3 mm
PhysitempT M HT-2 General Purpose Copper-constantan
T1
(type T) Thermocouple
PhysitempT M HT-2 General Purpose Copper-constantan
T2
(type T) Thermocouple
PhysitempT M HT-2 General Purpose Copper-constantan
T3
(type T) Thermocouple
PhysitempT M HT-2 General Purpose Copper-constantan
T4
(type T) Thermocouple
PhysitempT M HT-2 General Purpose Copper-constantan
T5
(type T) Thermocouple
OmegaT M Subminiature Transition Joint Probe, (type T)
T6
Stainless Steel Sheath 6, Ground Junction Thermocouple
OmegaT M Subminiature Transition Joint Probe, (type T)
T7
Stainless Steel Sheath 6, Ground Junction Thermocouple

were calibrated together. A calibration with 99.98% accuracy was obtained by


plotting pH vs. DC mV.

C.4.3 Reactor Level Calibration


The calibration of reactor level indicator (LT) was performed with following its
manual.

1. Connect the indicator into OPTO 22 unit to read the current output.

2. Verity that as the distance from the liquid to the indicator increases, the
current signal decreases.

3. Verity that as the distance from the liquid to the indicator decreases, the
current signal increases.

4. Plot distance (cm) vs. current (mA).


247

A calibration result with 99.99% accuracy was obtained with water as a solution in
a crystallizer without mixing.

C.4.4 Flow Rate Calibration


The careful calibration of flow rate was obtained with 99.93% accuracy by use
of Norprene (LS-14)1 tubing and MasterflexT M Console Analog L/S Pump Drive
using a Masterflex L/S Easy-LoadT M II Pump Head at various pump rpms. All of
pumps are connected to OPTO 22 measuring mV and controlling rpm with mA.

C.4.5 Mixer Speed Calibration


The mixer rpm was calibrated with OPTO 22 measuring mA. The accuracy
of 99.99% was obtained for mixing speed between 0 to 1200 rpm with water as a
solution in a crystallizer.

C.5 Start-up of Equipment Hardware


Each day before starting a experiment a check-list was followed for starting all
the required instruments needed in controlling this experiment. The list reads as
follows (refer to Figure C.1):

1. Before starting any equipments make sure that the crystallizer is filled with
1.4 ` of distilled water.

2. Turn on computer (PC).

3. Turn on OPTO 22T M (OPTO).

4. Turn on Pumps (P1, P2, P3 and P4) and verify that these pumps are all in
the external mode.

5. Fill the Thermo NESLABT M EX-510D bath (B) with distilled water.

6. Turn on Thermo NESLABT M EX-510D bath with digital controller (B) and
set the operating temperature to 75 C.
1
Brand name of Cole-parmer.
248

7. Turn on PolyScienceT M Digital Temperature Controller (PH1) and set the


operating temperature to 70 C.

8. Turn on PolyScienceT M Digital Temperature Controller (PH2) and set the


operating temperature to 20 C. Make sure the connection between PH1
and PH2 is disconnected.

9. Check the temperature of storage drum (D1). It should be in the range of


70 75 C.

10. Empty the water contents of the 1.4 ` crystallizer as soon as the
temperature reaches about 50 C. Refill with 1.4 ` of warm distilled water
and repeat this flushing step two additional times. This cleaning step is to
remove any contaminations as much as possible without dismounting the
crystallizer. On the last flush of the crystallizer leave the water in the
crystallizer and momentarily start the overflow pump (P4) so that the liquid
level is just at the bottom of the overflow pipe.

11. Wait for about 3 hrs to reach the desired temperature for all equipments.

C.6 Crystallization Operation and Sampling


C.6.1 Crystallization Experiment Procedures
A feed solution of CaCl2 , KCl, and NaCl which contains large amount of crystals
at room temperature can be solubilized by heating to 70 C. To ensure complete
solubilization, the solution was maintained at 70 C for more than 24 hr. Then the
solution is feed to the crystallizer which is maintained at 70 C. When the volume
of solution reaches 1400 ml, the temperature is kept for another 30 min. to have
steady state at 70 C and then the reactor is cooled to 25 35 C with cooling rate
of 0.7 1.5 C/min to induce crystallization. The CaCl2 and KCl are soluble at
around 30 C, and NaCl crystallizes out of solution in a CMSMPR crystallizer with
four baffles and a 6-bladed flat disc turbine (Rushton type) or 6-bladed pitch blade
249

impeller in the middle of reactor. A cooling crystallization to purify the CaCl2 is


required of this crystallizer. The operating procedure for crystallization follows as

1. When the desired temperatures for all equipments are reached after the last
steps in section C.5, turn on P1 to transfer solution from D1 to C1.

2. When solution in C1 container reaches same level with solution in B, turn


on P2.

3. When the solution in C2 reaches same level with solution in B, stop P1 and
P2.

4. When T2 reads temperature 70 C or above, empty the water in crystallizer


and turn on P1, P2, and P3.

5. When the solution in crystallizer reaches 1,400 ml, turn on P4 and open V1
valve such that the solution overflows to D2 storage container.

6. Turn on mixer (M) to a desired rpm and wait about 30 min to have steady
state at 70 C.

7. Set the temperature at PH1 to have a desired crystallization temperature


and open the connection between PH1 and PH2 if faster cooling is needed.

The controlling and reading of all equipments can be done with computer-interfaced
experimental set-up as showed in Figure 2.1 and explained in Section 2.5 except for
the valves which require manual controls.

C.6.2 Sampling Procedure at 10


For a typical run, time zero is started when the PolyScienceT M Digital Tem-
perature Controller (PH1) is set to the required temperature for crystallization.
Therefore for a mean residence time of 60 min, 10 can be taken after 600 min from
the time of temperature setting. The sampling process involved for a given value
is as follows (refer to Figure C.1 and Table C.2):
250

1. Make sure that the temperature of slurry sampling device (S1) is kept at the
same as that of crystallization.

2. One minute before the required sampling time, turn on vacuum pump (P5)
and open valve V2 such that the sampling line is ready.

3. At the required sampling time, open valve V3 and take 20 ml slurry


sample for a CSD measurement and then close valve V3. This sample goes
directly to the CSD measurement following the CSD sample preparation
procedure in C.6.3.2.

4. Put a 0.2 m filter (Supor-200 GelmanScienceT M ) in the filtering device


(S2). Turn the direction of V1 valve to S2 and open valve V4. Turn V1 back
to the initial position when the filtrate solution reaches 50 ml and then
close V4 after removing the excess liquid from the filter.

5. The filter sample was labelled CMSMPR mm/dd/yyyy and kept in dry for
further analysis such as SEM, EDX, and XRD. Also the filtrate solution was
labelled CMSMPR mm/dd/yyyy AF (where AF means after filtration) and
kept for ICP and AAS analysis.

6. Take 50 ml solution from the preheating solution container (C2). This


solution container was labelled CMSMPR mm/dd/yyyy BC (where BC means
before crystallization) and kept for further analysis.

7. After finishing all the above steps, follow the shut down procedure in C.7.

C.6.3 Measurement of CSD


The CSD was measured with Beckman Coulter LS230T M and verified for the
shape and size range with optical microscope and scanning electron microscopy
(SEM). The Beckman Coulter LS230T M is a system of multi-functional particle
characterization tool and has an experimental reproducibility of 1%.
251

C.6.3.1 Sampling Procedure


For a CSD measurement, a slurry sample ( 20 ml) was taken directly from the
reactor to the sample bottle with a vacuum pump in a very short time (< 1 s) for
every mean residence time ( ). The sampling process involved for a given value
is as follows (refer to Figure C.1 and Table C.2):

1. During this sampling, to avoid any disturbance to the particles the


temperature of sample device (S1) was kept the same temperature as the
slurry in crystallizer.

2. One minute before the required sampling time, turn on vacuum pump (P5)
and open valve V2 such that the sampling line is ready. Make sure valve V4
is closed.

3. At the required sampling time, open valve V3 and take 50 ml slurry


sample for a CSD measurement and then close valve V3. This sample goes
directly to the CSD measurement following the sample preparation
procedure in C.6.3.2.

