Anda di halaman 1dari 250

S D D

P F S T

Christopher Samuel Handscomb


Kings College

November, 2008

This dissertation is submitted for the degree of Doctor of Philosophy

Summary
Title: Simulating Droplet Drying and Particle Formation in Spray Towers
Author: Christopher Samuel Handscomb
This thesis presents a new modelling framework for the simulation of single droplet
drying. Focussing on droplets containing dissolved solids in an ideal binary solution with
additional suspended solids, the general model framework tracks the spatially resolved continuous phase composition using volume averaged transport equations and employs a population balance to describe nucleation, movement and crystallisation of the solid particles.
Beyond describing the rate of moisture loss from drying droplets, the key achievement
of the new model is the ability to describe structural development as influenced by evolving
droplet composition and drying conditions and, thereby, the ability to simulate multiple
dried-particle morphologies. This is achieved by combining various sub-models with the
core droplet description to produce an integrated simulation of structural evolution. The
thickening, wet shell, dry shell and slow boiling sub-models are developed, thus allowing
the simulation of morphological developments associated with shell and bubble formation
and growth. Further, extensive consideration is given to the physical motivation of these
sub-models, permitting the development of rational criteria for deciding which sub-model
to use at different stages in a droplets drying history. It is the development of such criteria
which allows, for the first time, structural evolution to be simulated as a function of evolving
droplet composition and external drying conditions.
The new droplet drying model, expressed as a coupled system of ordinary and partial
differential equations, is solved using NAG library routines. The computational and numerical issues associated with the model solution are discussed in depth, giving confidence
in the results obtained. Several systems of practical interest colloidal silica, sodium sulphate solution and detergent crutcher mix are simulated using the new model and the
results obtained are validated against experimental data from the literature. Droplet mass
and temperature profiles are accurately reproduced and the ability of the model to simulate
structural evolution during drying is demonstrated and investigated.

Declaration
This dissertation is the result of my own work and includes nothing which is the outcome
of work done in collaboration, except where specifically indicated in the text. The work
presented was undertaken at the Department of Chemical Engineering at the University of
Cambridge, UK, between September 2004 and October 2008. Chapters 2 and 3 of this
dissertation include work from the dissertation I submitted in August 2005 for a Certificate
of Postgraduate Study. No other part of this thesis has been submitted for a degree to this
or any other university. This dissertation contains approximately 59 000 words and 87
figures.
Some of the work in this dissertation has been published:
1. C. S. Handscomb, M. Kraft and A. E. Bayly. A new model for the drying of
droplets containing suspended solids. Chemical Engineering Science, in press, 2008.
doi:10.1016/j.ces.2008.04.051
2. C. S. Handscomb, M. Kraft and A. E. Bayly. A new model for the drying of droplets
containing suspended solids following shell formation. Chemical Engineering Science,
in press, 2008. doi:10.1016/j.ces.2008.10.019

Christopher Samuel Handscomb


December 13, 2008

Acknowledgements
I would like to acknowledge my supervisor, Dr Markus Kraft along with all members of the
Computational Modelling Group. Funding for this project was provided by the EPSRC
through the Smith Institute for Industrial Mathematics in collaboration with Procter and
Gamble; I should like to thank Dr Andrew Bayly from P&G and Dr Melvin Brown from
the Smith Institute for the help and guidance they provided over the course of my PhD.

ii

Quickly, bring me a beaker of wine, so that I may wet my mind and say something clever.
Aristophanes (456386BCE)

iii

Contents
1 Introduction
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Novel Elements of This Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Structure of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
1
2
4

2 Background
2.1 Drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1 Moisture in Gases and Solids . . . . . . . . . . . . . . . . . .
2.1.2 The Driving Force for Drying . . . . . . . . . . . . . . . . .
2.1.3 Stages of Drying . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.4 Industrial Drying . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Spray Drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Process Overview . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2 Applications of Spray Drying . . . . . . . . . . . . . . . . . .
2.3 Spray Dryer Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1 Heat and Mass Balances . . . . . . . . . . . . . . . . . . . . .
2.3.2 Equilibrium Based Models . . . . . . . . . . . . . . . . . . .
2.3.3 Rate Based Models using Simplifying Flow Assumptions
2.3.4 Rate Based Models Using CFD . . . . . . . . . . . . . . . .
2.4 Single Droplet Drying Models . . . . . . . . . . . . . . . . . . . . . .
2.4.1 Droplet Drying Behaviour . . . . . . . . . . . . . . . . . . .
2.4.2 Droplet Morphologies . . . . . . . . . . . . . . . . . . . . . .
2.4.3 Droplet-Averaged Drying Models . . . . . . . . . . . . . . .
2.4.4 Mechanistic Models . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Scope of this Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

5
5
6
7
8
11
11
11
14
15
16
18
19
20
22
22
24
27
28
32

3 A New Model for Drying Droplets Prior to Shell Formation


3.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1 Descriptions of Moisture Movement in Porous Media
3.1.2 Local Volume Averaging . . . . . . . . . . . . . . . . . . .
3.2 Overview of the New Model . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Model System . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Continuous Phase Description . . . . . . . . . . . . . . . . . . . .
3.3.1 Defining the Solids Volume Fraction . . . . . . . . . . .
3.3.2 Source Terms . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.3 Transport Term . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

33
33
33
37
43
43
46
48
48
49
49

iv

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

3.4

3.5

3.6
3.7

3.3.4 The Continuous Phase Equation . . . . . . . . . . . . . . .


3.3.5 Continuous Phase Density . . . . . . . . . . . . . . . . . . .
3.3.6 The Local Mass-Averaged Velocity . . . . . . . . . . . . . .
3.3.7 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . .
Discrete Phase Description . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1 Population Balance Equation . . . . . . . . . . . . . . . . . .
3.4.2 Moments of the Population Balance Equation . . . . . . .
3.4.3 Size Dependent Diffusion Coefficient . . . . . . . . . . . .
Heat and Mass Transfer From The Droplet . . . . . . . . . . . . . .
3.5.1 Evaporation from a Sphere . . . . . . . . . . . . . . . . . . .
3.5.2 Blowing Effects . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.3 Thermal Calculation . . . . . . . . . . . . . . . . . . . . . . .
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.6.1 Drying a Droplet of Colloidal Silica . . . . . . . . . . . . .
3.6.2 Drying a Droplet of Aqueous Sodium Sulphate Solution
Conclusions of the Chapter . . . . . . . . . . . . . . . . . . . . . . . .

4 Drying Droplets With an Outer Shell


4.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.1 Droplet Morphology . . . . . . . . . . . . . . . .
4.1.2 Formation of a Surface Shell . . . . . . . . . . .
4.1.3 Drying With a Rigid Shell . . . . . . . . . . . . .
4.2 Deciding Which Model to Use . . . . . . . . . . . . . . .
4.2.1 Physics of Drying Following Shell Formation
4.2.2 Criteria for the Different Models . . . . . . . .
4.2.3 Internal Boiling When Drying in Hot Air . .
4.3 Shell Thickening . . . . . . . . . . . . . . . . . . . . . . . .
4.3.1 Model Description . . . . . . . . . . . . . . . . .
4.3.2 Shell Growth . . . . . . . . . . . . . . . . . . . . .
4.3.3 Boundary Conditions . . . . . . . . . . . . . . . .
4.4 Dry Shell and Slow Boiling Models . . . . . . . . . . . .
4.4.1 Model Description . . . . . . . . . . . . . . . . .
4.4.2 Shell Growth . . . . . . . . . . . . . . . . . . . . .
4.4.3 Boundary Conditions . . . . . . . . . . . . . . . .
4.5 Wet Shell Model . . . . . . . . . . . . . . . . . . . . . . . .
4.5.1 Model Description . . . . . . . . . . . . . . . . .
4.5.2 Shell Growth . . . . . . . . . . . . . . . . . . . . .
4.5.3 Boundary Conditions . . . . . . . . . . . . . . . .
4.6 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6.1 Drying a Detergent Droplet . . . . . . . . . . .
4.6.2 Drying a Droplet of Colloidal Silica . . . . . .
4.7 Conclusions of the Chapter . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

5 Solution Methodology
5.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1 Equation Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.2 Solution Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

51
51
53
55
56
56
59
62
64
64
68
69
74
76
84
99

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

101
101
102
102
104
105
106
110
113
115
115
116
118
120
120
120
123
123
123
125
126
128
129
138
151

153
. 153
. 154
. 155

5.2

5.3

5.4

5.5
5.6

5.7

5.1.3 Moving Boundary Problems . . . . . . . . . . . . . . . . . . . . . . . . . 158


Model Equations and Boundary Conditions . . . . . . . . . . . . . . . . . . . . 162
5.2.1 Prior to Shell Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
5.2.2 Thickening Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
5.2.3 Wet Shell Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5.2.4 Dry Shell Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Solving the Model Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.3.1 The NAG Library Routine D03PLF . . . . . . . . . . . . . . . . . . . . . 170
5.3.2 Definition of the System of Partial Differential Equations pdedef 173
5.3.3 Definition of the Coupled ODEs odedef . . . . . . . . . . . . . . . 179
Boundary Conditions and Solution Grid . . . . . . . . . . . . . . . . . . . . . . . 181
5.4.1 Definition of the Boundary Conditions bndary . . . . . . . . . . . 181
5.4.2 Grid Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
5.4.3 Grid Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
The Numerical Flux Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
5.5.1 Definition of the Numerical Flux Function numflx . . . . . . . . 191
5.5.2 Roes Method for the Numerical Flux Function . . . . . . . . . . . . . 193
Numerical Parameter Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
5.6.1 Solution Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
5.6.2 Error Tolerances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
5.6.3 Maximum Time Step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
Conclusions of the Chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

6 Conclusion
6.1 Conclusions of the Thesis . . . . . . .
6.2 Suggestions For Future Work . . . . .
6.2.1 Extending the New Model .
6.2.2 Experimental Investigations .
6.2.3 Uses of the New Model . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Appendices

200
200
202
202
203
204
205

A Abel Transformation
206
A.1 Theory of the Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
A.2 Testing the Transformation Algorithm . . . . . . . . . . . . . . . . . . . . . . . . 207
B Roe Flux Function
B.1 Partial Differential Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B.2 Linearised Jacobian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B.3 Flux Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vi

210
210
211
214

List of Figures
2.1
2.2
2.3
2.4
2.5

Illustration of bound and unbound moisture within a drying material. . . .


Schematic drying rate curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mechanisms of moisture removal from a porous solid. . . . . . . . . . . . . .
The principle process stages of a generic spray drying process. . . . . . . . .
Product discharge from a co-current drying system with: (a) primary separation in the drying tower; and (b) total recovery in the dedicated separation equipment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.6 Size and segmentation of the laundry detergent market. . . . . . . . . . . . .
2.7 Schematic showing the structure of modern rate based spray dryer model
using CFD and particle sub-models. . . . . . . . . . . . . . . . . . . . . . . . . .
2.8 The different stages of drying for a liquid droplet containing solids. . . . .
2.9 Schematic showing some of the different dried-particle morphologies that
may result when drying droplets containing dissolved or suspended solids.
2.10 Images of hollow spray dried particles. . . . . . . . . . . . . . . . . . . . . . . .
Schematic showing a generic drying droplet containing suspended solids
and gas pockets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Every point, z , within the porous medium has an associated averaging surface, S , containing the volume, V . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Illustration of the terms used when volume averaging equations in a porous
medium consisting of a continuous and discrete phase. . . . . . . . . . . . . .
3.4 Schematic showing: (a) the model system; and (b) drying to form a hollow
shell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5 Illustration of a droplet with solids represented by a population of spherical
particles suspended in a continuous liquid phase. . . . . . . . . . . . . . . . .
3.6 Schematic illustrating the fluxes from an evaporating drop of A surrounded
by stagnant air. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.7 Plot of the saturated vapour pressure of water, comparing experimental
measurements with fitted equations. . . . . . . . . . . . . . . . . . . . . . . . . .
3.8 Comparison of simulated and measured (a) temperature and (b) mass profiles for the evaporation of a pure water droplet. . . . . . . . . . . . . . . . . .
3.9 Plot of the solids diffusion coefficient used to simulate the drying of a colloidal silica droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.10 Plot of the three sorption isotherms used in the investigation of colloidal
silica droplets drying. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.11 Plot showing the influence of the sorption isotherm on the predicted temperature and mass profiles when drying a colloidal silica droplet. . . . . . .

. 7
. 9
. 10
. 12
. 14
. 16
. 21
. 23
. 25
. 26

3.1

vii

. 34
. 38
. 40
. 44
. 57
. 65
. 73
. 75
. 77
. 79
. 80

3.12 Plot showing the influence of the relative air velocity, vrel , on the predicted
temperature and mass profiles when drying a colloidal silica droplet. . . . . .
3.13 Plot showing the influence of the air relative humidity, HR , on the predicted temperature and mass profiles when drying a colloidal silica droplet. .
3.14 Simulated solids volume fraction profiles during the drying of a colloidal
silica droplet in air at Tgas = 101 C. . . . . . . . . . . . . . . . . . . . . . . . . . .
3.15 Simulated drying of a colloidal silica droplet (lines) compared with experimental results from Neic and Vodnik (1991) (symbols) at Tgas = 178 C
with HR = 0% and v = 2.5 ms1 . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.16 Simulated solids volume fraction profiles during the drying of a colloidal
silica droplet in air at Tgas = 178 C. . . . . . . . . . . . . . . . . . . . . . . . . . .
3.17 Plot showing the binary diffusion coefficient, D(AB) , of Na2 SO4 in water at
25 C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.18 Water-Sodium Sulphate phase diagram adapted from Wetmore and LeRoy
(1951). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.19 The effect of initial particle number density and initial seed particle size on
the timing of shell formation, tshell . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.20 The effect of the nucleation rate, Nmax , on the timing of shell formation,
tshell when drying a droplet initially seeded with 10 nm crystals. . . . . . . . .
3.21 Plot showing the average solid particle size, L, at r = R at the point of shell
formation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.22 Plot showing how the maximum solid particle nucleation rate effects the
timing of shell formation for a droplet with initial m0(0) = 3.39 108 m3 s1 .
3.23 Plot showing spatially resolved profiles of the mean solid particle size, L, at
2 s intervals from t = 20 s until the point of shell formation, tshell . . . . . . .
3.24 Comparison of the simulated zeroth moment at the outer edge of a drying
droplet, (thick line) with that obtained from the first and second moments
using a closure approximation of the form given by (3.4.24), (symbols). . . .
3.25 Simulated drying of a 14 wt% sodium sulphate in water droplet at Tgas =
90 C, compared with experimental results from Neic and Vodnik (1991). .
3.26 Simulated solute mass fraction profiles plotted at 5 s intervals throughout
the drying of a droplet of aqueous sodium sulphate solution. In addition,
profiles are plotted at the point of shell formation, (dashed line) and solution at the edge of the droplet first becomes supersaturated, (bold line). . . .
3.27 Simulated normalised moments at the outer edge of the droplet. . . . . . . .
3.28 Simulated normalised moments integrated over the entire droplet. . . . . . .
3.29 Simulated solids volume fraction in a drying droplet of aqueous sodium
sulphate solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1
4.2
4.3
4.4
4.5

Schematic illustrating the process of shell thickening. . . . . . . . . . . . . . .


Illustration showing the drying process in the shell thickening period. . . .
Schematic demonstrating how the process of shell thickening can be likened
to the growth of a filter cake. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Illustration of the decision process used to select the appropriate drying
model to use following shell formation. . . . . . . . . . . . . . . . . . . . . . . .
Illustration of the pressure profiles within a drying droplet following formation of a surface shell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
viii

80
82
83

83
84
86
88
90
91
92
93
94
95
95

96
97
98
99

. 104
. 107
. 109
. 111
. 112

4.6
4.7
4.8
4.9
4.10
4.11
4.12
4.13
4.14
4.15
4.16
4.17
4.18
4.19
4.20
4.21
4.22
4.23
4.24
4.25
4.26

Illustration of a droplet of radius R drying through a dry shell with internal


radius, S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Evaporation from a droplet of radius R in the presence of a dry shell. . . .
Illustration of a droplet of radius, R drying through a wet shell with internal
radius, S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
(a) Initial composition of the crutcher mix droplets as measured by Griffith
et al. (2007); and (b) simplified description of the crutcher mix used to
model the system in the base case. . . . . . . . . . . . . . . . . . . . . . . . . . .
Water sorption isotherm for crutcher mix. The line shows the sorption
isotherm, (4.6.6), obtained by fitting to the experimental points obtained
by Bayly (2007). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simulated evolution of the moisture mass fraction in a crutcher mix droplet
(line) compared with experimentally measured values from Griffith (2008),
(symbols). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simulated moisture profiles in a drying detergent droplet (lines) compared
with experimental observations (symbols) from Griffith (2008). . . . . . . .
Simulated evolution of total droplet mass together with the mass of each of
the three components in the model droplet. . . . . . . . . . . . . . . . . . . .
Simulated time evolution of the dry shell interface during the drying of a
droplet of crutcher mix. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simulated solids volume fraction, ", profiles at 10 minute intervals during
the drying of a crutcher mix droplet. . . . . . . . . . . . . . . . . . . . . . . . .
SEM images showing detergent droplets dried in a spray dryer. . . . . . . .
Simulated lye phase sodium sulphate concentration profiles at 10 minute
intervals during the drying of a crutcher mix droplet. . . . . . . . . . . . . . .
Simulated drying of a colloidal silica droplet (lines) compared with experimental results from Neic and Vodnik (1991) (symbols) at: (a) Tgas =
101C and R0 = 0.972 mm; and (b) Tgas = 178C and R0 = 0.95 mm. . . .
Plot illustrating the influence of vapour pressure reduction as predicted by
the Kelvin equation on the simulated temperature profiles for a 1.9 mm
colloidal silica droplet drying in air at 178 C. . . . . . . . . . . . . . . . . . . .
Plot showing various pressures relating to a droplet of colloidal silica containing 16 nm particles drying in air at Tgas = 178C during the shell thickening regime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simulated morphological evolution of a droplet of colloidal silica containing 16 nm particles drying in air at 178C. . . . . . . . . . . . . . . . . . . . . .
Simulated solids volume fraction profiles during the drying of a colloidal
silica droplet containing 16 nm particles drying in air at Tgas = 178C. . . .
Plot showing various pressures relating to a droplet of colloidal silica containing 500 nm particles drying in air at Tgas = 178C during the shell
thickening regime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simulated morphological evolution of a droplet of colloidal silica containing 500 nm particles drying in air at 178C. . . . . . . . . . . . . . . . . . . . .
Simulated solids volume fraction profiles during the drying of a colloidal
silica droplet containing 500 nm particles drying in air at Tgas = 178C. . .
Schematic showing the effect silica particle size on the predicted drying
process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

ix

. 121
. 121
. 124
. 130
. 132
. 133
. 134
. 135
. 135
. 136
. 137
. 138
. 141
. 143
. 144
. 145
. 146
. 147
. 147
. 148
. 149

4.27 Plot showing various pressures relating to a droplet of colloidal silica containing 1500 nm particles drying in air at Tgas = 178C during the shell
thickening regime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.28 Plot showing how the Youngs modulus of the shell affects the size and
shell thickness of a dried-particle formed from a droplet of colloidal silica
containing 500 nm particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
5.12

5.13
5.14
5.15
5.16

Grid points, cells and cell boundaries on a one dimensional domain with
uniform grid spacing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Following shell formation, the drying droplet may be divided into two
adjoining solution domains, both of whose extents vary with time. . . . . . . 159
A spacetime illustration of a uniformly expanding solution domain showing the positions of the grid points, Ri (t ) and demonstrating the origins
of virtual flux terms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
Illustration of the co-ordinate transformation applied to a drying droplet
with a shell region. After transforming, the core and shell regions are both
fixed on the interval z [0, 1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
Illustration of the routines called by the time integrator, D03PLF. . . . . . . . 173
Numerical approximation of a Neumann boundary condition using the
two point, one sided difference formula. . . . . . . . . . . . . . . . . . . . . . . . 183
Construction of a solution grid placing grid points using (5.4.9) and taking
= 0.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Illustrating the effect of varying the parameter, in (5.4.9). Lower values
of result in more grid points clustered in the neighbourhood of the right
hand boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
An arbitrary distribution of grid points in the neighbourhood of the right
hand boundary of the z domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
Plot showing the predicted time to shell formation, tshell , when drying a
droplet of colloidal silica, illustrating the dependence on the number of
cells in the solution grid, npts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Plot illustrating the solids volume fraction within a droplet of colloidal
silica at the point of shell formation, showing the effect of increasing cell
density in a uniform grid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Plot illustrating the solids volume fraction within a droplet of colloidal
silica at the point of shell formation. The profiles obtained using a uniform
grid containing 20 and 500 cells are compared with that returned when
solving on a non-uniform grid containing 20 cells. . . . . . . . . . . . . . . . . 188
Demonstration of grid convergence as measured by the predicted timing of
shell formation, tshell , when simulating a drying droplet of sodium sulphate
solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Plot illustrating the influence of the finite difference formula used to approximate the boundary gradients on the convergence behaviour of the
solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
Plot demonstrating the efficiency of the solutions obtained using various
grid spacings and finite difference approximations to the boundary gradients.192
Plot illustrating the effect of varying the relative error tolerance, rtol. . . . . 196

5.17 Simulated profiles of the solids volume fraction is a droplet of crutcher mix
after 50 minutes of drying, illustrating the effect of reducing the maximum
integration step size, tmax . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
A.1 Numerical transformation of (A.2.1) with profiles showing the effect of
varying the number of data points, N . . . . . . . . . . . . . . . . . . . . . . . . . 208
A.2 The computation time, t , required to perform a double Abel transform
on (A.2.1) is proportional to N 2 , where N is the number of data points
used in the transform. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

xi

List of Tables
2.1
2.2
2.3
2.4

Summary of the major types of continuous driers used in the process industries, (Sinnott et al., 1999). . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Summary of the major advantages and disadvantages of spray dryers, (Masters, 1992). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sample process specification and results from applying a simple heat and
mass balance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sample process results and the sorption isotherm used to calculate the moisture equilibrium at the outlet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. 12
. 15
. 17
. 18

3.1

Values of the l 2 -norm of the deviations, "i , between predicted and measured values of the droplet mass for different drying air velocities, vrel . . . . . 81

4.1

The composition of the detergent slurry investigated by Griffith et al. (2008).129

5.1

Expressions for the various physical properties of water and air used in
droplet drying simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Coefficients of the one-sided n -point gradient approximations written for
a right hand boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
Comparison of accuracy and computational times for solutions obtained
on a uniform and non-uniform grid. . . . . . . . . . . . . . . . . . . . . . . . . . 188

5.2
5.3

A.1 Computational time to perform a double Abel Transform on (A.2.1). . . . . 207

xii

Chapter 1
Introduction
1.1 Motivation
Spray drying is the operation of choice for the production of many commercial products
ranging from high value pharmaceuticals to bulk commodities such as dried milk and detergent powders. The needs of these differing applications vary greatly. When producing
pharmaceuticals it is essential to maintain a sterile environment, whilst food products must
be dried in a way that ensures aromas and nutrients are retained. Detergent powders require
tightly controlled physical properties if customer demands concerning flowability and dissolution rate are to be met and, for any bulk drying operation, energy efficiency is likely to
be a principle concern. The spray drying operation may be tailored to suit all of these roles
and many more.
Spray drying works by contacting an atomised feed with drying air in a chamber. The
feed composition and process operating parameters are chosen to obtain the correct chemical and physical properties in the final product. In general, these properties will be influenced by the many processes occurring within the spray drying tower: in addition to
moisture removal and simultaneous particle formation, these may include particle build-up
on the dryer walls or inter-particle interactions such as agglomeration. A comprehensive
model of the spray drying operation would simulate all of these processes and such an accurate description is the holy grail of those working with spray dryers. The ideal model
would allow accurate control of product properties, prediction of dried-particle composition and morphology, specification of operating conditions for increased energy efficiency
and rigourous quantitative design of new towers allowing problems associated with wall deposition to be eliminated. However, despite their ubiquity and substantial research effort
over the past twenty years, spray driers continue to present substantial challenges to the
modeller and comprehensive dryer models remain elusive.
An alternative way of approaching the problem of spray dryer modelling is to focus on
simulating the drying behaviour of individual droplets. Simplified droplet drying models are
1

1.2. Novel Elements of This Thesis


essential components of top-down spray dryer simulations, wherein a computational fluid
dynamics description of the gas flow field is coupled with sub-models describing droplet
processes. More detailed single droplet drying models describe not only the rate of moisture
loss, but can also give valuable insights into the physical evolution of the drying droplets.
When droplets contain suspended or dissolved solids, such structural evolution can be complicated; an initial period of ideal shrinkage may be followed by formation of a surface shell,
internal bubble growth or inflation. Simulating the evolution of droplets undergoing such
processes is complicated, but necessary to predict the morphology and associated physical
properties of the final dried-particles. The development and validation of such a model is
the subject of this thesis.

1.2 Novel Elements of This Thesis


This thesis presents a completely new modelling framework for the simulation of single
droplet drying. This framework combines a physically motivated and rigorously derived
description of the droplet core, with various sub-models to describe processes following
shell formation. A number of these sub-models are developed, implemented and tested
within this thesis, but the general framework has the flexibility to allow future researchers
to extend the model with ease.
Beyond describing the rate of moisture loss from drying droplets, the key objective
of the new model is to describe the evolving droplet structure, allowing the simulation of
multiple dried-particle morphologies. Focussing on droplets containing dissolved solids in
an ideal binary solution with additional suspended solids, the general model framework:
tracks the spatially resolved continuous phase composition using transport equations
rigorously derived using the theory of volume averaging;
employs a population balance to describe the solid phase, allowing solid particles to
nucleate, move within the droplet and grow due to crystallisation from the continuous phase. The population balance is solved using a moment method, reducing
the description to a series of partial differential equations solved in parallel with that
describing the continuous phase.
This core droplet model is believed to be the most sophisticated generic description of
drying droplets produced to date.
Following shell formation, this thesis demonstrates how a number of sub-models may
be combined with the core droplet description to produce an integrated simulation of structural evolution. The sub-models investigated are:
Shell Thickening a completely new sub-model allowing for continued droplet
2

CSH

1.2. Novel Elements of This Thesis


shrinkage following shell formation, with evaporation proceeding from the droplet
surface;
Wet Shell Drying a sub-model from the literature including bubble growth and
describing drying in the presence of a rigid shell which is completely wetted by the
continuous phase;
Dry Shell Drying a common sub-model from the literature describing drying in
the presence of a rigid, dry shell where evaporation occurs at a front within the drying
droplet;
Slow Boiling a new sub-model describing moisture removal from a droplet at
temperatures above the boiling point of the continuous phase, without the droplet
shattering.
The thesis contains a detailed description of how these sub-models can be incorporated
within the general framework and validates this by comparing model predictions with experimental data from the literature. Further, for both the new sub-models and those taken
from the literature, extensive consideration is given to their physical motivation. This permits the development of rational criteria for deciding which sub-model to use at different
stages in a droplets drying history. It is the development of such criteria which permits,
for the first time, structural evolution to be simulated as a function of evolving droplet
composition and external drying conditions.
In summary, this thesis contains the following novel elements:
development of a framework to describe drying droplets based on volume averaged
transport equations and a population balance describing suspended solids;
development of the shell thickening sub-model;
extension of the wet and dry shell sub-models, providing physical motivation and
demonstrating compatibility with the new model framework;
development of physically motivated criteria to determine the appropriate sub-model
to use following shell formation and thus allow simulation of structural evolution
influenced by droplet composition and drying conditions;
simulation of systems of practical interest and validation of the new droplet drying
model against experimental data.

CSH

1.3. Structure of the Thesis

1.3 Structure of the Thesis


After this introductory chapter, Chapter 2 gives background information on the physics of
drying before discussing the use and modelling of spray dryers in more detail. The second
half of the chapter focusses on the literature relating to the simulation of single droplets, the
main subject of this work.
Chapter 3 describes the development of the core elements comprising the new droplet
drying model in detail. Comparisons with experimental data are presented for droplets
containing colloidal silica and dissolved sodium sulphate. Chapter 4 utilises the model
framework developed in the previous chapter and shows how this may be augmented with
various sub-models to describe droplet drying following formation of a surface shell. The
physics of this process is discussed and experimental comparisons are presented for three
types of droplet.
Chapter 5 discusses the numerical and computational methods used to solve the new
droplet drying model and thereby produce the results contained within this thesis. The
influence of the solution grid and numerical error tolerances are investigated at length.
Finally, Chapter 6 contains conclusions of the work presented and suggests ways in which
the novel elements developed herein might be taken forward.

CSH

Chapter 2
Background
In which the background material for the rest of this thesis is introduced. The basic
thermodynamics of drying are reviewed and an overview of industrial dryer selection
presented. There follows a more detailed discussion of the spray drying process and
a review of the literature relating to the simulation of spray drying towers. Single
droplet drying models are key to the success of such simulations and the development
of a new such model is the focus of this thesis. This chapter concludes with a critical
review of existing droplet drying models and indicates how the new model extends
the field.

2.1 Drying
Drying is the separation of solids and volatile substances most commonly moisture by
the application of heat to cause vaporisation1 , (Keey, 1975). The production of most solid
materials involves drying at some stage; the goals of such operations range from reducing
transport costs to providing specific product properties, such as porosity of a laundry detergent, (Coulson et al., 1996; Sinnott et al., 1999). As well as being a common industrial
operation, drying can also be an expensive one due to the inherent heat requirements, (Keey,
1978). A better understanding of the drying process therefore promises both cost savings
and enhanced product characteristics.
Before going any further, it is worth reviewing the ways in which moisture may be
present within solids and the basic thermodynamics of the drying process. Also, since most
dryers use hot air to remove moisture from solids, (Coulson et al., 1996), it is important to
know how to describe the moisture content of a gas.

Squeezing and adsorption are sometimes included in lists of drying processes although these do not
involve the use of thermal energy.

CSH

2.1. Drying

2.1.1 Moisture in Gases and Solids


Moisture in gases is described in terms of a number of different humidities. The absolute
humidity of a gas mixture is defined as the ratio of the mass of water vapour to the mass
of vapour-free, bone-dry gas, (Perry and Green, 1997). Assuming the gas to be ideal, this
may be written
H =

pAWA
(P pA)W g

(2.1.1)

where WA and W g are the molar masses of water and dry gas respectively. P is the total
pressure and pA is the partial pressure of water in the gas. The saturation humidity, H sat ,
is the humidity of saturated air, i.e., when pA = pAsat .
The percentage relative humidity is defined as the partial pressure of water vapour in
the air divided by the saturated vapour pressure of water at the same temperature, i.e.,
HR =

pA
pAsat

(2.1.2)

100% .