4. Place an extra bottle in S1 and lift the sampling tube out from the solution
in the crystallizer and then open valve V3 quickly such a way that the slurry
inside the tube can be removed for the next run.

5. Remove the bottle from S1, empty and flush it with distilled water for
couple of times.

6. Remove any moisture inside the bottle with lint free paper and then place
the bottle in S1 for the next run. This should be done at least 5 min. before
the next sampling so that the temperature of bottle is kept at the same as
that of crystallization.
252

C.6.3.2 Sample Treatment and CSD Measurement Pro-


cedure
High suspension of particles couldnt be handled in LS230 due to multiple
scattering, that is, light scattering off of more than one particle surface. Therefore
special sample preparation is needed for a given value as follows:

1. A saturated solution was prepared by filtrating the slurry with 0.2 m


Supor-200 GelmanScienceT M during the same crystallization experiment
and kept in a water bath at the same temperature as the slurry in
crystallizer. This solution was used to flush the LS230 small volume module
before and after CSD measurement.

2. Add small amount of highly concentrated CdCl2 solution into the saturated
solution making 100 ppm of Cd+ . This solution was used to stabilize
particles during the CSD measurement. The details with underlying theory
for the application of particle stabilization refer to Appendix F [117].

3. When LS230 is ready after the startup procedure (Section E.1), flush two
times with the saturated solution. After the last flush, fill the Cd+ modified
saturated solution ( 150 ml) and follow the sample analysis procedure in
Section E.2. This should be done at least 15 min. before the required time
for every CSD measurements. During this step, pay every attention to
remove bubbles from the solution inside LS230 performing de-bubbling
because bubbles are highly detrimental to the CSD measurement giving a
false result.

4. The slurry sample from step 3 in Section C.6.3.1 was added instantly to the
solution in LS230 right ahead of CSD measurement to obtain an allowable
concentration of particles. Extreme care should be taken to assure that the
temperature for sampling, diluting, and measuring of sample must be the
same as the slurry in the crystallizer at all time.

5. For the sample analysis procedure, refer to Section E.2.


253

C.7 Shut down Procedure


1. After finishing all the sampling steps and crystallization experiments, exit
from the OPTO runtime software.

2. Turn off B, PH1 and PH2.

3. Empty both preheating solution containers (C1 and C2) into the storage
drum (D1) and drain water in B.

4. The reactor contents are emptied into an overflow collection drum (D2) by
using an external tube and a pump.

5. Fill the reactor with distilled water and follow the cleaning step 10 in C.5.

6. Shut down OPTO and PC. Make sure that all hardware are turned off.

7. Transfer the contents from D2 into D1 and keep the solution in D1 for more
than 12 hr for the next run. The temperature of solution in storage drum
(D1) is kept at 70 75 C with the thermostat controlled drum heater (DH).

8. Make sure all the solution flow lines are cleaned with distilled water. Keep
them clean and dry for the next run.
Figure C.1. Operating flow and control diagram for a CMSMPR crystallization system.
254
APPENDIX D

CHARACTERIZATION OF NaCl
CRYSTALS

D.1 Particle Size distributions and Morphology of NaCl


Crystals

Figure D.1. A plot of differential Number % vs. particle size, L (m). After 1
hr from starting CMSMPR crystallizer. Note: scale 10 = 130 m in microscopic
photo.
256

Figure D.2. A plot of population density distribution vs. particle size, L (m).
After 1 hr from starting a CMSMPR crystallizer.

For the experimental number and condition, refer to Table 4.4 through 4.6 in
Section 4.3.3. In 3D plot of PSDs, x = particle diameter (m), y = mean residence
time ( ), and z = differential volume (%). For all experiments the volume of
crystallizer was 1400 ml.

D.1.1 Complete set of PSD plot for TS Samples


The details of experimental condition and data for TS sample are stated in
Tables 4.5 and 5.2 (see Sections 4.3.3 and 5.2). Figures on righthand side are
shown with a view from z-direction.
257

Figure D.3. A plot of differential Number % vs. particle size, L (m). After 6
hr from starting CMSMPR crystallizer. Note: scale 10 = 130 m in microscopic
photo.

D.1.2 Complete set of PSD plot for IS Samples


The details of experimental condition and data for IS sample are stated in
Table 4.6 and 5.3 (see Sections 4.3.3 and 5.2). Figures on righthand side are shown
with a view from z-direction.
258

(a) (b)

(c) (d)

(e) (f)

Figure D.4. 3D plot of PSDs, (a) and (b) for TS 04, (c) and (d) for TS 05, and
(e) and (f) for TS 06.
259

(a) (b)

(c) (d)

33
(e) (f)

Figure D.5. 3D plot of PSDs, (a) and (b) for TS 07, (c) and (d) for TS 08, and
(e) and (f) for TS 09.
260

(a) (b)

(c) (d)

(e) (f)

Figure D.6. 3D plot of PSDs, (a) and (b) for TS 10, (c) and (d) for TS 11, and
(e) and (f) for TS 12.
261

(a) (b)

(c) (d)

(e) (f)

Figure D.7. 3D plot of PSDs, (a) and (b) for TS 13, (c) and (d) for TS 14, and
(e) and (f) for TS 15.
262

(a) (b)

(c) (d)

(e) (f)

Figure D.8. 3D plot of PSDs, (a) and (b) for TS 16, (c) and (d) for TS 17, and
(e) and (f) for TS 18.
263

(a) (b)

(c) (d)

33
(e) (f)

Figure D.9. 3D plot of PSDs, (a) and (b) for TS 19, (c) and (d) for TS 20, and
(e) and (f) for TS 21.
264

(a) (b)

(c) (d)

(e) (f)

Figure D.10. 3D plot of PSDs, (a) and (b) for TS 22, (c) and (d) for TS 23, and
(e) and (f) for TS 24.
265

(a) (b)

(c) (d)

(e) (f)

Figure D.11. 3D plot of PSDs, (a) and (b) for TS 25, (c) and (d) for TS 26, and
(e) and (f) for TS 27.
266

(a) (b)

(c) (d)

(e) (f)

Figure D.12. 3D plot of PSDs, (a) and (b) for IS 04, (c) and (d) for IS 05, and
(e) and (f) for IS 06.
267

(a) (b)

(c) (d)

(e) (f)

Figure D.13. 3D plot of PSDs, (a) and (b) for IS 07, (c) and (d) for IS 08, and
(e) and (f) for IS 09.
268

(a) (b)

(c) (d)

(e) (f)

Figure D.14. 3D plot of PSDs, (a) and (b) for IS 10, (c) and (d) for IS 11, and
(e) and (f) for IS 12.
269

(a) (b)

(c) (d)

(e) (f)

Figure D.15. 3D plot of PSDs, (a) and (b) for IS 13, (c) and (d) for IS 14, and
(e) and (f) for IS 15.
270

(a) (b)

(c) (d)

(e) (f)

Figure D.16. 3D plot of PSDs, (a) and (b) for IS 16, (c) and (d) for IS 17, and
(e) and (f) for IS 18.
271

(a) (b)

(c) (d)

(e) (f)

Figure D.17. 3D plot of PSDs, (a) and (b) for IS 19, (c) and (d) for IS 20, and
(e) and (f) for IS 21.
272

(a) (b)

(c) (d)

(e) (f)

Figure D.18. 3D plot of PSDs, (a) and (b) for IS 22, (c) and (d) for IS 23, and
(e) and (f) for IS 24.
273

(a) (b)

(c) (d)

(e) (f)

Figure D.19. 3D plot of PSDs, (a) and (b) for IS 25, (c) and (d) for IS 26, and
(e) and (f) for IS 27.
APPENDIX E

LS230 SVM PLUS OPERATING


SUMMARY

E.1 STARTUP
It is recommended that the power remain on for day to day operation for LS230
small volume module plus (SVM+).

1. From LS230 Control Program main menu select Run Use Optical Module.

2. Check for suspension fluid at correct level (up to cone).

3. On SVM+ keypad, select On to turn on pump.

4. Wait 2 hours before running a control or sample if the instrument was


powered off before Startup.