This should not be confused with the percentage absolute humidity, Habs , which is defined as the humidity divided by the saturation humidity at the same temperature.
Moisture containing solids may be divided into three categories: hygroscopic porous,
non-hygroscopic porous and non-porous hygroscopic bodies, (Keey, 1975). Examples
of hygroscopic materials are salts, vegetal fibres, most metal oxides and many polymers
whilst metal powders and glass granules are examples of non-hygroscopic products. In
some materials such as gels moisture behaves as if it was dissolved in the solid and
moves by diffusion under ever decreasing concentration gradients. Air cannot penetrate
these non-porous hygroscopic bodies.
In non-hygroscopic porous materials, moisture is merely trapped in the spaces between
solid particles. This unbound moisture exerts an equilibrium partial pressure approximately equal to the partial pressure above pure water at the same temperature. In addition
to unbound moisture, hygroscopic porous materials may also contain some bound moisture. This may be in small capillaries2 , adsorbed on to solid surfaces, chemically bound in
2

The Kelvin equation,


pA =

pAsat exp

2V m,L cos
rM R g T

(2.1.3)

may be used to estimate the reduction in vapour pressure for moisture held in capillaries with radius, rM .
pA is the vapour pressure above the capillary and pAsat is the partial pressure above pure water at the same
temperature, T . V m,L is the molar volume of the liquid and is the contact angle. This equation shows
that for water at 50 C in a 1 m capillary and the vapour pressure reduction is less than 0.1%. However,
there are circumstances where the correction becomes important, as discussed in Section 4.3.2.

CSH

2.1. Drying

100
Unbound Moisture

E
Mo quilib
istu rium
re C
urv
e

Relative Humidity, HR / %

Bound Moisture

Drying air humidity, HR,air


Equilibrium
Moisture

Free/Active Moisture
(moisture removed during evaporation)

Dry Mass Moisture Content, u

Figure 2.1: Illustration of bound and unbound moisture within a drying material as plotted on a
schematic sorption isotherm. Adapted from Masters (1992).

the form of a hydroxyl ions or crystalline hydrates or in solutions within the solid, (Masters,
1992). The strength with which the absorbed water is bound to the product depends upon
the method of absorption.
The atmospheric humidity in equilibrium with a given solid moisture content is given
by a sorption isotherm. Such an isotherm is shown schematically in Figure 2.1, which
also illustrates some of the other terms defined in this section. The moisture content, u , is
given on a dry mass basis, that is, the ratio of the moisture mass to the mass of moisturefree solid material. The complexity arising from the numerous mechanisms for moisture
uptake just discussed means that, in practice, sorption isotherms for a given material must
be determined experimentally.

2.1.2 The Driving Force for Drying


Water activity reflects the active part of the moisture content. This is the part which,
under normal circumstances, can be exchanged between the product and its environment.
The water activity is a measure of the vapour pressure generated by the moisture present in
a hygroscopic product and is defined
aw =

pA
pAsat

(2.1.4)

CSH

2.1. Drying
Here, pA is the partial pressure of water calculated at the wet bulb temperature and pAsat is the
saturation pressure the partial pressure of water vapour above pure water at the product
temperature. Bound moisture has a water activity less than unity and therefore exerts an
equilibrium vapour pressure lower than the partial pressure above pure water at the product
temperature; in contrast, the activity of unbound moisture is unity.
Static equilibrium is defined as a set of conditions under which a material does not
exchange moisture with its environment. The water activity is defined under equilibrium
conditions which, by definition, implies that the partial pressure of water vapour at the
surface, pA,sur , is equal to the partial pressure of water vapour in the surrounding bulk
gas, pA, . In non-equilibrium situations, moisture exchange between the product and its
surroundings, i.e., drying, is driven by a difference in these partial pressures,
dM
dt

= KA pA,sur pA, ,

(2.1.5)

where K is the mass transfer coefficient based on a partial pressure driving forces. M is the
total mass of the droplet and A is the area available for mass transfer. Free or active moisture
is that in excess of the equilibrium level. This may consist of unbound and some bound
moisture. Only free moisture may be removed in a drying operation.

2.1.3 Stages of Drying


Experimental observations of drying porous solids typically record the variation of moisture
content with time; a typical such curve is shown schematically in Figure 2.2a. It is clear that
the drying rate varies as a function of time, a result arising as a consequence of the various
processes influencing moisture removal. It is possible to identify a number of different
drying periods, as was first done in the classic works of Sherwood (1929a,b, 1930). These
stages of drying are more clearly observed in Figure 2.2b, which is a plot of drying rate
against moisture content.3
As drying commences, there is a brief period (AB in Figure 2.2) during which the solid
heats and the drying rate increases. When drying droplets, this stage is typically very short:
Oakley (2004) gives an example showing the warming-up period lasts 0.1 s for a 300 m
droplet. The drying rate increases until point B where thermal equilibrium is established
between the solid and its surroundings. The line BC represents the continuation of this
condition of dynamic equilibrium during which the energy lost by evaporating moisture
equals that transferred to the solid from the drying gas. This is the constant-rate drying
period where the moisture flux from the surface remains near constant. The surface of the
drying material and thus the vapour directly above remains saturated with moisture
3

Note that the plots in Figure 2.2 refer to drying in the context of moisture removal from porous solids.
The situation is a little different when moisture is present in solution, as discussed in Section 2.4.1.

CSH

2.1. Drying

Dry Mass Moisture Content, u

D
E

Time, t
(a)

Falling-Rate Period

Second
Falling-Rate Period

vap
Drying Rate, m

Constant-Rate Drying Period


B
A

First
Falling-Rate
Period

Dry Mass Moisture Content, u


(b)

Figure 2.2: Schematic drying rate curves showing: (a) moisture content on a dry basis against time; and
(b) drying rate against dry mass moisture content.

CSH

2.1. Drying
Funicular

Flow

Diffusion

Flow

(a)

Pendular

Diffusion
Evaporation

(b)

Figure 2.3: Following the critical point, the drying front recedes within the porous solid. (a) During the
first falling-rate period, the liquid is in the funicular state and so flow to the surface is possible. Although
there is some diffusion in the vapour phase, most moisture is evaporated at the surface. (b) In the second
falling-rate period the fluid near the surface breaks up and enters the pendular regime. Evaporation from
the solid is now controlled by vapour diffusion from the funicularpendular boundary.

throughout the constant-rate period. Further, when convection is the sole mechanism of
heat transfer to the droplet, the solid surface is at the wet bulb temperature, (Perry and
Green, 1997).
The constant-rate period persists whilst moisture is supplied to the surface at a rate
sufficient to maintain saturated conditions. When this is no longer the case the drying rate
begins to decrease and the temperature rises, (Scherer, 1990). This is the start of the first
falling-rate period, CD, during which air enters the pores and the drying front recedes
beneath the solid surface. The transition between these two regimes, point C in Figure 2.2,
occurs at the critical moisture content. The critical moisture content is a function of the
drying history and since it is an averaged value the mass and geometry of the material
being dried.
During the first falling-rate period, the liquid remains in the funicular condition and
so contiguous pathways to the surface persist, (Figure 2.3a). As a result, the majority of
the evaporation continues to occur from the external surface with some moisture evaporating within the unsaturated pore-space and moving to the surface by diffusion, (Schlnder,
2004). It is for this reason that the first falling-rate period is sometimes referred to as the
period of unsaturated surface drying, (Perry and Green, 1997).
Eventually, the drying front recedes so far that the pathways to the surface break up
and the liquid near the exterior enters the pendular condition, (Figure 2.3b). The liquid
flow to the surface then stops and moisture is removed solely by vapour diffusion. This is
said to be the second falling-rate period, (DE in Figure 2.2), during which the drying rate
drops further and the external surface of the solid begins to approach the temperature of the

10

CSH

2.2. Spray Drying


drying air. The rate of mass transfer is now controlled by the resistance of the solid layer.
The drying rate falls to zero once the equilibrium moisture content is reached, (E).

2.1.4 Industrial Drying


The differing demands made by various industries of their drying equipment has led to a
wide variety of dryer types, (Keey, 1978; Perry and Green, 1997). Dryer selection is primarily driven by consideration of the needs of the product. For example, granular detergent
manufacturers need to produce a powder with the right dissolution rate, placing great importance on the porosity, density and pore size of the dried product. They also need a
drying system that can handle a high throughput of material at a relatively low cost. In
contrast, the food industry is interested in flavour retention and may often be working with
smaller quantities of high-value, heat sensitive products. All these factors, along with others such as how the product is conveyed and the heating medium used, will play a part in
dryer selection, (Coulson et al., 1996). Table 2.1 lists some of the main dryer types used in
industry.

2.2 Spray Drying


This thesis is concerned with spray drying and, in particular, the development of a new
model to describe the drying of individual droplets within a spray drying tower. In this
section, the spray drying process and typical industrial applications are discussed in more
detail.

2.2.1 Process Overview


At its simplest level, spray drying involves the feed, in liquid or slurry form, being sprayed
into a drying medium. This is normally hot air. In more detail, the spray drying unit
operation may be considered as being composed of the four stages illustrated in Figure 2.4.
The atomisation of a pumpable feed to form a spray is the key characteristic of spray
drying. Two principle types of atomiser are used in industry: rotary atomisers making use of
centrifugal energy; and pressure nozzles which exploit pressure energy to atomise the spray.
Multiple injection levels may be used to handle higher flowrates. Whichever the type of
injector chosen, the initial droplet diameter will be in the range 20-500 m. The result of
the atomisation must be a spray which provides optimum evaporation conditions leading
to the desired characteristics in the dried product.
The manner in which the spray droplets contact the drying medium determines their
subsequent drying behaviour and, in turn, greatly influences the properties of the final
product. The form of spray-air contact is determined by the location of the atomiser relative

11

CSH

2.2. Spray Drying


Table 2.1: Summary of the major types of continuous driers used in the process industries, (Sinnott et al.,
1999).

Dryer Type

Description

Conveyor/
Tunnel

Solids move along a tunnel on trays or a continuous belt. Drying air is


passed over the solids or up through the belt. The temperature profile can
be controlled and the thermal efficiency is good, but such driers have long
retention times and high set-up and maintenance costs.

Drum

Used for liquid and dilute slurries which form a film on the surface of a
heated, rotating drum. Drum driers have high thermal efficiencies and,
due to short residence times, are suitable for heat sensitive materials.

Fluidised
Bed

Suitable for granular and crystalline solids with a diameter of 1-3 mm.
These driers have short retention times, rapid and uniform heat transfer
and good temperature control, but power requirements are high.

Pneumatic

Upward flowing hot gas pneumatically conveys and rapidly dries suspended particles. Such driers have poor thermal efficiency, but are useful
for particles which must be dried rapidly but are too fine for a fluidised
bed. Short contact times mean large particles cannot be dried.

Rotary

Free-flowing solids move along a rotating, inclined cylinder and are dried
by contact with drying air and the cylinder wall. High throughput, high
thermal efficiency and relative low capital and running costs are set against
difficulty achieving uniform drying.

Spray

Can handle any pumpable material. The feed is atomized to droplets


which then fall through drying air forming particles, which are then collected by cyclones or bag filters. Spray driers have high throughputs, short
residence times and good control over product properties. However, energy requirements are high.

Atomisation

Spray-Air Contact

Feed
rotary atomizer

pressure nozzle

co-current

countercurrent

mixed

Separation
Powder

total product
product discharge
from chamber and discharge from
separation unit
separation unit

Spray Evaporation

Figure 2.4: The principle process stages of a generic spray drying process, adapted from Masters (1992).

12

CSH

2.2. Spray Drying


to the air inlet. Broadly speaking, the flow may be considered either co- or countercurrent.
In the co-current arrangement the product and air pass through the dryer in the same
direction. This is by far the most common arrangement, (Zbicinski and Zietara, 2004),
and is especially suited to the drying of heat sensitive products. As the wet feed immediately
contacts the hottest air, drying is rapid and the drying air cools accordingly. The product
temperature remains around the wet bulb temperature throughout the initial drying period.
Subsequently, the product is in contact with cooler air and is at no point subject to thermal
degradation. Counter-current operation offers greater thermal efficiency as the liquid feed
and air enter at opposite ends of the drier. However, this means that the driest material is
exposed to the hottest air. Consequently, the set-up is only suitable for products which are
non-heat-sensitive. There are also dryer designs which combine co- and counter-current
flow patterns and these are termed mixed-flow driers.
The choice of how to contact the spray with the drying air is determined by the material
being dried and the desired product properties. Co- and counter-current arrangements give
different particle morphologies due to the different particle temperature histories. This can
lead to counter-current set ups producing a less porous product with higher bulk density.
Slower evaporation reduces the tendency to puff, lowering the particle porosity. However,
as mentioned above, the configuration may only be used for products which can withstand heat treatment. Conversely, co-current driers feature rapid evaporation preventing
high particle temperatures. The downside is that such high drying rates are more likely to
cause particle expansion or fracture, producing non-spherical, porous particles. The degree
of agglomeration will also be affected by the dryer arrangement and, in turn, will influence product properties such as coarseness. The morphological development of droplets is
considered further in Section 2.4.
Evaporation occurs as the fluid feed comes into contact with the drying air. As discussed
previously, the drying rate will be determined, in part, by the physical layout of the spray
drying operation. However, it is also important to understand how the composition of the
feed will affect the drying behaviour of the particles in the spray. The unique feature of spray
drying is that drying and formation of a powdered product occur simultaneously. Hence
an understanding of the physics of droplet drying is essential in order to describe the unit
operation.
Once the dried product has formed, a final separation stage is necessary. Two principle systems may be identified. In the first, (Figure 2.5a), primary separation occurs in the
drying tower itself, with the majority of the product being removed from the base of the
tower. The remaining product exits entrained in a separate air discharge stream, which is
sent to secondary separation equipment, e.g., cyclones, bag filters or electrostatic precipitators, (Perry and Green, 1997). The second system, (Figure 2.5b) operates with total

13

CSH

2.2. Spray Drying


Product In
Air In

Product In
Air In

Air In
Air Out

ATOMIZER

Air In
Air Out

ATOMIZER

Air and
entrained
powder

CYCLONE

DRYING CHAMBER

CYCLONE

DRYING CHAMBER

Total Product
Discharge

Secondary
Product
Discharge

Primary
Product
Discharge

(a)

(b)

Figure 2.5: Product discharge from a co-current drying system with: (a) primary separation in the drying
tower; and (b) total recovery in the dedicated separation equipment. Diagram adapted from Masters
(1992).

recovery of the dried product in the separation equipment. This places great importance on
the efficiency of the separation system employed and, for obvious reasons, can only be used
with a co-current set-up.

2.2.2 Applications of Spray Drying


The range of spray drying technology in current usage reflects the diversity of the industries
that dry their products in this way, (Masters, 1992). The ubiquity of this drying technology results from a number of factors. Of the dryer types listed in Table 2.1, the spray and
drum dryers are the only two that can handle a pumpable fluid feed. Of these, the spray
dryer is the only one which can produce powders of a specific particle size, porosity and
moisture content. Spray dryers are also capable of handling a very wide range of feed materials and flow-rates: a pharmaceutical company may use a lab-scale system to dry a few
kilograms of high-value product whereas the mining industry might use a much larger drier
to continuously process over 100 tonnes of material per hour. The principle advantages and
disadvantages of spray dryers are outlined in Table 2.2.
The production of laundry detergents is one of the best known applications of spray
drying and continues to be a large market, (de Groot et al., 1995). Figure 2.6 shows that
spray dried laundry products powder and tablets account for almost 80% of the
UK laundry detergent market, (Snapdata, 2005a). In the US, the percentage is less, but
this still represented a $1.2 billion market in 2004, (Snapdata, 2005b). Most detergents
are formed in counter-current towers with multi-level nozzle atomisation, (Masters, 1992).

14

CSH

2.3. Spray Dryer Modelling


Table 2.2: Summary of the major advantages and disadvantages of spray dryers, (Masters, 1992).

Advantages

Disadvantages

Specific product properties can be achieved High fabrication and installation costs.
consistently throughout the dryer run.
Operation is continuous and easily adapt- Large size requires expensive supporting
able to automatic control.
structures.
Unit can be designed for virtually any ca- Poor thermal efficiency (except with very
pacity required.
high inlet-air temperatures).
Can control product density and porosity.
Can handle heat-sensitive and heatresistant materials.
Can handle flammable, explosive, malodorous and toxic materials and those requiring hygienic conditions.
Can dry feedstocks in solution, slurry, paste
or melt form, including corrosive and abrasive feeds.
The product is relatively heat insensitive, allowing drying air temperatures of up to 400 C
and, consequently, relatively high thermal efficiencies are achieved. High throughput and
good reliability are especially important when producing such a bulk commodity. Spray
drying allows this, consistently producing powder with the correct physical properties to
yield the desired characteristics in the finished product.

2.3 Spray Dryer Modelling


Despite the importance and wide application of spray drying technology discussed in the
previous section, the theoretical modelling of spray dryers is still relatively poorly developed.
Spray dryers are more difficult to model than other dryer types4 ; the complex interactions
between the swirling gas flow and atomised drying droplets mean that simple scale-up techniques common elsewhere in chemical engineering design cannot be used, (Oakley, 1994).
Marshall and Seltzer (1950a,b) considered the basic principles and design aspects of
spray drying, reviewing the knowledge and industrial practice at the time. This was, on
the whole, based on correlations and qualitative observations of different drying systems.
One of the first attempts to model the behaviour of a spray drying tower was by Chaloud
4

Most spray dryer models deal with the first three stages presented in Figure 2.4, i.e., the separation system
is not considered. This is the approach adopted in this work.

15

CSH

2.3. Spray Dryer Modelling

Others
7%

Liquid
14%

Powder
27%

Powder
43%
Liquid
73%

Tablet
36%

Total UK Market Value (2004):


820 million

Total US Market Value (2004):


$4.5 billion

(a)

(b)

Figure 2.6: Size and segmentation of the laundry detergent market in: (a) the UK; and (b) the
US, (Snapdata, 2005a,b). Data refers to the 2004 financial year and segmentation is based on value %.

et al. (1957). They demonstrated the logical application of chemical engineering principles
to deduce a number of operating lines for spray drying towers. These related the moisture
content and bulk density of the dried powder to the drying rate and, beyond this, to independent variables such as slurry moisture content, tower height and drying-air throughput.
Such an approach formalised previous intuition-based understanding of tower behaviour,
but was still essentially qualitative.
To classify more quantitative descriptions of spray dryer behaviour, it is helpful to adopt
the hierarchy of modelling levels introduced by Oakley (2004):
Level 0
Level 1
Level 2A
Level 2B

Heat and Mass Balances;


Equilibrium Based Models;
Rate-based with simplifying assumptions about fluid flow;
Rate-based with simulation (CFD) of the continuous gas phase and particle motion.

The most appropriate choice of modelling level to use depends on the detail and accuracy
required from the solution, that is, on the purpose of the model and problem to be solved.
The remainder of this section discusses each of these modelling levels in turn and reviews
their application to problems reported in the literature.

2.3.1 Heat and Mass Balances


The simplest modelling layer is to apply appropriate energy and mass balance equations
to the system. Such models require no detailed knowledge of the dryer geometry, or the

16

CSH

2.3. Spray Dryer Modelling


processes occurring therein, (Coulson et al., 1996; Masters, 1992). However, the inlet
streams must be specified, along with the exit moisture content.
As an example, consider a co-current device drying a product containing 32.8% moisture on a wet basis at a rate of 1800 kghr1 . The required outlet moisture content is 7%,
again on a wet basis. There are energy losses of 50 kW and it is assumed that the outlet
product is at the temperature of the surrounding gas. The heat and mass balance shows
that the air stream specified along with the rest of the problem in Table 2.3, is capable of
performing this drying operation: the outlet gas temperature will be 74 C and HR = 30%.
Whilst the predictive power of heat and mass balances is limited, they are useful to
assess the thermodynamic feasibility of an operation at the earliest design stage. There
are also some examples of such simple models being used on their own in the literature.
For example, Baker and McKenzie (2005) surveyed a number of industrial spray drying
installations and used a model based on a simple heat balance to estimate the wasted energy.
Goffredi and Crosby (1983) used a set of heat and mass balances to develop simplified
models of co- and counter-current driers and investigate their sensitivity to variations in
product and drying gas flow-rates. More recently, Montazer-Rahmati and Ghafele-Bashi
(2007) modelled a counter-current dryer by applying heat and mass balances over multiple
stages within the tower. They applied this approach to simulate a detergent production
tower with some success.
Table 2.3: Sample process specification and results from applying a simple heat and mass balance.

Product
Moisture
Feed moisture content
Feed rate (wet solids)
Feed temperature
Specific heat capacity

Generic Slurry
Water
0.328 kgkg1 wet solids
1800 kghr1
50 C
2 kJkg1 K1

Inlet gas
Absolute humidity
Flow rate (wet)
Temperature

Air
0.05 kgkg1 bone-dry gas
4500 kghr1
300 C

Product moisture content 0.070 kgkg1 wet solids


Exit air temperature
74 C
Relative humidity at exit 30%

17

Product In
Air In

Air In

Air Out

Product Out

CSH

2.3. Spray Dryer Modelling

2.3.2 Equilibrium Based Models


An alternative to specifying the outlet product moisture content is to use a phase equilibrium relationship to relate the product moisture content to the humidity of the surrounding
gas. Assuming that dried-particles at the outlet are in equilibrium with their surroundings
allows use of a sorption isotherm to predict the exit product moisture content, (2.1.1).
Equilibrium based models clearly require that the sorption isotherm is determined, but in
return they yield more information about the feasibility of a given drying operation.
A sorption isotherm could be used with the example summarised in Table 2.3 to predict
the product moisture content at the dryer outlet. Using the generic sorption isotherm shown
in Table 2.4 together with the feed conditions specified previously, it is simple to calculate
that the product will leave with a wet-basis moisture content of 8%. That is, using the
sorption isotherm has shown that the 7% figure guessed previously is thermodynamically
unattainable. The air leaving the dryer will be at 82 C with HR = 17%.

Feed specification as in Table 2.3.


Product moisture content 8%
Exit air temperature
82 C
Relative humidity at exit 17%

Relative Humidity, HR / %

Table 2.4: Sample process results and the sorption isotherm used to calculate the moisture equilibrium at
the outlet.

0.8

0.6

0.4

0.2

0
0

0.2

0.4

0.6

0.8

Wet Mass Moisture Content, u

An approach-to-equilibrium factor may be employed to handle situations where the


product outlet stream is not in equilibrium with its surroundings. However, as a user specified parameter this provides little insight, (Oakley, 2004). Hence, the equilibrium models
are really only of any use where equilibrium is attained. Intuitively, this is more likely to
be valid for small droplets with long residence times in the dryer. Ozmen and Langrish
(2003a) provide experimental evidence that this is indeed the case for dried-particles with a
final size of 30 m.

18

CSH

2.3. Spray Dryer Modelling

2.3.3 Rate Based Models using Simplifying Flow Assumptions


When it is not valid to assume equilibrium conditions at the outlet, the rate at which
moisture is removed from the sprayed droplets must be considered. Alongside this, it is
necessary to model the droplet residence time in the drying unit. Together, these allow the
product moisture content at the exit to be predicted. However, achieving both of these
objectives presents problems to the modeller.
Droplet residence time in a spray drying tower may be modelled in two main ways:
the first is by invoking simplifying assumptions about droplet motion whereas the second,
discussed in the next section, is to use computational fluid dynamics to simulate the whole
flow field.
Invoking simplifying assumptions about the droplet motion equates to finding shortcut
methods to model droplet residence times. Gauvin and Katta (1976) presented a design
methodology whereby laboratory-scale experiments were used to develop correlations for
droplet residence times in full-size spray towers. Their analysis allowed both drying rate
and residence time to be functions of droplet diameter.
Clement et al. (1991) modelled the gas in a spray dryer as a single well-mixed mass,
allowing the standard expression for the residence time distribution (RTD) of a perfectly
mixed reactor to be used. This was coupled to a shrinking core model to describe the
drying of single droplets which were all assumed to be the same size. Birchal and Passos
(2005) used the same physical model to simulate the drying of milk emulsions, although
they present an improved solution algorithm. Both papers present results which exhibit
fair agreement with experimental milk drying data. Later, Birchal et al. (2006) compared
this model with a more sophisticated computational fluid dynamics (CFD) simulation,
obtaining similar results from both for some key parameters. Palencia et al. (2002) used a
similar RTD-based approach, modelling the gas as a series of well mixed vessels in series.
Their method reproduced experimental data on milk drying with moderate accuracy.
The strengths of using simplified descriptions of the gas phase include the ability to
investigate dynamic changes in inlet conditions and flowrates and relatively fast computational times. The main weakness of this approach lies in the assumed residence time
distribution. Perfect mixing is nothing more than a first approximation which, for many
towers, will prove insufficient. Whilst it is simple to substitute an alternative RTD, this will
almost certainly be unique and therefore need to be measured for each unit modelled. Thus,
whilst this approach is useful for investigating changes to operating conditions on existing
spray drying installations, it is not an adequate tool for detailed design, (Reay, 1988).

19

CSH

2.3. Spray Dryer Modelling

2.3.4 Rate Based Models Using CFD


The most sophisticated level of spray dryer modelling is to extend rate based models by
introducing detailed descriptions of the fluid flow and particle processes. The advances in
computing power and the rise of computational fluid mechanics has made such detailed
descriptions possible, (Oakley, 1994). Today, such approaches represent the most comprehensive simulations of spray drying behaviour, (Fletcher et al., 2003).
Crowe (1980) was the first to use computational techniques to simulate a spray drying
tower. He used his particle-source-in-cell method, (Crowe et al., 1977) to handle the energy and momentum coupling between the gas and drying droplets. This is a LagrangianEulerian modelling approach, whereby the gas phase is treated as a continuum an Eulerian viewpoint with the droplets being tracked through the flow field in a Lagrangian
manner. Such an approach is sensible where one phase occupies a small fraction of the total
volume of the solution domain, (Huang et al., 2003), as is the case throughout most of a
spray drying tower.
It is possible to treat both phases as a continuum the fully Eulerian or two-fluid
model as was done by Platzer and Sommerfeld (2003) when modelling the dense spray
region around a spray nozzle. The method delivers velocities and volume fractions of both
phases as output parameters, but no information about the droplet size distribution. For this
reason the approach is rarely used where the spray is dilute and the distribution of droplet
sizes is important. A purely Lagrangian approach is also possible, although rare. Salman and
Soteriou (2004) presented such a model, which they claim has advantages when modelling
high volume-fraction, evaporating spray systems.
The vast majority of spray drying simulations use the combined Lagrangian-Eulerian
approach. Such a framework facilitates the incorporation of various physically motivated
sub-models to describe droplet processes. Fletcher et al. (2003) identified three such droplet
processes that merited inclusion in spray dryer simulations: droplet drying; dropletdroplet
interaction; and dropletwall interaction. A generic Lagrangian-Eulerian spray dryer model
incorporating all these elements is illustrated in Figure 2.7. Whilst CFD-based spray drying
models are relatively common in the literature, simulations incorporating all of these submodels are rare.
There exist a large number of attempts to use CFD to model gas flow patterns in spray
dryers, (e.g., Southwell and Langrish, 2000; Fletcher et al., 2003; Oakley, 2004). Langrish
and Fletcher (2003) look forward and conclude that commercially available CFD packages
are already capable of producing adequate simulations of the gas phase, although there is
still considerable discussion regarding the most appropriate turbulence model, (Bayly et al.,
2004; Oakley and Bahu, 1991; Zbicinski and Zietara, 2004). More recent papers focus on
coupling one or more of the droplet sub-models to the CFD simulation.
The description of droplet drying is the most important of the three droplet sub 20

CSH

2.3. Spray Dryer Modelling

Full Spray Dryer Model

Gas Flow

Sub Models

Particle Drying

Particle-Wall
Interaction

Particle-Particle
Interaction
(Agglomeration,
Breakage)

Figure 2.7: Schematic showing the structure of modern rate based spray dryer model using CFD and
particle sub-models.

models. In its simplest form, a drying model gives the rate at which moisture is lost
and, from this, the average droplet moisture content. More sophisticated models give the
spatially distributed moisture content, the temperature profile and perhaps even describe
the morphological development of the droplet. Many researchers have included such a
sub-model in their CFD simulations of spray dryers. These range from evaporating pure
liquid droplets (e.g., Papadakis and King, 1988; Huang et al., 2003; Li and Zbicinski,
2005; Huang and Mujumdar, 2005), through models using a material dependent drying
curve (e.g., Langrish and Kockel, 2001; Harvie et al., 2002; Huang et al., 2004; Zbicinski
et al., 2005; Zbicinski and Li, 2006), to fully spatially resolved approaches, (Verdurmen
et al., 2004). Single droplet drying models are discussed in detail in Section 2.4.
Langrish and Fletcher (2003) state that considerable work is needed to model the adhesion and cohesion of particles due to stickiness. This is a pre-requisite to producing
physically realistic sub-models for both dropletdroplet and, especially, dropletwall interactions. Although stickiness is a common concept, its nature is highly complex and it
is still poorly understood, (Kudra, 2003). Consequently, sub-models describing droplet
interactions are poorly developed at present.
When dealing with dropletwall interactions the typical approach is to assume all droplets
stick on collision with a wall, (Langrish and Zbicinski, 1994). While such a methodology
can produce useful results, (e.g., Straatsma et al., 1999), it is simplistic and fails to capture
any of the complexities found in experimental observations of the modelled systems, (e.g.,
Kota and Langrish, 2006; Ozmen and Langrish, 2003b). Similarly, agglomeration is normally handled using traditional kernels which only consider droplet size and possibly velocity, (Sommerfeld, 2001; Verdurmen et al., 2004). However, as with dropletwall interactions, other droplet properties in particular the surface moisture content have a strong
influence on the adhesion probability. Metzger et al. (2007) identify the development of

21

CSH

2.4. Single Droplet Drying Models


new kernels to reflect these dependencies as a priority for the future; such descriptions will
clearly be dependent on the predictions of the drying sub-model.
The most comprehensive simulation of spray dryers to date is the Efficient Design and
Control of Agglomeration in Spray Drying Machines (EDECAD) project, (Verdurmen
et al., 2004). The principle deliverable of the project was a Design-Tool to relate the dryer
geometry, process conditions, product composition and final powder properties. A spatially
resolved drying sub-model was used, although evolving droplet morphology was not considered. Dropletdroplet collisions are treated using the model of Sommerfeld (2001), with
agglomeration treated using the classical model of Brazier-Smith et al. (1972). The project
did not explicitly address droplet interaction with the wall.
In conclusion, it is clear that rate-based models using CFD coupled with droplet submodels provide the most detailed description of the spray drying process. Whilst considerable progress has been made towards the development of a unified dryer model, there is
room for improvement in all three droplet sub-models. The most fundamental of these is
that describing droplet drying; a sub-model giving the surface moisture content of drying
droplets is a pre-requisite for accurate prediction of particle agglomeration and wall build
up.