5. Change to the directory or folder needed (existing or newly created).

6. Load a Preference file of other than the default.prf file if needed.

7. (Optional) Run a control. For the first run of the day check that:

(a) Offset is within 6 mV for each channel except 127.

(b) Alignment produces a pattern similar to the dfr.$ file in the calfile
folder.

(c) Background is 2106 for each channel.

Power Up
275

(a) Check for correct suspension fluid level.

(b) Turn on the power switches at the LS bench, printer, monitor, and
computer.

(c) Check Use Optical Module and select OK.

E.2 SAMPLE ANALYSIS


1. (De-bubble) Turn pump off, and then push FILL from SVM keypad to
perform routine (takes approximately 30 seconds).

2. Prepare the sample. (If existing, run settings can be loaded for type of
sample: select Run Load Run Setting. . .).

3. From main menu select Run Run Cycle. . . New Sample Start or from
the toolbar select Run Cycle. . . icon then click on New Sample Start.

4. When Measure Loading screen appears, slowly add sample and observing
the title bar for correct obscuration of 8% to 12%.
For PIDS samples, < 0.5 m, check for obscuration of 45 to 55%.
Measure Loading: Obscuration = 0%; PIDS = 0% Low, add sample
Add sample until prompt after % level says OK (8-12%)
The runfile title bar will appear as follows:
Measure Loading: Obscuration = 10%; PIDS = 50% OK

5. From the open RunFile menu select Done.

6. Fill in Sample Info dialog box and select OK.

7. Fill in the Run Info dialog box and select OK.

8. The runs start and continue until the last runs data appears on the screen.

9. Always rinse sample from module when finished with run, and fill with fresh
suspension fluid.

10. To run additional samples that use this same sample and run info:
276

(a) From the main menu select Run Start or from the toolbar select
Start icon.

(b) Repeat steps 3 & 4.

11. After the last sample of the day, perform Shutdown.

E.3 SHUTDOWN
1. On SVM+ keypad, select Off to turn pump off.

2. Close all open RunFiles by clicking on the control menu item until you are
at the LS main menu.

3. Select Run Shut Down Optical Module.

4. Turn off power switch at the monitor.

Power Down

(a) Perform Shutdown procedure.

(b) Click on the control menu icon on the LS main window, and from the
Windows taskbar, select:
Start Shut Down. . . select the radio button for shut down and select
OK.

(c) Turn off the power switches at the LS bench, printer, monitor and
computer.
APPENDIX F

PARTICLE STABILIZATION FOR EX


SITU PSD MEASUREMENT

The measurement of a particle size distribution (PSD) of a soluble salt, NaCl,


dispersed in a saturated solution using laser light scattering (LLS) is fraught with
difficulty since this apparatus and others like it do not have strict temperature
control in the sample chamber. A slight increase or decrease in temperature
produces conditions where the crystals grow or dissolve spontaneously from or into
background solution during PSD analysis. The difficulties introduced in particle size
analysis can be overcome by adding 100 ppm CdCl2 to the saturated solution used
to disperse the crystals for analysis. The Cd2+ ion is known to adsorb on the surface
of NaCl in solution, thus inhibiting the crystal growth/dissolution rate by pinning
steps of NaCl crystals (Kern, T., Growth Cryst. 8, 3, 1969). With this additive
the PSD of NaCl produced during the crystallization from a NaCl contaminated 38
wt% CaCl2 solution cooled from 70 to 30 C in a 1.4 liter jacketed-baffled-stirred
crystallizer is stabilized so that it can be reproducibly measured with the LLS.

F.1 Introduction
Cabrera and Vermilyea [54] hypothesized that strongly absorbing, immobile im-
purities can be adsorbed on terraces of crystal surfaces and pin steps, thus drastically
impeding the step movement relative to that expected without impurities present.
The mechanism of pinning and impeding steps has been widely practiced in many
metal alloys, e.g., precipitation hardening in aluminum copper and aluminum silver
alloys and others [66]. Figure F.1 is a schematic picture showing the qualitative
effect of immobile impurities on step movement. The step movement, and hence,
growth rate, would be halted altogether when the distance (d) between impurities
278

on the surface was < 2rc apart, where rc was the medium radius curvature of a
step. For spacings greater than this, the step may squeeze past the adsorbed species
if the step growth velocity is high enough. The impurity spacing is also called a
critical distance, whose magnitude is approximately given by the Gibbs-Thomson
critical diameter [54, 118].

Figure F.1. A schematic showing the pinning of an advancing step by absorbed


immobile impurities [118]. Top-down view of a growing layer.

Recently, use of atomic force microscopy (AFM) to investigate impurity-step


interactions [57, 119] has shown clear images of the impurity poisoning process of
the steps and proved that below a certain supersaturation (d ) the crystal exhibits
a dead zone where no growth occurs. When the supersaturation () is increased
beyond d , the surface begins to grow with a step speed that is weakly dependent
on . Above a sufficiently high supersaturation ( > d ), the step velocity rises
rapidly, approaching a linear value with respect to . Supersaturation was defined
by /kT = ln(a/ae ), where is the change in chemical potential per mole-
cule upon crystallization, k is Boltzmanns constant, T is the absolute temperature,
and a and ae are the actual and equilibrium solution activities, respectively.
Based on the underlying assumptions of the model, it appears especially suited
to describing systems containing tailor-made additives or otherwise chemically sim-
ilar impurities that strongly interact with the product during crystallization. Such
279

Figure F.2. Dependence of step velocity on supersaturation in the presence of Fe3+


contamination. Measurements were made interferometrically [120] for the 100 face
of KDP. Below a supersaturation given here by d , the crystal exhibits a dead zone
where no growth occurs. As the supersaturation increases beyond d , the surface
begins to grow. Above a sufficiently high supersaturation (denoted here by ),
the step velocity rises rapidly, approaching a linear dependence on supersaturation
(dashed line) that is common for all curves and extrapolates to zero. Similar curves
are obtained for other impurities, but the detailed shape of the curves depends on
impurity type as well as the purity and pH of the starting solution. The results
of AFM measurements fall very close to these curves. For clarity, points a to f are
placed along the curves determined from interferometry. The dotted line shows the
prediction of standard models for impurity pinning of steps [57].

impurities characteristically contain functional groups that bear strong structural


and chemically resemblance to the host crystalline phase, resulting in strong surface
adsorption, high levels of incorporation, and completely blocked growth rate at high
concentration.
Several experimental investigations have been undertaken to clarify the situ-
ation. For example, it was shown that Pb+2 acts as a nucleation agent in a
NaCl system, whereas Co+2 inhibits nucleation in a KNO3 solution. The effect
of the presence of chromium ions on the crystallization of magnesium sulfate was
shown to decrease in both the nucleation and growth rate in increasing chromium
concentration. Although it is difficult to generalize, certain trends can be observed
from the experimental evidence. The inhibiting effect appears to increase with
increasing charge of the cation, e.g., Cr+3 > Fe+3 > Ni+2 > Na+1 . The inhibiting
280

effect also appears to decline above a certain critical impurity concentration. This
was shown in the behavior of the nucleation exponent, which is unaffected at low
contaminant concentration (ppm) but changes erratically at higher concentrations.
The use of 100 ppm CdCl2 was chosen to stabilize NaCl particles since Cd2+
is known to adsorb on the crystal surface [121, 122, 123], pinning and immo-
bilizing the steps in the surface of the NaCl crystal. Cadmium was chosen as
the divalent impurity because its ionic radius is within five percent of that of
sodium [122, 123, 124, 125]. Also the value of is determined not only by impurity
concentration (number of impurity ions per mole of NaCl), but by its type as well.
The strongest hindering action was observed with a smaller ionic radius [120]. In the
solution containing small amounts of Cd2+ and NaCl crystals, the Cd2+ ions enter
the N a+ lattice substitutionally on the surface of NaCl crystals. Since the crystal
must be electrically neutral, each divalent impurity ion replaces two sodium ions
with the result that an excess concentration of positive ion vacancies equal to the
concentration of impurity ion is introduced into the crystal [122]. However, at low
enough temperature the concentration of vacancies is so low that the impurity ions
are mostly unassociated and permanently fixed at their positions in the lattice.
The unassociated impurity ions have no mobility by themselves. However, an
impurity ion may move whenever there is an adjacent vacancy into which it can
jump, although not otherwise.
The use of step pinning Cd2+ additive to stabilize NaCl crystals, to measure PSDs
using LLS, in nearly saturated solutions undergoing temperature variations that
under or super saturate the solution is discussed below.