2.4 Single Droplet Drying Models


As discussed in the previous section, a droplet drying submodel is integral to any detailed
simulation of the spray drying process, (Langrish and Fletcher, 2003; Oakley, 2004). The
droplet moisture content specifically the surface moisture determines the rate of
other processes occurring within a spray tower, such as agglomeration and wall deposition, (Fletcher et al., 2003; Blei and Sommerfeld, 2006). Building on the theory presented
in Section 2.1.2, this section starts by reviewing the way in which single droplets dry, focussing on those containing suspended solids. The possible morphologies of spray dried
particles are then discussed, before the literature relating to single droplet drying models is
reviewed.

2.4.1 Droplet Drying Behaviour


The theory introduced in Section 2.1 provides a basis for understanding the drying behaviour of sprayed droplets. However, a number of modifications must be made before
such droplets can be modelled successfully.
The evaporation of pure liquid droplets has been extensively studied for many years, (e.g.,
Frssling, 1938; Ranz and Marshall, 1952). Droplets in real spray dryers generally contain
dissolved or suspended solids, but understanding the simpler pure liquid system forms the

22

CSH

2.4. Single Droplet Drying Models

Temperature, Td

Drying Gas Temperature, Tgas

Boiling Temperature,
Tboil
B

0 A
0

E
D

Wet-bulb Temperature

Time, t

Figure 2.8: The different stages of drying for a liquid droplet containing solids.

basis for describing more complicated drying mechanisms, (Masters, 1992; Oberman et al.,
2004).
It is well known that evaporation from droplets containing dissolved solids is slower
than that given by (2.1.5) the rate at which mass is lost from pure liquid droplets.
Solutes lower the vapour pressure of water, pA,sur , thus lowering the driving force for evaporation, (Masters, 1992). Annamalai et al. (1993) present a neat analysis of evaporation from
two-component droplets, considering the case where both components evaporate. Keey
(1992) discusses selective evaporation effects which may occur in such circumstances.
Whilst dissolved solids reduce the driving force for evaporation, Ranz and Marshall
(1952) showed that the presence of suspended, insoluble solids has a negligible vapour pressure lowering effect. Drying during the constant rate period may be treated in the same
way as for a pure liquid droplet. However, following formation of a rigid crust, the droplet
behaves more like a porous solid and the discussion in Section 2.1.3 applies.
Figure 2.8 shows a typical temperature history for an individual droplet containing
dissolved or suspended solids drying in a spray tower; many of the stages are similar to
those outlined for generic solid drying in Section 2.1.3. The droplet rapidly heats to the
wet-bulb temperature, (AB) and then remains at the wet-bulb temperature and drying
proceeds at a near constant rate whilst the surface remains saturated with moisture, (B
C). However, the presence of dissolved or suspended solids can alter this behaviour. A high
initial solids loading results in the constant rate drying period being brief, if observed at
all, (Cheong et al., 1986; Dolinsky, 2001). Any dissolved solids will concentrate as moisture
is removed, reducing the moisture vapour pressure and causing the surface temperature to

23

CSH

2.4. Single Droplet Drying Models


rise above the thermodynamic wet bulb.
The falling rate drying period begins when moisture can no longer be supplied to the
surface at a rate sufficient to maintain saturated conditions, (CD). The transition between
these two regimes occurs at the critical moisture content and, in the presence of suspended
solids, may also coincide with the start of crust formation, (Cheong et al., 1986).
During the falling rate period the droplet will heat above the wet bulb temperature.
Vaporisation will commence if the temperature reaches the boiling point of the solution.
Considerable energy is required for vaporisation and so the sensible heating of the droplet
halts (DE). This phase is termed the boiling regime and the drying rate is now controlled
by external heat transfer to the droplet. The presence of a dissolved solute will typically
raise the boiling point of a solution; the droplet temperature therefore increases slowly in
the boiling regime as the solution becomes more concentrated. Once all the free moisture
has been removed, the temperature will again rise, asymptotically approaching that of the
surrounding gas, (EF).

2.4.2 Droplet Morphologies


The removal of moisture from a spray of droplets involves simultaneous heat and mass
transfer and, uniquely in a spray dryer, this process is coupled to concurrent particle formation. When drying droplets containing dissolved or suspended solids, a wealth of different dried-particle morphologies may form. Which type of particle forms depends upon
the composition, size, temperature and drying history of the droplet, (Ranz and Marshall,
1952). Further, because droplet drying and particle formation occur simultaneously, the
drying mechanism and resultant kinetics are, in turn, strongly dependent on the evolving
droplet microstructure, (Huntington, 2004).
Figure 2.9 illustrates the main dried-particle morphologies that may result when drying droplets containing dissolved or suspended solids. At very low solids concentrations,
droplets continue to evaporate like pure liquid spheres and no particle is formed. However,
in any spray drying application of interest one of the other drying routes will be followed.
As described in the previous section, droplets initially shrink ideally in the constant rate period. Crust formation commences when the moisture content falls below a critical value at
a single preferential surface site, usually the point of maximum mass transfer as determined
by the surrounding flow field, (El-Sayed et al., 1990). Once initiated, the crust spreads
rapidly over the surface of the droplet, forming a structured solid shell and stabilising the
droplet diameter, (Cheong et al., 1986). The behaviour after this point strongly depends
on the nature of the shell formed and the drying conditions.
Walton and Mumford (1999a) identify three distinct categories of droplet, which behave
in morphologically similar ways when dried: crystalline, skin-forming and agglomerate.
Within these categories, the major differences in morphology result from the drying air
24

CSH

2.4. Single Droplet Drying Models

No particle
formation

Shattered
Particle

Solid Particle
Collapse

Low solids
concentration,
<1%w/w

High
temperature

Re-inflation
Dry Shell

Wet Shell

Initial Droplet

Saturated Surface
Drying

Crust Formation

Internal Bubble
Nucleation

Uninflated Shell

High temperature

Blistered Particle

Shrivelled Particle

Inflated Puffed
Particle

Figure 2.9: Schematic showing some of the different dried-particle morphologies that may result when
drying droplets containing dissolved or suspended solids.

temperatures and, consequently, the drying rate. At lower temperatures, the mechanisms
allowing for droplet shrinkage and deformation are more pronounced, (Alamilla-Beltrn
et al., 2005); moisture loss and the rate of shrinkage are slower, allowing more time for
structures to deform, shrink and collapse, (Oakley, 1997).
A solid dried-particle often forms when the drying gas temperature is below the moisture boiling point, (El-Sayed et al., 1990). Once a rigid crust has formed, such droplets
dry somewhat like a porous solid medium with moisture menisci receding into the droplet,
(2.1.3). However, low temperature drying does not always result in the formation of solid
dried-particles, (Walton and Mumford, 1999a,b). With aerated feeds, bubbles or voids
can arise as a result of two mechanisms. The droplet can become super-saturated with
any dissolved air as a result of increasing solute concentration. Bubbles typically nucleate around the transition from the constant- to falling-rate drying period, (El-Sayed et al.,
1990). Greenwald and King (1981, 1982) present results showing internal voidage formed
in this way. Alternatively, entrained air pockets can coalesce and expand during drying
to produce hollow particles. Verhey (1972) conducted an extensive study on drying milk
which demonstrated that the gas vacuoles in this system originated from air entrainment
during atomization.
At high temperatures, droplets tend to inflate, form crusts and blister or break, (AlamillaBeltrn et al., 2005). For aqueous solutions, this happens as the droplet temperature ap-

25

CSH

2.4. Single Droplet Drying Models

(a)

(b)

Figure 2.10: Images of hollow spray dried particles taken with (a): a scanning electron microscope; and
(b) an optical microscope. Courtesy of Cheyne et al. (2002)

proaches 100 C, corresponding to the boiling regime discussed in the previous section,
(Greenwald and King, 1981, 1982). Inflation results from large partial pressures of water vapour joining inerts in a bubble, (El-Sayed et al., 1990; Oakley, 1997). Subsequent
drying behaviour and final dried-particle morphology are determined by the chemical and
physical properties of the shell or film regions, (Walton and Mumford, 1999b). For example, the rheological properties of skin-forming materials allow such droplets to undergo
multiple inflation-collapse cycles. Such behaviour is observed when drying coffee extract,
(Charlesworth and Marshall, 1960; Hecht and King, 2000a) and skim-milk, (Walton and
Mumford, 1999a), amongst many others. This behaviour may be contrasted with less pliable crystalline droplets that tend to undergo only partial inflation or form hollow or semihollow dried-particles. As an example, Figure 2.10a shows an image of a dried detergent
droplet from Cheyne et al. (2002). The droplet has undergone partial inflation and contains a large central void. This is seen more clearly in Figure 2.10b where a dried-particle
has been cut open and imaged following capture in wax.
The morphology of the dried-particles produced depends strongly on the nature of the
shell formed. The droplet drying models discussed in the next section can say very little
about this, thus limiting the scope of the morphological predictions they are able to make;
in general, a given drying model is only capable of simulating the morphological evolution
towards one type of dried-particle. It would be a great achievement for a single drying model
to be capable of simulating multiple dried-particle morphologies, with structural evolution
determined by the evolving droplet composition and drying conditions.

26

CSH

2.4. Single Droplet Drying Models

2.4.3 Droplet-Averaged Drying Models


There are many single droplet drying models in the literature. At the highest level, these
can be divided into those which only model droplet averaged quantities such as moisture
content and temperature and those which are based on mechanistic pictures of droplet
drying. This second class of model returns some morphological information such as
dried-particle size and may give spatially resolved moisture profiles.
The simplest droplet-averaged approach to use is a characteristic drying curve, (Langrish
and Kockel, 2001; Chen and Lin, 2004; Huang et al., 2004). Essentially an empirical
method, the approach relies upon first identifying an unhindered drying rate and then
measuring a drying curve for each material considered. Once normalised, it is assumed that
this curve is unique for a given material and independent of external drying conditions,
sample size and geometry. If accepted, this implies that drying curves obtained from labbased single droplet studies may be used to describe drying of much smaller droplets in
industrial scale equipment.
Keey (1992) provides an overview of the experimental tests of the characteristic drying
curve. The results show that the method works well for several materials over a modest
range of temperatures, air humidities and velocities. However, the concept did not work
for large droplets with R0 > 10 mm. Also, there is no data reported for drying of slurry
droplets.
Huang et al. (2004) investigated the use of characteristic drying curves in CFD models of spray dryers. They demonstrated that the choice of drying model influences both
the droplet mass history and the path it takes through the spray dryer, but presented no
comparison with experimental data. The minimal computational expense of this method
makes it attractive for use in CFD applications, especially where transient flow patterns are
being investigated, (Langrish and Kockel, 2001). However, only droplet averaged properties
are returned; no information is obtained about morphological changes that may occur as a
result of drying, or even final dried-particle sizes.
The reaction engineering approach, introduced by Chen and Xie (1997) maintains the
simplicity of implementation associated with the characteristic drying curve whilst claiming
a sounder physical basis. The method considers drying as a competitive process between an
activation type evaporation reaction and a condensation reaction. A normalised curve of
the evaporation activation energy against moisture content is considered to be characteristic
of a given material.
Chen et al. have successfully applied the reaction engineering approach to a number
of spray drying systems, (Chen et al., 2001; Lin and Chen, 2005, 2006, 2007). Chen
and Lin (2004) showed the reaction engineering approach gives better predictions than
a characteristic drying curve when applied to drying milk droplets. Recently, Woo et al.
(2008) conducted a comparison of the two models and again appear to demonstrate the
27

CSH

2.4. Single Droplet Drying Models


superiority of the reaction engineering approach, especially when using an extension which
allows the surface moisture to be predicted. However, wider take up of the method is so far
lacking, perhaps because considerable experimental effort is required to obtain the activation
energy curve and sorption isotherm for each system.
Both the characteristic drying curve and reaction engineering approaches return only
droplet averaged properties; they give no spatially resolved or morphological information.
Nevertheless, such models are still common in CFD simulations as they are simple to implement and computationally cheap.

2.4.4 Mechanistic Models


Spatially resolved, or mechanistic models are based upon simplified pictures of moisture
movement within a drying droplet. A number of such simplified pictures exist, although
almost all simulations of solid containing systems are based on just two. These may be
loosely termed the effective diffusion and shrinking core approaches. Models containing
a bubble may be considered a third class of model which builds on aspects of the first two
approaches. This section reviews the many different drying models found in the literature,
grouping them into these three broad categories. All the models discussed here assume that
the droplets remain spherical throughout their drying history.
Effective Diffusion Coefficient Models
Effective diffusion coefficient models assume that moisture transport within a drying droplet
can be described by Fickian diffusion. It is possible to model binary systems exactly using
the convection diffusion equations, (e.g., Hecht and King, 2000a,b), but in most cases the
diffusion coefficient matrix is unknown or if there are more than two components the
system is simply too complicated. In such cases it is found that an effective diffusion coefficient is required to adequately reproduce experimental results, (Whitaker, 1977). This
effective diffusion coefficient is normally a strong function of local moisture concentration
and temperature and, as such, needs to be determined experimentally for each system investigated, (Charlesworth and Marshall, 1960; Ferrari et al., 1989; Kentish et al., 2005).
There is no explicit formation of a shell region in such models. However, a reduction in
the effective diffusion coefficient at low moisture contents can achieve a reduction in mass
transfer similar to that caused by shell resistance. Whilst it is generally necessary to determine effective diffusion coefficients experimentally, Whitaker (1977) demonstrated that,
under certain assumptions, the same equations can be derived from a rigourous treatment
of mass transfer.
Van der Lijn (1976) was the first to apply an effective diffusion coefficient approach to
individual droplets drying in spray driers, although the general use of such methods goes

28

CSH

2.4. Single Droplet Drying Models


back much further, (Lewis, 1921; Sherwood, 1929a,b, 1930). Since then, effective diffusion
models have been used by a large number of researchers to simulate single droplets drying.
Wijlhuizen et al. (1979) based their model on an effective diffusion coefficient approach
and applied it to investigate the inactivation of phosphate during spray drying of skim
milk. Ferrari et al. (1989) used experimental data to determine an appropriate effective
diffusion coefficient for drying milk powder and subsequently obtained close agreement
with measured drying curves. Adhikari et al. (2003, 2004) used a model based on that
described by Van der Lijn (1976) to investigate parameters affecting surface stickiness during
drying. All these studies would be impossible without a spatially resolved model.
The model of Sano and Keey (1982) has been used by many researchers to simulate
droplets in spray driers. Etzel et al. (1996) and Kentish et al. (2005) both used variants of
this to simulate milk drying and a simplified version of the model is used as the droplet
drying sub-model for the EDECAD project, (Verdurmen et al., 2004). Similarly, the model
of Frey and King (1986) has been used many times, for example, by both El-Sayed et al.
(1990) and Wallack et al. (1990) to simulate coffee drying. Recently, Porras et al. (2007)
have presented a treatment of a binary liquid in a porous medium using an effective diffusion
coefficient approach and volume averaging.
The effective diffusion coefficient approach is the most common method used when
spatial moisture profiles are desired. However, on its own, the method yields relatively little
information about droplet morphological development. For this reason, several models
base their description of moisture transport on such an effective diffusion approach which
they then combine with a further model for morphological changes. Some such models are
discussed below.
Shrinking Core Models
The shrinking core model consists of two stages. Initially the droplet shrinks ideally, with
solids accumulating at the droplet surface. After some time, a crust forms and the subsequent size of the droplet is fixed. The point at which crust formation occurs may be given
by a critical moisture fraction averaged across the particle (e.g., Mezhericher et al., 2007), or
at the surface (e.g., Liang et al., 2001), or by some critical droplet-averaged porosity, (e.g.,
Kadja and Bergeles, 2003).
After the formation of a surface shell, the model enters the second stage in which the
solution-crust interface recedes into the porous particle. Evaporation occurs at this receding
interface and water must be transported to the surface by vapour diffusion through the dried
shell. The mass transfer resistance of the crust therefore becomes the factor limiting the rate
of continued drying.
The simplest application of the shrinking core approach assumes that the droplet outside
the crust region is well mixed and, consequently, the need to track the spatial distribution

29

CSH

2.4. Single Droplet Drying Models


of water is removed. Such models have been used to simulate droplets of sodium sulfate
decahydrate, (Cheong et al., 1986) and coal slurries, (Kadja and Bergeles, 2003) amongst
many others, (Audu and Jeffreys, 1975; Dolinsky, 2001; Neic and Vodnik, 1991).
If spatial information on the droplet moisture content is required, a shrinking core
model can be combined with an effective diffusion approach. Prior to shell formation moisture movement within the droplet is described using an effective diffusion coefficient. The
trigger for shell formation can then be based on a value at the droplet surface. A dry shell
subsequently grows inwards as per the shrinking core model, with an effective diffusion coefficient continuing to describe the wet core. Such an approach is adopted by Dalmaz et al.
(2007) to simulate droplets of colloidal silica and skim milk and by Werner et al. to model
an amorphous polymer solution, (2008a) and maltodextrin, (2008b; 2008c). Neic and
Vodnik (1991) presented the largest selection of experimental comparisons, investigating
colloidal silica, sodium sulphate and skim milk systems. Whilst agreement with experimental measurements is good in all these papers, the comparison is limited to droplet mass and
temperature histories. Consequently, the spatial predictions of such models have not been
experimentally verified.
Elperin and Krasovitov (1995) extend the shrinking core model further, by solving transport equations within the shell region itself. Seydel et al. (2006) present the most advanced
model to date: they employ a population balance to describe the suspended solids and
couple this to an effective diffusion equation for the droplet moisture. A shrinking core
type approach then describes morphological evolution after the formation of a rigid shell.
Whilst impressive, the implications of this model are not fully explored and comparison
with experimental data is limited to qualitative observations.
One of the key postulates of the shrinking core model that evaporation proceeds at
the interface between the dried shell and the wet core implies that this interface remains
at or around the wet bulb temperature. This results in the prediction of a large temperature
gradient across the crust region. Recognising this, Farid (2003) developed a model based
on the shrinking core approach in which heat transfer due to both internal conduction and
external convection controlled the drying rate. However, the temperature gradients such an
approach implies are unphysical and contradict experimental findings, (Schlnder, 2004;
Dalmaz et al., 2007). These typically show that their are no major temperature variations
within drying particles, (Chen and Peng, 2005). To overcame this objection, Schlnder
(2004) introduces a wet surface model which invokes capillary transport through the crust
and only allows vaporisation at the particle-air interface. However, to date no work has been
done towards implementing such an approach within a droplet drying simulation.
Lee and Law (1991) point to a second problem with the shrinking core approach: such
models conserve mass only if the porosity of the dried shell is the same as that of the original
slurry. This condition is not generally true. Lee and Law overcome this by introducing a

30

CSH

2.4. Single Droplet Drying Models


continuously expanding vapour-saturated space located at the centre of the particle.
Models With a Bubble
The third class of single droplet drying model discussed here encompasses those models
which include a centrally located bubble. Such models allow the simulation of hollow
particles; as discussed in Section 2.4.2, such a dried-particle morphology is common in
practice.
Wijlhuizen et al. (1979) were among the first to publish a drying model with a bubble.
They postulated the presence of a bubble from the start, which can expand and contract as
a result of droplet temperature variations. Moisture profiles within the droplet are tracked
assuming Fickian diffusion of moisture. However, their model does not simulate drying
into the boiling regime, i.e., inflation. Sano and Keey (1982) presented a model that can
simulate inflation but is otherwise similar to that of Wijlhuizen et al.. Results are compared
with experimental measurements of milk drying and good agreement is obtained. As the
model is incapable of simulating inflation-deflation cycles, a key question raised is how to
set the maximum particle size. Without some attempt to do this, the inflating droplet would
continue expanding until it was unrealistically large. This same problem is faced by Hecht
and King (2000b) who present a further model containing a bubble. They introduce a
check which considers the surface tension of the shell and prevents expansion if the pressure
within the bubble is insufficient to over come this. They also impose an arbitrary maximum
particle size to prevent excessive expansion.
Frey and King (1986) presented a model capable of simulating a droplet with multiple
small internal bubbles. This worked by homogenising the entire droplet and dealing with
the equivalent binary, ideal homogeneous mixture. El-Sayed et al. (1990) made experimental measurements of sucrose and maltodextrin solutions, coffee extract and skim milk drying
and applied the model of Frey and King to make satisfactory predictions prior to droplet
boiling.
Minoshima et al. (2001, 2002) developed a relatively simple model to predict the formation of hollow granules. Their model uses the structural properties of the crust to the
predict dried-particle size and shell thickness. Tsapis et al. (2005) considered the electrostatic stabilisation forces between colloidal silica particles to produce impressive simulations
of shell formation and buckling. Neither of these models simulated spatial moisture profiles, but are interesting because they demonstrate how the physical properties of droplets
might be used inform the structural and morphological simulation.

31

CSH

2.5. Scope of this Thesis

2.5 Scope of this Thesis


There have been numerous attempts to simulate droplets drying, as reviewed in Section 2.4.
Whilst these previous models contain several good ideas, there are a number of obvious deficiencies. Droplet moisture content is often resolved spatially, but variations relating to
the solids within the droplet beyond inclusion of an explicit shell region are normally ignored. Morphological development may be simulated by means of the shrinking
core or bubble-based approaches, but no existing model is capable of simulating multiple
dried-particle morphologies; once the use of a particular drying model has been specified,
the type of dried-particle is determined. Related to this inability to predict dried-particle
morphologies is the lack of any clear rational for choosing a particular model for morphological development. Given the importance of dried-particle morphology in determining
the properties of the dried-powder, the combination of these shortcomings amounts to a
serious weakness in existing models.
The new model presented in this thesis attempts to address the major weaknesses identified in existing descriptions of droplet drying. In particular, the new droplet drying model
will:
simulate spatially resolved moisture profiles and, further, be capable of modelling the
local concentration of a dissolved solute;
utilise a population balance approach, based on that developed by Seydel et al. (2006),
to simulate the nucleation, growth and transport of suspended solids;
provide for the inclusion of a centrally located bubble and explicit shell region;
incorporate existing and develop new sub-models to describe morphological development following shell formation;
provide a physical basis for the use of different sub-models following shell formation
and develop coherent criteria for choosing the appropriate description to use.
The combination of these features produces a model that, for the first time, is capable of
describing structural development as influenced by evolving droplet composition and drying
conditions and is thereby able to simulate multiple dried-particle morphologies.
The more sophisticated, stand-alone, droplet drying model developed here gives useful
physical insights into the droplet drying process. These elucidate the relationship between
droplet composition, drying rate and dried-particle structure. Eventually, it is intended
that the model may be incorporated within a comprehensive computational fluid dynamics
simulation of a spray drying tower, or may help with the formulation of improved reduced
models for this purpose.

32

CSH

Chapter 3
A New Model for Drying Droplets Prior
to Shell Formation
In which a new droplet drying model is introduced. This novel drying model incorporates a population balance to describe the suspended solids and uses volume
averaging to obtain appropriate transport equations for the continuous phase. After
demonstrating the formulation and discussing the benefits of this new model, drying
droplets of colloidal silica and sodium sulphate are simulated up until the point of
shell formation. The results obtained are compared with experimental data from
the literature. The material presented in this chapter is based upon work published
in Handscomb et al. (2008a).

3.1 Background
Spray towers require a pumpable feed, (2.2.1) which typically contains suspended solids
or dissolved materials which crystallise during drying, (Masters, 1992). Droplets may also
contain air present in the feed or entrained in the atomiser, (2.4.2) and will therefore, in
general, constitute a three phase system as illustrated in Figure 3.1. The literature contains
a number of methods for describing moisture transport and drying in such systems. Before
proceeding with a detailed exposition of the new droplet drying model, this section briefly
discusses these various approaches. The reasons for selecting a volume averaged description
are expounded and the key results underpinning such an approach are introduced.

3.1.1 Descriptions of Moisture Movement in Porous Media


The drying of individual droplets can be described in a number of different ways. It is
perhaps convenient to consider the various length-scales on which such drying could be
simulated, (Kohout et al., 2006b). At the largest scale, whole droplets can be considered

33

3.1. Background

Vapour and Inert


Gas Phase

Solid Phase
Liquid Phase

Figure 3.1: Schematic showing a generic drying droplet containing suspended solids and gas pockets.

with the drying rate prescribed as some function of droplet size and droplet-averaged quantities such as moisture concentration or temperature. These average quantities are, in turn,
evolved yielding droplet temperature and moisture histories. The characteristic drying curve
and reaction engineering approaches discussed in Section 2.4.3 are examples of modelling
at this scale. Although simple to implement, such models provide very limited information.
At the other extreme, drying can be modelled at scale of individual pores. The traditional approach is to use pore network models, of which Prat (2002) and, more recently,
Metzger et al. (2007) have written thorough reviews. The pore networks in such models
consist of regularly or randomly located pores connected by throats. The geometry of both
pores and throats can be chosen to simulate different aspects of real porous media. Assuming that capillarity is the process controlling moisture transport, drying can be simulated
as an invasion-percolation process, (Yiotis et al., 2005, 2006). The objective of such models is to predict the dependence of parameters such as the permeability, capillary pressure
and effective diffusion coefficient on moisture concentration. The effect of the pore size
distribution on drying rate can also be investigated, although typically only in a qualitative
manner, (Bray and Prat, 1999). This is due to both computational limits on the size of networks that can be simulated and experimental limitations analysing the nano-scale structure
of porous materials.

34

CSH

3.1. Background
A more sophisticated development of pore network models is to work with more realistic
geometries and use a volume of fluid approach5 to track the fluid menisci, (Orr et al.,
1977). Such reconstructed medium methods are capable of simulating three phase, solid
liquidgas systems. For example, tepnek et al. (2001) investigated the effect of porous
structure on the evolution of gas pockets trapped within a granular sludge. The realistic
geometries used are typically assemblies of randomly packed spheres, (e.g., Bryant and
Johnson, 2003; Mayer and Stowe, 2006), but recent developments have seen the move
towards more physically accurate descriptions. Kohout et al. (2006b) digitally encoded
a porous structure recorded by X-ray micro-tomography, thus running simulations on a
geometry identical to that used for comparative experiments.
The most detailed approach to modelling porous media is to perform actual hydrodynamic simulations at the pore-scale. The appropriate laws of physics are applied directly to
determine the moisture distribution as a function of time, (Whitaker, 1977). Each phase in
the system must obey the overall continuity equation,

+ v = 0 ,

(3.1.1)

and, where a phase consists of more than one component, the continuity equation for
species j ,
j
t



+ j vj = rj .

(3.1.2)

Here is the total mass density and v is the mass averaged velocity. j is the mass density
of species j and v j is the velocity of species j with respect to a stationary frame of reference.
The volumetric mass rate of production of species j is represented by r j . The third important physical law is the requirement that linear momentum is conserved. This is expressed
in the equation of motion,

v
t

(3.1.3)

+ v v = g + ,

where g is the acceleration due to gravity and is the total stress tensor. This later quantity
depends on the fluid, with the Navier-Stokes equation being obtained when the Newtonian
constitutive relation is employed.
The equations above are sufficient to describe mass transfer in a porous medium, but
only in conjunction with a complete geometric configuration of the phase interfaces. Even
when such information is available, two problems face workers using such an approach:
5

The volume of fluid method is a numerical technique to track the movement and development of free
liquid surfaces, (Hirt and Nichols, 1981).

35

CSH

3.1. Background
constructing a suitable solution grid and limited computer power, (Benzi, 2003). To a
varying degree, the second of these problems is an issue facing all numerical methods for
hydrodynamic simulation. The difficulty of meshing an intricate pore-space is a major issue for traditional CFD techniques, but can be overcome in the lattice Boltzmann method.
In this approach, the fluid is represented by an array of computational particles and their
associated number density functions. These particles move across a regular lattice with a
discrete set of velocities, according to rules ensuring that the required conservation laws are
satisfied, (Benzi et al., 1992; Succi, 2001). Macroscopic quantities are obtained by averaging over this particle ensemble, with the Navier-Stokes equation being recovered in the
hydrodynamic limit.6 Because a regular lattice is used irrespective of the geometric complexity of the flow domain, this approach is considerably more robust than traditional CFD
approaches in applications like the simulation of pore-scale flow in porous media, (Chen
et al., 2003; Sullivan et al., 2006). However, Vogel et al. (2005) note that some degree of
lattice refinement may be necessary to capture the narrow films of moisture which play an
important role when drying in the pendular regime.
In practice, the detailed geometric description required for hydrodynamic simulation of
porous media is rarely available and may not even be measurable. Moreover, it also likely to
prove unnecessary, as, for most applications, the detailed velocity distribution within each
pore is not required. Rather, it is more important to know the variation of the average
velocity over distances which are large compared with a typical pore diameter. Certainly for
the present application where radial moisture profiles are sought, the coarse fidelity of such
an effective medium length scale is sufficient. To this end, it is chosen to model moisture
movement within the drying droplets at this intermediate scale through the use of volume
averaging.
The fundamental continuity equations describing transport in a porous medium, (3.1.1
3) are point equations they are true everywhere in the porous medium since mass and
momentum must be conserved at each point. The idea behind the volume averaging
technique is to average these fundamental continuity equations over some suitably chosen region, (Whitaker, 1977, 1980; Slattery, 1999). The resulting description is that of a
homogenised- or effective-medium approximating the true porous structure. Volume averaging smears out information about flow at the pore-scale and such homogenised models
cannot be expected to predict phenomena resulting from large-scale heterogeneities in the
liquid phase distribution, such as surface dry spots, (Laurindo and Prat, 1996). However,
spatial concentration profiles for each species in the system can be obtained and, for many
purposes, this is sufficient.
As with most averaging techniques, a number of terms requiring closure are introduced
6

, where is the
The hydrodynamic limit is that in which " 0, (Benzi et al., 1992). Here, " = F
F
mean free path and F is a general macroscopic field. The parameter, ", can be likened to a local Knudsen
number, i.e., a ratio of molecular kinetic to hydrodynamic length scales.