F.2 Experiments
F.2.1 Crystallization Experiments
A feed solution of (C0 ) = CaCl2 (46.83 wt%), KCl (0.92 wt%), and NaCl
(0.49 wt%) which contains a large amount of crystals at room temperature can
be dissolved by heating to 70 C. To ensure complete dissolution, the solution
was maintained at 70 C for more than 24 hr. Then the solution is fed to the
281

crystallizer and cooled to 30 C to induce crystallization. The CaCl2 and KCl are
soluble at around 30 C, and NaCl crystallizes out of solution in a continuous mixed
suspension, mixed product removal (CMSMPT) crystallizer with four standard
baffles radially-directed around the reactor periphery running the length of the
reactors straight side and a turbine in the middle of reactor. A standard baffle
design and detailed baffling information as well as experimental set-up can be
found elsewhere [87, 88, 92]. The solute concentration of solutions before and
after crystallization was determined by atomic adsorption spectroscopy (AAS)
and inductively coupled plasma-atomic emission spectrometry (ICP-AES). Also
scanning electron microscopy (SEM), energy-dispersive spectrometry (EDS), and
X-ray diffraction (XRD) were performed on solid particles for crystal morphology,
elementary, and phase analyses.

F.2.2 Measurement of Particle Size Distribution (PSD)


The PSD was measured with Beckman Coulter LS230T M and verified for the
particle shape and size range with an optical microscope and scanning electron
microscopy (SEM). The Beckman Coulter LS230T M is a system of multi-functional
particle size characterization tool and has an experimental reproducibility of 1%.
Initial experiments to find an inert liquid in which to disperse the NaCl particles
produced in these crystallization experiments were unsuccessful. Those liquids that
allow for particle stabilization against aggregation were ones that also dissolved the
particles. Those liquids that were inert (e.g., no dissolution) did not disperse the
particles well enough for PSD analysis.
For PSD measurement, a slurry sample ( 10 ml) was taken directly from
the reactor to the sample bottle with vacuum pump in a very short time (< 1
s). The temperature was kept constant during sampling to avoid any change in
the crystals. This sample was diluted in the LS230 to lower the number density
of particles so that light scattering measurement could be made without multiple
scattering. A saturated solution at the temperature of the crystallizer was used
for dilution. It was prepared by filtering the product of the same crystallization
282

experiment with a 0.2 m Supor-200 GelmanScienceT M filter making sure that


the temperature remained constant. Thus the temperature of the solution and
suspension for sampling, diluting, and measuring of sample was kept constant at
all times at 30 C, the same temperature as the slurry in the crystallizer. This is
feasible for all of the steps except measurement in the LS230 where the sample is
heated by 2 to 5 C by the light source and mixer inside the sample chamber. Due
to this heating the measurement of the PSD is not stable over a 10-minute period
as is shown in Figure F.3. The particles tend to dissolve slowly.

Figure F.3. Result of PSD measurements without the addition of CdCl2 . Each of
the measurements was done for 1 min..

To stabilize the PSD measurement cadmium impurities were introduced into


the filtrate solution, which was kept at 30 C by adding precisely determined levels
of Cd2+ ions in distilled water to the saturated solution placed in the LS230 sample
cell before the addition of 10 ml of the slurry. At a concentration of 100 ppm Cd2+
in the final solution, the PSD measurement was stable as shown in Figure F.4. In
Figure F.3 and F.4 each of the PSD measurements was performed for 1 min. with
a 1 min. time interval between measurements. The tri (or quadra) modal form is
due to the nucleation, crystal growth, aggregation and breakage in the crystallizer.
283

Figure F.4. Result of PSD measurements with the addition of CdCl2 . Each of the
measurements was done for 1 min..

Figure F.5. SEM image of particles showing the crystal habit and size after
crystallization at 30 C.

The measured PSD was validated with SEM (Figure F.5) and microscopic photos,
which show detailed particle size and morphology, that do not change with time in
the cadmium modified saturated solution.
284

F.3 Results and Discussion


With the application of Cd2+ impurities to stabilize particles during PSD mea-
surement, there was no appreciable change in the PSD measured repeatedly over a
10-minute period. Table F.1 shows a comparison of particle volume mean diameter
(Dv ) and percentile volume diameters (Dv10 , Dv50 , and Dv90 ) as a function of time to
illustrates more clearly the effects of dissolution on the PSD in Figure F.3 and F.4.
Here we see that there is only a 1.1% change in the mean size for the stabilized
crystals based upon the differential volume size distribution and a 6.8% change in
the mean crystal size for the unstabilized particles over 10 minutes. Note that the
reproducibility of the instrument was 1%.

Table F.1. Comparison of volume statistics (Arith-


metic). Summary of 10 measurements. Calculations
from 0.04 to 2000 m. Dv : volume mean diameter.
Dv10 , Dv50 , and Dv90 : 10, 50, and 90 percentile
volume diameters.

Figure 2 Figure 3
(without Cd2+ ) (with Cd2+ )
0 min. 118 108
Dv (m) 5 min. 108 106
10 min. 103 106
C.V. (%) 6.8 1.1
0 min. 38.9 38.6
Dv10 (m) 5 min. 36.9 38.4
10 min. 36.1 38.3
C.V. (%) 3.8 0.4
0 min. 114 103
Dv50 (m) 5 min. 104 101
10 min. 99.6 101
C.V. (%) 6.7 1.0
0 min. 200 184
Dv90 (m) 5 min. 183 181
10 min. 175 180
C.V. (%) 6.9 1.13
285

Using this experimental protocol with its step immobilization additive and
careful temperature control, accurate particle size distributions can be measured
for materials that are unstable due to crystal growth/dissolution from/to the back-
ground solution of the light scattering apparatus. We have used this experimental
protocol to measure the effluent of a crystallizer for various materials [59]. For
practical purposes, particle stabilization with impurities could be a useful industrial
practice when particle size measurement is required during crystallization if no inert
solvent is available to stabilize the particles. Step immobilizing additives and crystal
habit modifiers for other particle systems are tabulated in Myersons book [89].

F.4 Conclusions
Measuring the PSD produced by crystallizers is often difficult due to the dis-
solution or growth of the particles during analysis in traditional light scattering
instruments. If an inert liquid can not be found into which the particles can be
dispersed the only technique available is to use a saturated solution. Even with
dispersion in a saturated solution, the PSD measured can be effected by dissolution
or growth of the particles due to slight variations in the temperature of the sample
in the instrument. Dissolution and growth of the particles in a slightly under
or over saturated solution can be arrested by the addition of step immobilizing
additives. These additives effectively stabilize the crystals by opening up a window
of saturation where the steps of the surface of the crystal lattice are immobile.
APPENDIX G

THERMODYNAMIC MODEL (OLI


SYSTEMS)