36

CSH

3.1. Background
when volume averaging the mass transport equations for a porous medium. Whitaker
(1977) describes how these can be treated in a rigourous way and this approach has been
used to model textile dyeing, (de Souza and Whitaker, 2003a), catalytic reactions in a
packed bed reactor, (de Souza and Whitaker, 2003b) and drying of porous sandstone, (Wei
et al., 1985a,b). However, in practice, most researchers group the terms requiring closure
into a number of effective transport coefficients such as an effective thermal conductivity,
effective vapour-phase diffusivity, and effective liquid-phase permeability, (Erriguible et al.,
2005; Kohout et al., 2006a,b). These parameters which are typically strong functions
of local microstructure and moisture content can then be measured from well designed
experiments, (Perr and Turner, 1999; Metzger et al., 2007) or obtained by simulations at
the pore space length-scale, (Bray and Prat, 1999; Kohout et al., 2004).
As droplets dry in a spray tower, they form particles and any of the methods discussed
in this section can be applied to model their continued drying. However, a different approach is required whilst the droplets are more accurately described as droplets containing
suspended solids rather than porous particles containing moisture. In this case, the pore
network and reconstructed medium methods are not applicable as they require the identification of a porous structure. Methods based on hydrodynamic simulation can still be
used, but the additional need to describe solid phase growth renders them even less attractive from a computational perspective. Consequently, the new model introduced in this
thesis uses the idea of volume averaging to describe moisture transport in a drying droplet.
Such an approach can handle droplets containing suspended solids and moisture transport
in a porous medium with equal ease. The method is therefore ideally suited to modelling
droplets in spray dryers, where simultaneous drying and particle formation is the defining
feature of the process.

3.1.2 Local Volume Averaging


As discussed in the previous section, the model introduced in this thesis uses local volume
averaging to describe transport within a drying droplet. This section gives an overview of
the basic mathematical concepts underpinning the approach and which are used later in
developing the new model. The theory of volume averaging as applied to porous media is
discussed in detail in several sources, (Slattery, 1967, 1969, 1970, 1999; Whitaker, 1969,
1977, 1980). The reader is referred to these works for more detail on any of the methods
or theoretical results introduced in this section and, consequently, references for individual
results are not supplied.

37

CSH

3.1. Background

Figure 3.2: Every point, z , within the porous medium has an associated averaging surface, S , containing
the volume, V .

Definitions
Every point, z , within the porous medium whether in the solid, fluid or gas has,
associated with it, an averaging surface, S . A spherical averaging surface is illustrated in
Figure 3.2, but any shape can be used, provided that the dimensions and orientation are
invariant. The minimum acceptable size for S is, roughly speaking, such that the average
of the averages taken within S is equal to the average taken at the central point. The total
volume enclosed by S is denoted V , with V (i ) denoting the volume occupied by phase i
within the averaging surface.
Consider a quantity B associated with the phase i , where B may be a scalar, vector or
tensor valued quantity. The average of B within the volume enclosed by S may then be
defined in three ways. Firstly, the superficial volume average for phase i of B is defined
B

(i )

1
V

(3.1.4)

B dV .
V (i)

Note that, where appropriate, the phase with which a given averaged quantity is associated
is written in superscripted parentheses. The superficial volume average may therefore be
understood as the mean value of B (i ) in V . This notation is used throughout the remainder
of the thesis.
The intrinsic volume average for phase i of B the mean value of B (i) in V (i ) is
defined
Z
1
(i )
B (i )
B dV .
(3.1.5)
V
V (i)

38

CSH

3.1. Background
From their definitions, it is clear that the superficial and intrinsic volume averages are related
according to
B

(i )

V (i )

B(i ) .

(3.1.6)

Finally, the total volume average of B over all I phases present the mean value of B in
V is given by
B
=

V
I
X

B dV
V

(i)

(3.1.7)
.

i =1

These three methods of volume averaging will be used frequently in the remainder of this
thesis.
Having defined superficial and intrinsic volume averages, the minimum acceptable size
of the averaging volume enclosed by S can be specified more precisely. If the averaging
surface is characterised by a dimension, L0 , then the minimum acceptable size of S is such
(i )

that B is nearly independent of position over distances of order L0 . As a consequence of


this, the following two relations are true:
(i )

B (i) = B
and

(i)

(3.1.8a)

B(i ) (i) = B(i ) .

(3.1.8b)

Averaging Theorem
The key result which enables local volume averages to be taken of interesting equations7 is
the theorem for the local volume average of a gradient. Consider again the quantity B (i)
and take the superficial volume average of its gradient,8
B

(i )

= B

7
8


B d V =

V (i)
(i)

1
V

1
V


B dV

V (i)

1
V

Z
Bn
dA
Sw

(3.1.9)

Bn
dA .
Sw

Here, interesting equations is taken to mean equations of relevance to transport in porous media.
B
Here B = i j ...k...m
, and B n
is evaluated appropriately.
x
k

39

CSH

3.1. Background

V (c)

Se

Sw

Figure 3.3: Illustration of a porous medium consisting of a continuous and discrete phase. S represents
the boundary of the averaging volume associated with the point, z . The fraction of this volume occupied
by the continuous phase, V (c) , is bounded by walls and exits. Walls are denoted Sw and represent those
portions of the bounding region defined by phase interfaces; exits comprise the remainder of the bounding
surface and are those areas where the boundary of V (c) coincides with S .

A special case of this theorem is


divB

(i )

divB d V
Z
1
(i)
= divB +
Bn
dA ,
V Sw

V (i)

(3.1.10)

where B is to be interpreted as a spatial vector field or second-order tensor field. In these


equations, Sw represents the portions of the surface bounding V (i ) comprised of phase interfaces, termed walls. The remaining regions, i.e., those regions coincident with S , are termed
exits and denoted Se . Figure 3.3 illustrates these terms for a porous system comprising a
continuous and discrete phase.
Volume Averaged Differential Mass Balance
Having defined the volume averages and stated the averaging theorem, it is now possible to proceed with averaging the conservation equations. The key equation for the purposes of the drying model developed below is the continuity equation for an individual
species, (3.1.2). Written for a specific component, A, this reads
A
t

+ AvA rA = 0 ,

(3.1.11)

40

CSH

3.1. Background
and is also known as the differential mass balance for species A. As a point equation, the
differential mass balance is applicable in the solid, liquid and gas phases, although it may
take on simpler forms in some of these. For example, if the solids are assumed to be at rest
and not growing, the differential mass balance for this phase becomes redundant.
Consider a system comprising a continuous liquid and discrete solid phase. It is sought
to associate a local volume average of (3.1.11) with each point in the continuous phase.
Taking the superficial volume average of this equation gives,
1

A
t

V (c) (t )

(3.1.12)

+ AvA rA d V = 0 ,

where V (c) (t ) is volume of the continuous phase within the local averaging surface. Although the solids are presently assumed not to move, V (c) may still change as a function
of time due to solid phase growth through crystallisation. Denoting the velocity of this
growing interface by w, (3.1.12) becomes
d

dt
1

V
Z

V (c) (t )

Sw

A d V

A w n
dA+

1
V

V (c) (t )

AvA d V

1
V

Z
V (c) (t )

rA d V = 0 .

(3.1.13)

Using the definition of a superficial volume average, (3.1.4), it is possible to introduce


(c)
rA

Z
(c)

V (c) (t )

(3.1.14)

rA d V ,

which is the homogeneous rate of production of A by, for example, chemical reaction.
Recognising the first term in (3.1.13) as the time derivative of (c)
, and substituting for r (c)
A
A
gives9
(c)

A
t

1
V

Sw

A w n
d A + AvA

h R
i
Here dtd V1 V (c) (t ) A d V =
fixed point in space.

(c)

dA
dt

(c)

(c)

rA = 0 .

(3.1.15)

(c)

A
t

, because the time derivative of an average is associated with a

41

CSH

3.1. Background
Using the theorem for the volume average of a gradient, (3.1.9), on the third term gives
(c)

A
t

Z
(c)

A w n
d A + AvA +

(c)
AvA n
dA rA = 0

V Sw
Z

1
(c)
+ AvA(c)
A (w vA) n
dA rA = 0 .
t
V Sw

Sw

(c)
A

(3.1.16)
The third term on the left hand side may be written
rA00

1
V

1
V
Z

A (w vA) n
dA

(3.1.17a)

Sw

()

Sw

(3.1.17b)

rA d A ,

where rA() denotes the rate of production of species A per unit area of interface by heterogeneous chemical reactions. rA00 therefore denotes the volume averaged rate at which species
A is produced per unit area of the fluid-solid interface. This gives the final locally averaged
mass balance for species A as
(c)

A
t

(c)

+ AvA(c) rA00 r A = 0 .

(3.1.18)

Alternatively, defining
(3.1.19)

nA = AvA ,

as the mass flux of species A with respect to a stationary frame of reference, (3.1.18) may be
written
(c)

A
t

(c)

(c)

+ nA rA00 r A = 0 .

(3.1.20)

This equation concludes the introduction of required background material. In the following
sections, this knowledge is applied to the development of a new model for single droplet
drying.

42

CSH

3.2. Overview of the New Model

3.2 Overview of the New Model


3.2.1 Model System
Spray drying is unique in that it combines moisture removal with particle formation; any
model for the drying of such droplets must be able to describe both these concurrent process. The model must therefore simulate the liquid phase containing most of the moisture and the suspended or dissolved solids. In general, drying droplets may also contain
a dispersed gas phase, but the present model does not consider this possibility further.
As the feed to the spray dryer must be pumpable, it is reasonable to assume that the
liquid phase is initially continuous with discrete suspended solids. This will change as
a solid dried-particle structure forms, but throughout this work the liquid and solids are
referred to as the continuous and discrete phases respectively.
Continuous Phase
Both the liquid and the solid phases may be complex mixtures, e.g., the crutcher mix feed to
a tower producing detergent powder contain may contain more than ten components, (Appel, 2000). Simulating such a complicated system in detail would be a daunting task and it
is not sought to do so here. Rather, the new droplet drying model introduced in this thesis
approximates the continuous phase as an ideal binary solution consisting of a solvent
normally water and single solute and considers the solids phase to be comprised of a
single component.
The decision to model the continuous phase as a binary solution permits the description of initially liquid droplets where solids crystallise as drying progresses. Whilst clearly
not capturing the great complexity possible in many real systems, this approach does allow
modelling of some key morphological developments associated with dried-particle formation. Further, the ideal binary assumption means that the continuous phase composition
can be described by a single equation, derived from the volume-averaged differential mass
balance for one of the species. This derivation is presented in Section 3.3.
Discrete Phase
The description of the discrete solid phase is key to modelling the morphological evolution of the droplet. As stated above this, along with predicting the drying rate, is one of
the prime deliverables from a droplet drying model. In reality, droplet structure and moisture transport are strongly interdependent and, consequently, the solid phase description
also influences the predicted drying rate and the continuous phase description affects the
morphological simulation. In the present framework, the solid phase is modelled by a population of discrete solid particles, assumed spherical, which evolve according to a population

43

CSH

3.2. Overview of the New Model


Ideal Shrinkage
T (t )
b (t )

Ideal Binary
Solution

R(t )

Vapour
Bubble

Slurry

Shell

Discrete
Solid Phase

Hollow
Shell

Vapour
Bubble
(b)

(a)

Figure 3.4: Schematic showing: (a) the model system; and (b) drying to form a hollow shell.

balance equation. The volume fraction of the solid phase, along with other quantities of
interest, can be extracted and used to inform the predicted morphological development.
The approach used to model the solid phase is discussed in detail in Section 3.4.
Shell and Bubble
In addition to the description of the suspended solids, two additional features are included
in the model enabling simulation of a greater range of experimentally observed dried droplet
morphologies, (2.4.2). The first of these is provision for an explicit shell region, within
which the equations describing the continuous and discrete phases can be modified. As
almost all spray dried droplets form some sort of solid structure, provision for a structured
shell region is essential. Secondly, it is observed that many experimentally observed morphologies are based on hollow particles. Whilst a dispersed gas phase is not included in the
present model, there is provision for a single centrally located bubble. This bubble can expand (and contract) with time, thus allowing the simulation of hollow droplets. Depending
on the nature of the droplet when the bubble expands, inflated shells or uninflated hollow
spheres can be simulated.
Figure 3.4a shows a schematic of the droplet system outlined so far in this section,
highlighting the ideal binary continuous phase, discrete solid phase, external shell region
and central bubble. Figure 3.4b illustrates just one morphological history capable of being
simulated by the new model drying to an inflated hollow shell.

44

CSH

3.2. Overview of the New Model


Droplet Temperature
In general, drying models do not attempt to model the spatially resolved temperature distribution within a droplet. This approximation is generally considered valid provided that the
Biot number is less than 0.1, (Incropera and DeWitt, 2002). The Biot number a ratio
of internal and external heat transfer resistances may be defined
Bi =

hL
drop

(3.2.1)

where h is the heat transfer coefficient of the mass transfer film and drop is the thermal
conductivity of the droplet. Alternatively, Bi may be given as a function of the Nusselt
number,
Bi = Nu

drop

(3.2.2)

where is the thermal conductivity of the film. In this way, the effect of flowing drying
air can be accounted for through correlations for Nu. Farid (2003) gives an example of a
200 m milk droplet drying in an air stream at 90 C with a relative velocity of 1 ms1 . Here,
the Biot number is initially 0.15, and so the temperature distribution within the droplet can
be ignored.
Provided the thermal conductivity of the droplet doesnt change, the Biot number will
only decrease from this initial value as the droplet shrinks. This is confirmed in the simulations of pure water droplets conducted by Oberman et al. (2004). Here, the temperature
profiles start relatively flat, with any variation across the droplet getting increasingly small
with time. However, Farid (2003) points to some systems, such as skimmed milk, where
the thermal conductivity of the droplet decreases by as much as an order of magnitude as
moisture is removed. In such cases, Farid argues that the internal temperature distribution
becomes important as the droplet dries and presents a model based upon this hypothesis.
The simulation results show surface temperatures up to 10 C greater than those averaged
over the relatively large, 2 mm, droplets at low drying air temperatures. However, as the
author himself notes, it is difficult to validate these predictions against experiments. Indeed, the measured experimental temperatures, which are droplet averages, fit the predicted
surface and averaged temperatures equally well.
The effect of decreasing droplet thermal conductivity may, to some extent, be offset by
the effect of surface evaporation. This idea is explored by Chen and Peng (2005), who
conclude that the Bi < 0.1 criterion for uniform droplet temperature may be relaxed for
evaporating particles.

45

CSH

3.2. Overview of the New Model


For the new model presented in this thesis, it was decided not to calculate the internal
droplet temperature distribution. As predicted temperature variations where they are
significant at all are very hard to verify experimentally, it was felt that there was little
to gain from spatially resolving the temperature profiles. Clearly it is important to model
the averaged droplet temperature and the details of how this is done are to be found in
Section 3.5.