G.1 Outline of Thermodynamic Model


Development of models for electrolyte systems has been an important subject
of research in applied thermodynamics because of the considerable role that elec-
trolytes play in separation processes, environmental applications, production of
energy sources, electrochemical processes, hydrometallurgy and other applications.
Various models for electrolyte solutions have been reviewed by Zemaities et al. [80],
Pitzer [126], Rafal et al. [81], Loehe and Donohue [127] and Anderko et al. [128].
A characteristic feature of electrolyte systems is the fact that phase equilibria
and other thermodynamic properties are often inextricably linked to speciation
equilibria, which may be due to ion pairing, acid-base reactions, complexation, and
other phenomena. For many applications, speciation-related properties such as pH,
oxidation-reduction potential, distribution of complexed or hydrolyzed species, etc.,
are of primary importance. Also, speciation can have a significant effect on phase
equilibria, such as the solubility of salts in multicomponent systems. Recently,
OLI Systems developed a comprehensive thermodynamic model for mixed-solvent
electrolyte systems (Anderko et al. [128], Wang et al. [129]). In the course of this
project, this model was refined to address all the requirements described above.
The fundamentals of the mixed-solvent electrolyte model have been described in
detail by Wang et al. [129]. The basic features of the model are outlined.
The thermodynamic framework has been established by combining an excess
Gibbs energy model for mixed-solvent electrolyte systems with a comprehensive
287

treatment of chemical equilibria. In this framework, the excess Gibbs energy is


expressed as
Gex Gex Gex Gex
= LR + M R + SR (G.1)
RT RT RT RT
where Gex ex
LR represents the contribution of long-range electrostatic interactions, GSR

is the short-range contribution resulting from intermolecular interactions, and an


additional (middle-range) term Gex
M R represents primarily ionic interactions (i.e.,

ion/ion and ion/molecule) that are not accounted for by the long-range term. The
rationale and derivation of Equation G.1 was discussed in detail by Wang et al. [129].
The long-range interaction contribution is calculated from the Pitzer-Debye-
Hckel formula (Pitzer [130, 131]) expressed in terms of mole fractions and symmet-
rically normalized, i.e.,
! !
1/2
Gex
DH
X 4Ax Ix 1 + Ix
= ni ln P  0 1/2
 (G.2)
RT i
i xi 1 + (Ix,i )

where the sum is over all species, Ix is the mole fraction-based ionic strength,
0
Ix,i represents the ionic strength when the system composition reduces to a pure
component i, i.e., i0x,i = 1/2; is related to the hard-core collision diameter ( = 14)
and the Ax parameter is given by
3/2
e2

1
Ax = (2NA ds )1/2 (G.3)
3 40 s KB T

where ds and s are the molar density and the dielectric constant of the solvent,
respectively. The dielectric constant is calculated from an expression developed
previously at OLI Systems for mixed solvents (Wang and Anderko [132]).
For the short-range interaction contribution, the UNIQUAC equation (Abrams
and Prausnitz [133]) is used. The middle-range interaction contribution is calcu-
lated from an ionic strength-dependent, symmetrical second coefficient-type expres-
sion (Wang et al. [129]):
!
Gex
MR
X XX
= ni xi xj Bij (Ix ) (G.4)
RT i i j
288

where Bij (Ix ) = Bji (Ix ), Bii = Bjj = 0 and the ionic strength dependence of Bij is
given by
 p 
Bij (Ix ) = bij = Cij exp Ix + a1 (G.5)

where bij and cij are adjustable parameters and a1 is set equal to 0.01. In general,
the parameters bij and cij are calculated as functions of temperature as
b2,ij
bij = b0,ij + b1,ij T + (G.6)
T
c2,ij
cij = c0,ij + c1,ij T + (G.7)
T
In practice, the middle-range parameters are used to represent ion-ion and ion-
neutral molecule interaction. The short-range parameters are used primarily for
interaction between neutral molecules.
To account for speciation, the chemical effects due to the formation of ion pairs
and complexes are explicitly taken into account using chemical equilibria. Also,
the same chemical equilibrium formalism is used to calculate solid-liquid equilibria.
For this purpose, the standard-state chemical potential 0i needs to be computed
for all species together with the activity coefficients. For solids, the values of 0i are
calculated using the reference-state Gibbs energy of formation, entropy and heat
capacity according to standard thermodynamic relationships. For aqueous species,
the standard-state chemical potentials are calculated as functions of temperature
and pressure using a comprehensive model developed by Helgeson and cowork-
ers (Helgeson et al. [134, 135, 136, 137]), commonly referred to as the Helgeson-
Kirkham-Flowers equation of state. The parameters of this model are available for
a large number of aqueous species including ions, associated ion pairs, and neutral
species (Shock and Helgeson [138, 139, 140], Shock et al. [141, 142]). It should be
noted that standard-state property data and the model of Helgeson et al.are based
on the infinite-dilution reference state and on the molality concentration scale.
At the same time, the activity coefficient model is symmetrically normalized and is
expressed in terms of mole fractions. To make the two reference systems consistent,
the activity coefficients calculated from Equation G.1 are converted to those based
on the unsymmetrical reference state, i.e. at infinite dilution in water, as described
289

by Wang et al. [129]. At the same time, the molality-based standard-state chemical
potential is converted to a corresponding mole fraction-based quantity.
A similar procedure is used for enthalpy and heat capacity calculations. In the
unsymmetrical normalization, the total enthalpy is expressed as
X
h= xi hi = hex, (G.8)
i

where hi is the standard-state partial molar enthalpy of species i, which can be


computed from the Helgeson-Kirkham-Flowers equation and hex, is the excess
molar enthalpy, which is calculated by differentiating the excess Gibbs energy
with respect to temperature and taking into account the change of speciation with
temperature (Wang et al. [129]). Heat capacities are found by differentiating the
enthalpy (Equation G.8) with respect to temperature.

G.2 OLI StreamAnalyzer


Solubility studies can be conveniently performed using StreamAnalyzer, a soft-
ware tool developed by OLI systems. The StreamAnalyzer makes it possible to
make parametric analyses, which can be used to simulate solubility behavior in
multicomponent system and study the effects of temperature, solvents, possible
co-solvents, etc., on solubility. Also, the StreamAnalyzer predicts other thermo-
dynamic properties (such as speciation, pH, activity coefficients, enthalpy, heat
capacity, density, etc.) and transport properties (electrical conductivity, viscosity,
and diffusivity). Current information about the StreamAnalyzer is available on the
website www.olisystems.com.
REFERENCES

[1] T. A. Ring, Fundamentals of Ceramic Powder Processing and Synthesis,


Academic Press, New York, 1996.
[2] A. D. Randolph and M. A. Larson, Theory of Particulate Processes:analysis
and techniques of continuous crystallization, Academic Press, San Diego, 2nd
edition, 1988.
[3] J. W. Mullin, Crystallization, Butterworth-Heinemann, Oxford, 4th edition,
2001.
[4] A. D. Randolph and M. A. Larson, Transient and steady state size distri-
bution in continuous mixed suspension crystallizers, AIChEJ, vol. 8, pp.
639645, 1962.
[5] R. Becker and W. Doring, , Ann. Physik, vol. 24, pp. 719752, 1935.
[6] M. Volmer and A. Weber, , Z. Phys. Chem, vol. 119, pp. 227, 1926.
[7] A. E. Nielsen, Kinetics of Precipitation, vol. 18 of International Series of
Monographs on Analytical Chemistry, Pergamon Press, New York, 1964.
[8] T. A. Ring, J. Dirksen, K. N. Duval, and H. Scheel, Effect of rotation on
surface nucleation in crystal growth, 1998.
[9] J. Garside, , Chem. Eng. Sci., vol. 40, no. 1, pp. 326, 1985.
[10] D. Turnbull and J. C. Fisher, , J. Chem. Phys., vol. 17, pp. 71, 1949.
[11] D. Turnbull, Solid state phys., Science, vol. 112, pp. 448, 1950.
[12] D. Elwell and H. J. Scheel, Crystal growth from High-Temperature Solutions,
Academic Press, New York, 1975.
[13] Zeldovich, .
[14] H. Schubert, PhD thesis, Technische Universitat Munchen, 1998.
[15] A. Eble, PhD thesis, Technische Universitat Munchen, 2000.
[16] C. Heyer, PhD thesis, Technische Universitat Munchen, 2001.
[17] G. D. Botsaris, Secondary nucleation: a review, in Industrial Crystalliza-
tion (6th Symposium, Usti nad Labem), J. W. Mullin, Ed., New York, 1976,
pp. 322, Plenum Press.
291

[18] J. Estrin, Preparation and Properties of Solid State Materials, vol. 2, Dekker,
New York, 1976.
[19] J. Garside and R. J. Davey, Secondary contact nucleation:kinetics, growth
and scale-up, Chem. Eng. Commun., vol. 4, pp. 393424, 1980.
[20] A. D. Randolph, J. R. Bechman, and K. Kraljevich, Crystal size distribution
dynamics in a classified crystallizer: Part I. experimental and theoretical
study of cycling in a potassium chloride crystallizer, AIChE J., vol. 23, no.
4, pp. 500510, 1977.
[21] B. C. Shah, W. L. McCabe, and R. W. Rousseau, Polyethylene vs. stainless
steel impellers for crystallization processes, AIChE J., vol. 19, no. 1, pp.
194196, 1973.
[22] J. Garside and M. B. Shah, Crystallization kinetics from msmpr crystalliz-
ers, Ind. Eng. Chem. Process. Des. Dev., vol. 19, pp. 509514, 1980.
[23] C. Gahn and A. Mersmann, Brittle fracture in crystallization processes:
Part a. attrition and abrasion of brittle solids, Chem. Eng. Sci., vol. 54, no.
9, pp. 1273, 1999.
[24] C. Gahn and A. Mersmann, Brittle fracture in crystallization processes:
Part b. growth of fragments and scale-up of suspension crystallizers, Chem.
Eng. Sci., vol. 54, no. 9, pp. 1283, 1999.
[25] A. F. Wells, Crystal growth, in Annual Reports on the progress of
Chemistry, London, 1946, The Chemical Society, vol. 43, pp. 6287.
[26] H. E. Buckley, Crystal Growth, Chapman and Hall, London, 1952.
[27] R. F. Strickland-Constable, Kinetics and Mechanism of Crystallization,
Academic Press, London, 1968.
[28] B. Lewis, Nucleation and growth theory, chapter In Crystal Growth, pp.
2363, Pergamon Press, Oxford, 1980.
[29] A. A. Chernov, Modern Crystallography, Springer-Verlag, Berlin, 1980.
[30] A. A. Chernov, Formation of crystals in solution, Contemporary Physics,
vol. 30, pp. 251276, 1989.
[31] J. N`
y vlt, O. Sohnel, M. Matuchova, and M. Broul, The Kinetics of Industrial
Crystallization, Elsevier, Amsterdam, 1985.
[32] D. E. Temkin, Crystallization Processes, Consultant Bureau, New York,
1964, p.15.
[33] K. A. Jackson, Liquid Metals and Solidification, Am. Soc. Metals, Cleveland,
1958, p.174.
292

[34] J. W. Gibbs, Thermodynamics, vol. I. Yale University Press, New Haven,


1948.
[35] M. Volmer, Kinetics of Phase Formation, Theodor Steinkoff, Dresden,
Germany, 1939.
[36] J. A. Dirksen and T. A. Ring, Fundamentals of crystallization:kinetic effects
on particle size distributions and morphology, Chem. Eng. Sci., vol. 46, no.
10, pp. 23892427, 1991.
[37] M. OHara and R. C. Reid, Modelling Crystal Growth Rates from Solution,
Prentice-Hall, Englewood Cliffs, 1973.
[38] J. P. Van der Eerden, P. Bennema, and T. T. Cherepanova, , Prog. Crystal
Growth Characterization, vol. 1, pp. 219, 1978.
[39] W. K. Burton, N. Cabrera, and F. C. Frank, , Phil. Trans. Roy. Soc., vol.
A243, pp. 299358, 1951.
[40] G. Springholz and K. Wiesauer, Nanoscale dislocation patterning in
pbte/pbse (001) heteroepitaxy, Physical Review Letters, vol. 88, pp. 015507
1/4, 2002.
[41] A. R. Verma, Crystal growth and Dislocations, Butterworth, London, 1953.
[42] A. J. Forty, , Phil. Mag., vol. 42, pp. 670, 1951.
[43] F. C. Frank, , Disc. Fraday Soc., vol. 5, pp. 48, 1949.
[44] P. Bennema and G. H. Gilmer, Crystal Growth, North Holland, Amsterdam.
[45] N. Cabrera and M. M. Levine, , Phil. Mag., vol. 1, pp. 450, 1956.
[46] A. A. Chernov, , Sov Phys Uspeki, vol. 4, pp. 116, 1961.
[47] A. A. Noyes and W. R. Whitney, Rate of solution of solid substances in
their own solution, J. Am. Chem. Soc., vol. 19, pp. 930934, 1897.
[48] H. A. Miers, The concentration of the solution in contact with a growing
crystal, Phil. Trans., vol. A202, pp. 492515, 1904.
[49] A. Berthoud, Theorie de la formation des faces d,un crystal, Journal de
Chimique Physique, vol. 10, pp. 624635, 1912.
[50] J. J. P. Valeton, Wachstum und aufl osung der kristalle, Zeitschrift f
ur
Kristallographie, vol. 59, pp. 483, 1924.
[51] E. Sobczak, A simple method of determination of mass transger coefficients
and surface reaction constant for crystal growth, Chemical Engineering
Science, vol. 45, pp. 561564, 1990.
293

[52] J. Garside, The concept of effectiveness factors in crystal growth, Chem.


Engng. Sci., vol. 26, no. 9, pp. 14251431, 1971.
[53] J. Garside and N. S. Tavare, Non-isothermal effectiveness factors for crystal
growth, Chem. Eng. Sci., vol. 36, no. 5, pp. 863866, 1981.
[54] N. Cabrera and D. Vermilyea, The Art and Science of Growth Crystals,
chapter Growth and Perfection of Crystals: The Growth of Crystals from
Solution, pp. 393410, Wiley, New York, 1958.
[55] E. N. Slavnovna, , Growth of Crystals, vol. 1, pp. 117, 1958.
[56] E. N. Slavnovna, , Growth of Crystals, vol. 2, pp. 166, 1958.
[57] T. A. Land, T. L. Martin, S. Potapenko, G. T. Palmore, and J. J. De
Yoreo, Recovery of surfaces from impurity poisoning during crystal growth,
Nature, vol. 399, pp. 442445, 1999.
[58] G. W. Sears, , J. Chem. Phys., vol. 29, pp. 104, 1958.
[59] T. A. Ring, J. A. Dirksen, K. N. Duval, and N. Jongen, LiBr2H2 O
crystallization inhibition in the presence of additives, J. colloid Interface
Sci., vol. 239, pp. 399408, 2001.
[60] A. Mersmann, M. Angerhofer, R. Gutwald, R. Sangl, and S. Wang, , Sep.
Technol., vol. 2, pp. 8597, 1992.
[61] C. F. Abegg, J. D. Stevens, and M. A. Larson, Crystal size distributions in
continuous crystallizers when growth rate is size-dependent, AIChE J., vol.
14, no. 1, pp. 118122, 1968.
[62] Z. Rojkowski, Two parameter kinetic equation of size dependent crystal
growth, Kristall und Techmik, vol. 13, pp. 12771284, 1978.
[63] S. J. Jan
ci
c and P. A. M. Grootscholten, Industrial Crystallization, Univer-
sity Press, Delft, 1984.
[64] C. W. Bunn, Chemical Crystallography, Oxford University Press, London,
2nd edition, 1961.
[65] B. D. Cullity, Elements of X-Ray Diffraction, Addison-Wesley Publishing
Co., Reading, MA, 2nd edition, 1978.
[66] D. A. Porter and K. E. Eastering, Phase Transformations in Metals and
Alloys, Chapman and Hall, New York, 2nd edition, 1992.
[67] R. E. Reed-Hill, Physical Metallurgy Principles, Van Nostrand, New York,
2en edition, 1973.
[68] B. Chalmers, Principles of solidification, Wiley, New York, 1964, p.116.
294

[69] M. V. Smoluchowski, Mathematical theory of the kinetics of the coagulation


of colloidal solutions, Z. Phys. Chem., vol. 92, pp. 129, 1917.
[70] K. A. Kusters, The Influence of Turbuluence on Aggregation of Small
Particles in Agitated Vessels, PhD thesis, Technical University Eindhoven,
The Netherlands, p.19.
[71] H. M. Ang and J. W. Mullin, Crystal growth rate determinations from
desupersaturation measurements: Nickel ammonium sulphate hexahydrate,
Trans. Inst. Chem. Eng., vol. 57, no. 4, pp. 237243, 1979.
[72] K. C. Lim, M. A. Hashim, and B. Sen Gupta, The effect of volume shape
factor on the crystal size distribution of fragments due to attrition, Cryst.
res. Technol., vol. 34, no. 4, pp. 491502, 1999.
[73] A. Dey and B. Sen Gupta, Effects of growth rate dispersion in industrial
crystallizers. a classical approach, Cryst. Res. technol., vol. 28, no. 7, pp.
937944, 1993.
[74] K. Sedlatschek and L. Bass, , Powder Metall. Bull., vol. 6, pp. 148153,
1953.
[75] L. Austin, K. Shoji, V. Bhatia, V. Jindal, K. Savage, and R. Klimpel, Some
results on the description of size reduction as a rate process in various mills,
Ind. Eng. Chem. Process Des. Dev., vol. 15, no. 1, pp. 187196, 1976.
[76] L. G. Austin, R. R. Klimpel, and A. N. Beattie, Design and Installation of
Communinution Cricuits, AIME, New York, 1982.
[77] C. L. Prasher, Crushing and Grinding Process Handbook, Wiley, Chickch-
ester, 1987, p.282.
[78] R. A. Broadbent and T. G. Callcott, , J. Inst. Fuel, vol. 29, pp. 524539,
1956.
[79] R. R. Klimpel and L. G. Austin, , Ind. Eng. Chem. Fund., vol. 9, pp.
230237, 1956.
[80] J. F. Jr. Zemaitis, D. M. Clark, M. Rafal, and N. C. Scrivner, Handbook of
Aqueous Electrolyte Thermodynamics, AIChE, New York, 1986.
[81] M. Rafal, J. W. Berthold, N. C. Scrivner, and S. L. Grise, Models for
Electrolyte Solutions in Models for Thermodynamic and Phase Equilibria
Calculations, M. Dekker, New York, 1994.
[82] A. F. Izmailov and A. S. Myerson, Diffusion in supersaturated solutions:
Application to the case of supersaturated protein solutions., Journal of
Chemical Physics, vol. 112, no. 9, pp. 43574364, 2000.
[83] M. M. Lencka, A. Anderko, S. J. Sanders, and R. D. Young, , Int. J.
Thermophysics., vol. 19, pp. 367, 1998.
295

[84] J. A. Dirksen, The Precipitation of Basic Zirconium Sulfate in a Continuous


Stirred Tank Reactor as a Precurson to Zirconia Ceramics, PhD thesis, Swiss
Federal Institute of Technology Lausanne (EPFL), 1992.
[85] H. Yamasaki, K. Tojo, and K. Miyanami, Effect of power consumption on
solids concentration profiles in a slurry mixing tank, Powder Technology,
vol. 64, pp. 199206, 1991.
[86] S. M. Walas, Rules of thumb, Chemical Engineering, pp. 7581, March 16
1987.
[87] R. L. Bates, Impeller Characteristics and Power, Chapter 3 in Mixing:
Theory and Practice, Academic Press, New York, 1966.
[88] K. J. Myers, M. F. Reeder, and J. B. Fasano, Optimize mixing by using the
proper baffles, CEP, pp. 4247, February 2002.
[89] A. S. Myerson, P. Y. Lo, Y. C. Kim, and R. Ginde, In proceedings of the 11th
Symposium on Industrial Crystallization, p. 847, European Fed. of Chemical
Engineers, Munich, Germany, 1990.
[90] R. M. Ginde and A. S. Myerson, , AIChE Symposium Series, vol. 87, no.
284, pp. 124129, 1991.
[91] H. Scott Fogler, Elements of Chemical Reaction Engineering, Prentice Hall
PTR, New Jersey, 3rd edition, 1999.
[92] B. S. Choi, B. Wan, S. Philyaw, K. Dhanasekharan, and T. A. Ring, Res-
idence time distributions in a stirred tank comparison of cfd predictions
with experiment, Industrial and Engineering Chemistry Research, vol. 43,
no. 20, pp. 65486556, 2004.
[93] O. Levenspiel, Chemical Reaction Engineering, John Wiley and Sons, New
York, 2nd edition, 1972.
[94] T. Allen, Particle Size Measurement, 3rd edition, 1981.
[95] A. S. Myerson, Handbook of Industrial Crystallization, Butterworth-
Heinemann, Boston, 2nd edition, 2002.
[96] K. A. Ramanarayaran, K. A. Berglund, and M. A. Larson, Growth kinetics
in the presence of growth rate dispersion from batch crystallizers, Chem.
Eng. Sci., vol. 40, pp. 16041608, 1985.
[97] J. R. Bourne and M. Zabelka, The influences of gradual classification on
continuous crystallization, Chem. Eng. Sci., vol. 35, pp. 533542, 1980.
[98] P. F. Liao and H. M. Hulburt, Agglomeration processes in suspension
crystallization, AIChE Meeting, Dec. 1976.
296

[99] M. Nicmanis and M. J. Hounslow, Finite-element methods for steady-state


population balance equations, AIChE J., vol. 44, no. 10, pp. 22582272,
1998.
[100] V. Pavelic and U. Saxena, Basics of statistical experimentdesign, Chem-
ical Engineering, October 6 1969.
[101] E. T. White and A. D. Randolph, Graphical solution of the material balance
constraint for msmpr crystallizers, AIChE J., vol. 33, no. 4, pp. 686689,
1987.
[102] S. J. Khang and O. Levenspiel, New scale-up and design method for stirrer
agitated batch mixing vessels., Chem. Eng. Sci., vol. 31, no. 7, pp. 569577,
1976.
[103] P. Zaloudik, Mixed model for continuous stirred tank reactor with viscous
fluid from experimental age distribution study., React. Eng., vol. 14, no. 5,
pp. 657659, 1969.
[104] A. Gianetto and F. Cazzulo, Continuous-flow mixers: working efficiency
and conditions for attaining scale-up; Notes I, II and III, Sci. Chim., vol.
38, no. 4, pp. 322342, 1968.
[105] J. C. R. Turner, The interpretation of residence-time measurements in
systems with and without mixing, Chem. Eng. Sci., vol. 26, pp. 549557,
1971.
[106] O. Levenspiel, B. W. Lai, and C. Y. Chatlynne, Tracer curves and the
residence time distribution., Chem. Eng. Sci., vol. 25, pp. 16111613, 1970.
[107] O. Levenspiel and J. C. R. Turner, The interpretation of residence-time
experiments., Chem. Eng. Sci., vol. 25, pp. 16051609, 1970.
[108] F. J. Beltran, M. Gonzalez, B. Adedo, and C. Gomez, Kinetic modeling of
aqueous trichloroethylene direct ozonation in a continuous tank., J. Environ.
Sci. Health Part A, vol. 32, no. 910, pp. 24712482, 1997.
[109] A. Amanulla, C. M. McFarlane, A. N. Emery, and A. W. Nienow, Scale-
down model to simulate spatial pH variations in large-scale bioreactors,
Biotechnol. Bioeng., vol. 73, no. 5, pp. 390399, 2001.
[110] K. T. Li and J. D. Lee, Mixing and control of a cstr with series-parallel
reactions., Chin. Inst. Chem. Eng., vol. 22, no. 2, pp. 6169, 1991.
[111] B. Newell, J. Bailey, A. Islam, L. Hopkins, and P. Lant, Characterizing
bioreactor mixing with residence time distribution tests., Water Sci. Tech-
nol., vol. 37, no. 12, pp. 4347, 1998.
[112] J. L. Johnson and L. T. Fan, Observation concerning pulse testing of flow
systems., AIChE J., vol. 12, pp. 1026, 1966.
297

[113] R. H. Perry and C. H. Chilton, Chemical Engineers Handbook, McGraw-Hill,


New York, 5th ed. edition, 1973.
[114] R. L. Curl and M. L. McMillan, Accuracy in residence time measurements,
AIChE J., vol. 12, no. 4, pp. 819822, 1966.
[115] ASTM, Philadelphia, PA, Manual on the Use of Thermocouples in Temper-
ature Measurement, 1993.
[116] Standard Test Method for Calibration of Thermocouple by Comparison Tech-
niques, E220, vol. 14, Annual Book of ASTM Standards, 2001, pp. 98109.
[117] B. S. Choi and T. A. Ring, Stabilizing nacl particles with Cd2+ in a saturated
solution during ex situ psd measurement, J. Cryst. Growth, vol. 269, no.
24, pp. 575579, 2004.
[118] J. P. van der Eerden and H. Muller-Krumbhaar, Formation of macrosteps
due to time dependent impurity adsorption, Electrochimica Acta, vol. 31,
no. 8, 1986.
[119] A. J. Malkin, Yu. G. Kuznetsov, and A. McPherson, In situ atomic force
microscopy studies of surface morphology, growth kinetics, defect structure
and dissolution in macromolecular crystallization, J. Cryst. Growth., vol.
196, pp. 471488, 1999.
[120] L. N. Rashkovich and N. V. Kronsky, Influence of Fe3+ and Al3+ ions on
the kinetics of steps on the {100} faces of KDP, J. Cryst. Growth., vol. 182,
pp. 434441, 1997.
[121] R. J. Davey, Adsorption of impurities at growth steps, J. Cryst. Growth,
vol. 29, no. 2, pp. 212214, 1975.
[122] H. W. Etzel and R. J. Maurer, The concentration and mobility of vacancies
in sodium chloride, J. Chem. Phys., vol. 18, pp. 10031007, 1950.
[123] R. Kern, Crystal growth and adsorption, Growth Cryst., vol. 8, pp. 323,
1969.
[124] C. Kittel, Introduction to Solid State Physics, Wiley, New York, 1996.
[125] A. R. Allnatt and M. H. Cohen, Statistical mechanics of defect-containing
solids. II. ionic crystals, J. Chem. Phys., vol. 40, pp. 18711890, 1964.
[126] K. S. Pitzer, Activity Coefficients in Electrolyte Solution, CRC Press, Boca
Raton, FL, 2nd edition, 1991.
[127] J. R. Loehe and M. D. Donohue, , AIChe J., vol. 43, pp. 180, 1997.
[128] A. Anderko, P. Wang, and M. Rafal, Scale-down model to simulate spatial
ph variations in large-scale bioreactors, Fluid Phase Equilibria., vol. 123,
no. 5, pp. 194197, 2002.
298

[129] P. Wang, A. Anderko, and R. D. Young, , Fluid Phase Equilibria., vol. 141,
pp. 203, 2002.
[130] K. S. Pitzer, Thermodynamics of electrolytes. i. theoretical basis and general
equations, J. Phys. Chem., vol. 77, no. 2, pp. 268277, 1973.
[131] K. S. Pitzer, Electrolytes. from dilute solutions to fused salts, J. Am.
Chem. Soc., vol. 102, no. 9, pp. 29022906, 1980.
[132] P. Wang and A. Anderko, , Fluid Phase Equilibria., vol. 186, pp. 103, 2001.
[133] D. S. Abrams and J. M. Prausnitz, , AIChE J., vol. 21, pp. 116, 1975.
[134] H. C. Helgeson, D. H. Kirkham, and G. C. Flowers, , Amer. J. Sci., vol.
274, pp. 1089, 1974.
[135] H. C. Helgeson, D. H. Kirkham, and G. C. Flowers, , Amer. J. Sci., vol.
274, pp. 1199, 1974.
[136] H. C. Helgeson, D. H. Kirkham, and G. C. Flowers, , Amer. J. Sci., vol.
276, pp. 97, 1976.
[137] H. C. Helgeson, D. H. Kirkham, and G. C. Flowers, , Amer. J. Sci., vol.
281, pp. 1241, 1981.
[138] E. L. Shock, H. C. Helgeson, and D. A. Sverjensky, , Geochim. Cosmochim.
Acta, vol. 52, pp. 2009, 1988.
[139] E. L. Shock, H. C. Helgeson, and D. A. Sverjensky, , Geochim. Cosmochim.
Acta, vol. 53, pp. 2157, 1989.
[140] E. L. Shock, H. C. Helgeson, and D. A. Sverjensky, , Geochim. Cosmochim.
Acta, vol. 54, pp. 915, 1990.
[141] E. L. Shock, D. C. Sassani, M. Willis, and D. A. Sverjensky, , Geochim.
Cosmochim. Acta, vol. 61, pp. 907, 1997.
[142] D. A. Sverjensky, E. L. Shock, and H. C. Helgeson, , Geochim. Cosmochim.
Acta, vol. 61, pp. 1359, 1997.
INDEX
Activity BCF model, 35
in equilibrium with the condensate, chemical reaction, 45
12 continuous growth, 25
of the condensing species, 11 diffusion-reaction theories, 40
Aggregation kernel, 58 impurity effect, 46
size-independent, 0 , 155 lateral growth, 25
Aggregation rate, Ka , 157 surface energy theories, 27
Atomic absorption spectroscopy (AAS), Crystal shape, 53
79 Crystallization, 1
Attrition, 192 Crystallization temperature, TC , 157
Crystallography, 53
CMSMPR, 1
CSTR, 4
Computational fluid dynamics (CFD),
Cumulative number oversize distribu-
94
tion, N (L), 190
Contact angle, 16
Contact nucleation rate, 19 Damkohler number, 50
Conversion factor for crystal growth, Da, 46
surface area, A , 12 Density, 76
volume, V , 12 suspension density, 4
Correlation analysis, 164 Diffusion
Crystal coefficient, 50n
habit, 54 EDS, 88
population density, 5 Effectiveness factor, 45
size distribution for crystal growth rate, c , 45
CSD, 3 Element analysis
Crystal growth kinetics, 20 of solid, 88, 107
adsorption layer theories, 27 of solution, 79, 97
300

Embryo mixing speed, M S, 157


concentration, 13 Molar
Equilibrium number density enthalpy, 2
of embryo, 14 entropy, 2
Gibbs free energy, 1
Factorial design, 68
Moments of size distribution, 8
analysis, 167
Multiple linear regression, 161
feed flow rate, Q, 157
model, 184
Free energy change
three factor interaction, 163
per embryo, 11
of aggregate, 11 Nucleation, 9
critical size, r , 12
Gibbs free energy, 12
kinetic order of, 8
energy barrier, 12
primary, 9
of a critical cluster, Gmax , 12
heterogeneous, 16
Growing crystal structure
homogeneous, 10
kink site, 21
secondary, 10
step site, 21
Nucleation rate, B0 , 7, 157
Growth rate, G, 7, 157
Arrhenius type of, 11
kinetic order of, 8
Nucleation theory, 10
Hounslow equation, 154 embryo, 10
Gibbs-Kelvin equation, 10
Ionic crystals, 53
Nuclei
Linear regression analysis, 161 standard critical size of, 13
Number density distribution
McCabes L law, 7
exponential, 8
Mean retention time, 7
Number of nuclei, n0 , 157
Mean size of particles (Lmean ), 192
Mechanical stress, 192 Particle number density, n, 156
Mie theory, 90 Particle number fraction, nL , 156
Mixer power number, 226 Particle size, L, 190
301

Population balance equation atomically flat close-packed inter-


homogeneous nucleation rate, 14 face, 23
PBE, 3 Surface nucleation, 28
Population density, n(L), 190 population balance equation, 31
Precipitation point analysis, 82 Suspension density, MT , 191
Pycnometer, 76
Total mass of particles, MT , 156
Reactor steady state, 84 Two-dimensional Growth, 32
Residence time distribution, 94 birth and spread, 34
RTD, 80 mononuclear, 32
Reynolds Number, 225, 226 polynuclear, 33
Ripening, 13
Viscometer, 78
Saturation ratio, 3, 11 capillary viscometer, 78
Secondary nucleation, 17, 197 rotational viscometer, 78
apparent nucleation, 17 Viscosity, 77
attrition fragment, 154 dynamic viscosity, 79
collision frequency, 154 kinematic viscosity, 78
contact nucleation, 17 Newtons law of, 78
effective rate, Na,ef f , 154 non-Newtonian fluid, 78
true nucleation, 17
XRD, 91
Shape factor, 54
Bragg equation, 91
area shape factor (A ), 54
volume shape factor (V ), 54
Solid
amorphous, 53
crystalline, 53
noncrystalline, 53
Specific gravity, 76
Structure of crystal surface, 23
atomically diffuse interface, 24

Anda mungkin juga menyukai