3.2.2 Notation
Throughout this work, quantities relating to the solvent, solute and solid components are
given the subscripts A, B and D respectively. Quantities relating to the continuous and
discrete phases are indicated with the superscripts (c) and (d ) respectively. The volume
fraction of the discrete phase is given the symbol, " and, assuming volume additivity, this
means the volume fraction occupied by the continuous phase is (1 ").
The term droplet always refers to the overall droplet being dried, even when physically
this entity might more accurately be described as a particle. This is to enable the term
particle to be used solely in reference to those solid particles suspended within the drying
droplet and described by the population balance equation. Finally, dried-particle is used
to describe the droplet at the end of drying.
The liquid phase mass fraction of component j is denoted j . These mass fractions are
related to the corresponding liquid phase mole fractions by
x WA
A = PJ A
,
x
W
j
j =1 j

(3.2.3)

where W j is the relative molecular mass of species j and J is the number of species in the
phase. For a binary system, this reads
A =

xAWA
xAWA + xB WB

(3.2.4)

Similarly, mass fractions may converted to mole fractions using


/WA
xA = PJ A
,

/W
j
j =1
j

(3.2.5)

which, for a binary solution reads


xA =

A/WA
A/WA + B /WB

(3.2.6)

46

CSH

3.2. Overview of the New Model


Mass and mole fractions of component j in the gas phase are denoted w j and y j respectively.
The mass density of a particular component in phase i is denoted (iA ) . The mass density
of the associated phase is then given by
(i )

J
X

(i )

(3.2.7)

j ,

j =1

where J is again the total number of components in phase i . Mass fractions are related to
mass densities by
(i )

(i )
A

(i )

(3.2.8)

The material density of a given component, A that is, the density of the pure substance
is denoted 0A.
The molar flux of component A with respect to a fixed co-ordinate frame is notated NA.
The corresponding mass flux is nA, and the two are related by
NA =

nA
WA

(3.2.9)

The velocity of this component with respect to the same fixed frame of reference is then
denoted vA, and the mass-averaged velocity of the entire phase is
v=

J
X
j =1

(3.2.10)

j vj .

The mass flux of the component is then given by


(3.2.11)

nA = AvA ,

and the total mass flux of the phase by


n=

J
X

(3.2.12)

n = v .

j =1

Having defined the core notation used, the following three sections detail the mathematical development of the new droplet drying model outlined above. This chapter focusses
on the core features of the new model: the continuous and discrete phase descriptions, together with an analysis of heat and mass exchange with the bulk. The features of the new
model concerned with morphological development after shell formation the shell region

47

CSH

3.3. Continuous Phase Description


and central bubble are introduced in Chapter 4.

3.3 Continuous Phase Description


The composition of a binary solution is uniquely specified once the mass fraction of one
of the components is known. Therefore, it is only necessary to solve the differential mass
balance for the solvent or the solute. The equation for the solute is chosen as this enables
the crystallisation process to be described more easily. Recalling (3.1.20), the differential
mass balance for the solute can be written
(c)

B
t

(c)

(c)

+ nB rB00 r B = 0 .

(3.3.1)

This equation is completely general and not specific to the present problem of simulating
drying droplets. To be implemented in the droplet drying model, the reaction source terms
(c)
rB00 and r B need to be specified, together with an expression for the averaged mass transfer
flux, n(c)
. It is also useful to transform (3.3.1) written in terms of superficial volume averages
B
to its equivalent in terms of intrinsic volume averaged quantities.

3.3.1 Defining the Solids Volume Fraction


Before proceeding with manipulations of (3.3.1), it is first necessary to carefully define ", the
volume fraction occupied by the discrete phase. This is done by introducing the function

(z) =

0
1

if z lies in the continuous phase,


if z lies in the discrete phase,

(3.3.2)

and defining the solids volume fraction at the point z to be the total volume average of this
function i.e., " . Recalling (3.1.6), the following relations are immediately obvious
"=

V (d )
V

=1

V (c)
V

(3.3.3)

where V (d ) and V (c) denote the volume of the discrete and continuous phase within an
appropriately defined averaging surface enclosing total volume, V . From this, it follows
that
(c)

= (1 ") B(c) ,

(3.3.4a)

(d )

= "B(d ) ,

(3.3.4b)

B
B

48

CSH

3.3. Continuous Phase Description


where B is any scalar, spatial vector or tensor valued function associated with the system.
Using these results in (3.3.1) gives


(c)
(1 ") B (c) + (1 ") nB (c) rB00 r B = 0 .
t

(3.3.5)

3.3.2 Source Terms


As there are no chemical reactions, the homogeneous rate of production of solute, r (c)
,
B
is zero. However, the volume average interfacial production rate is non-zero due to the
possibility of a crystallisation process. Relating the mass rate of crystallisation of the solute,
B , to the changing mass of the solid phase gives

"

rB00 = 0D

t

(3.3.6)

,
crys

where 0D is the material density of the solid phase. The local solids volume fraction may
evolve as a result of crystallisation from the continuous phase or spatial transport of existing
solids from elsewhere in the droplet. The volume average interfacial production rate only
contributes to the former process, as indicated by the subscript on the derivative in (3.3.6).
A more detailed expression is given for rB00 in Section 3.4.2, following discussion of the
discrete phase description.
It has been assumed in writing (3.3.6) that there is no change in the total volume on
crystallisation. This is equivalent to saying that the material density of the solute, 0B is
equal to that of the crystallised solid, 0D . Such an assumption is not valid in general, but
is often invoked in the modelling of crystallisation, (Gerstlauer et al., 2002) and is a good
approximation in certain systems. Substituting into (3.3.5) gives

"

(1 ") B (c) + (1 ") nB (c) + 0D
=0 .
t
t crys

(3.3.7)

3.3.3 Transport Term


The next task is to introduce an expression for the intrinsic solute mass flux, nB (c) . Ficks
first law can be used to describe binary diffusion of a gas within a porous medium and it is
current common practice to always use Ficks first law when analysing binary diffusion in
liquids, (Slattery, 1999). Ficks first law may be written

nB = nA + nB B D(AB) B ,

49

(3.3.8)

CSH

3.3. Continuous Phase Description


where D(AB) is a binary diffusion coefficient. Recognising

nA + nB = n = v ,

(3.3.9)

and taking the intrinsic volume average with respect to the continuous phase gives
nB (c) = (c) B v(c) D(AB) (c) B (c) ,

where (c) is the density of the continuous phase. Whilst D(AB) may vary appreciably
within the drying droplet, it seems reasonable to neglect variations within the averaging
volume. This allows the previous equation to be written
nB (c) = (c) B v(c) D(AB) (c) B (c) .

(3.3.10)

The relation given in (3.3.4a) can be used to cast the averaging theorem, (3.1.9), in terms
of intrinsic volume averages,
(i)

(i )

(1 ") B = (1 ") B

(3.3.11)

Bn
dA ,

Sw

which allows (3.3.10) to be written


(c)

nB

(c)

(c)

(c)

= B v

(c) D(AB)
1"

(1 ") B (c) (B) .

(3.3.12)

where the mass tortuosity vector is given by


(B) = (c) B (c) v(c) D(AB) (c) B (c)
(c)

(c)

(c)

(c)

B v + D(AB) B +

(c) D(AB) Z
V (1 ")

Sw

B n
dA .

(3.3.13)
It is often helpful to think of the mass flux in terms of an effective diffusivity tensor, i.e.,
nB (c) = (c) B (c) v(c)

(c)
1"

(e)
D (AB) (1 ") B (c) ,

(3.3.14)

and this is the approach adopted for the present model. Here, the terms on the right hand
side represent convective and diffusive mass transport respectively. The material presented
in this section is well known in the field and the interested reader is referred to Slattery
(1999) for a thorough discussion.

50

CSH

3.3. Continuous Phase Description

3.3.4 The Continuous Phase Equation


Having obtained an expression for the solute mass flux, it is now possible to write the
equation describing the composition of the continuous phase. Substitution of (3.3.14)
into (3.3.7) gives


(1 ") (c) B (c)
t



(e)
+ (1 ") (c) B (c) v(c) (c) D (AB) (1 ") B (c)

"

0
+ D
=0 .
t

(3.3.15)

crys

Assuming spherical symmetry allows this equation to be simplified to


(1 ") (c) B (c)
t

2
(1 ") B (c)
+ 2
r (1 ") v r (c) (c) B (c) r 2 Deff (c)
{z
} |
r {z
r r |
}
A

"

0
+ D
=0 .

t crys
| {z }

(3.3.16)

This is the equation used to describe the continuous phase in the new droplet drying model.
In assuming spherical symmetry, the effective diffusion tensor, D (e)
has been reduced to a
(AB)
scalar effective diffusion coefficient, Deff . In reality, the only way to obtain an expression for
this diffusion coefficient is to use an empirical form measured from experiments.
It still remains to consider appropriate expressions for v r(c) , the volume-averaged, massaveraged velocity in the radial direction the only non-zero component assuming spherical
symmetry. Also, an appropriately volume averaged expression for the continuous phase
density, (c) , is required. These two tasks are addressed in the following sections.

3.3.5 Continuous Phase Density


The density of the continuous phase may be expressed in terms of the material densities of
the solvent and solute 0A and 0B and the mass fractions of each. It is relatively trivial
to show that

(c)

0A0B
0B A + 0AB

(3.3.17)

51

CSH

3.3. Continuous Phase Description


Making use of the fact that mass densities must sum to one, and defining the density ratio
0B

B =

0A 0B

0A 6= 0B ,

(3.3.18)

gives
(c) =

0AB
B + B

(3.3.19)

It is necessary to take the intrinsic volume average of this expression for use in (3.3.16).
This is done by first expressing the point solute mass fraction as
(c)

B ,
B = B (c) +

(3.3.20)

(c)
B is the deviation from the intrinsic averaged mass fraction within the averaging
where
(c)
B  B (c) . Substituting for B from (3.3.20)
volume. In general, it is expected that
in (3.3.19) and taking the intrinsic volume average gives

(c)

0AB

(c)
(c) 1
(c)
B
B + B +
.

(3.3.21)

Expanding the term in square brackets and making use of (3.1.8b) gives

(c) = 0AB

1
B (c) + B

(c)

B (c) + B

B (c)
0AB

2 ,
B (c) + B
B (c) + B
0AB

(c) 2 (c)

A
+O

(3.3.22)

(c)2
A have been dropped. Further, it may be argued that the
where terms of the order
B (c) may also be ignored since
B (c)  B (c) < 1. Alternatively, the
term involving
correction could be incorporated into the effective diffusion tensor in (3.3.14).10 Either way,
the expression used for (c) in the remainder of this work is

(c)

10

0AB
B (c) + B

(3.3.23)

It is possible to expand (3.3.10) in terms of intrinsic means and variations in a similar fashion to that
demonstrated here. The equation can then be re-arranged to develop an expression for the mass tortuosity vector in terms of these fluctuations. This provides some insight regarding the likely magnitude of
the various terms in (B) but does not remove the eventual need for empirical closures. More information
on this approach can be found in Whitaker (1977).

52

CSH

3.3. Continuous Phase Description

3.3.6 The Local Mass-Averaged Velocity


The advective transport term in (3.3.16) is seen to arise naturally from the assumption of
Fickian diffusion. It reflects the continuity requirement for the continuous phase, (3.1.1),
and it is seen from (3.3.9) that this term only vanishes in the case of equi-mass counter
diffusion where nA + nB = 0. For all situations in which the material density of the solute
is not equal to that of the solvent, the local mass-averaged fluid velocity, v (c) , is non-zero
and needs to be calculated. In general, this requires that the equation of motion, (3.1.3), be
solved for the fluid. However, with the assumption of spherical symmetry it is possible to
make use of an overall continuity equation to determine the mass-averaged radial velocity,
v r(c) the sole remaining non-zero component of v (c) .
The simplest way to obtain an expression for the mass-averaged radial velocity is to
consider conservation of continuous phase volume within an averaging surface, S . Recalling
the terms illustrated in Figure 3.3, it is clear that material may only enter or exit the volume
occupied by the continuous phase, V (c) , through the surfaces denoted Se . Assuming volume
additivity in the ideal binary liquid, a volume balance may thus be written
Z

nA

0A

Se

nB

(3.3.24)

n
dA= 0 ,

0B

where it is assumed that V (c) doesnt change with time. Since the fluxes through the surfaces
Sw are both zero,11 it is possible to write
I

nA
S

0A

nB

(3.3.25)

n
dA= 0 ,

0B

where the integral is now over a closed surface. The divergence theorem then gives
I
div
V (c)

nA
0A

nB

0B

(3.3.26)

dV = 0 .

Since the averaging surface, S , was associated with an arbitrary point, z , the integrand
in (3.3.26) must be identically equal to zero. Further, assuming spherical symmetry and
reducing to one dimension gives
1
2

r r

11

r 2

nAr
0A

nB r
0B

!
=0 ,

The flux through Sw will be non-zero due to crystallisation, but it is reasonable to ignore this for the
present discussion since the material involved doesnt cross the averaging surface, S .

53

CSH

3.3. Continuous Phase Description


or, upon integrating,
nAr

0A

nB r
0B

f (t )
r2

(3.3.27)

Taking the intrinsic volume average of this equation and using the 1-D form of (3.3.14) to
substitute for nAr (c) and nB r (c) gives
!
(c)

eff
(1 ") A(c)
v r (c) (c) A(c)
1" r

1
0A

!
(3.3.28)
(c)

f
(t
)
eff
(1 ") B (c)
= 2 .
v r (c) (c) B (c)
1" r
r

0B

Multiplying through by
v r (c) (1 ")

0A0B (1")
(c)

, and collecting like terms yields


Deff

0B A(c) + 0AB (c) r

(1 ") 0B A(c) + 0AB (c)

= (1 ")

f (t )
r2

(3.3.29)
.

Recognising that A(c) = 1 B (c) , and using the density ratio


B =

0B
0A 0B

(3.3.30)

gives the required velocity as


(c)

v r

= Deff

B (c)

B + B (c)

"

1" r

!
+

f (t )
r2

(3.3.31)

The function, f (t ), is determined by the boundary conditions on the droplet. In the absence of a central bubble, the symmetry condition at r = 0 immediately gives f (t ) = 0.
However, this is not always the case. For example, consider the case of a droplet containing
a centrally located bubble of radius b , which is expanding at a given rate. Further, assuming that the spatial gradients of the solute mass and solids volume fractions are zero at the
bubble interface, (3.3.31) gives
db
dt

f (t ) = b

f (t )

b2
db
2
dt

54

CSH

3.3. Continuous Phase Description


Such a situation is discussed in Section 4.5.
Substituting (3.3.31) into the 1-D form of (3.3.14) gives, on taking f (t ) = 0,
(c)

nB r

(c)

= Deff

B (c)

B
B + B (c)

(3.3.32)

Comparison with Ficks law demonstrates that this expression has the standard form for a
diffusive flux. The factor in square brackets modifies the flux to take account of density


differences between the species. As 0A 0B , B , B / B + B (c) 1 and
Ficks law is recovered.

3.3.7 Boundary Conditions


Having derived the equation describing the continuous phase, (3.3.16), and given expressions for all the relevant quantities involved, it only remains to specify the boundary conditions.
In the absence of a central bubble, symmetry considerations require that the gradient of
the solute profile is zero at the centre of the droplet, i.e.,

B (c)

r

(3.3.33)

=0 .
r =0

A full discussion of the boundary conditions once a bubble has formed is deferred to Section 4.5.3.
Basic drying theory, (2.1.2) explains how water will leave a drying droplet at the surface
due to a higher water activity adjacent to the droplet surface than in the bulk. In contrast,
it is assumed that the solute does not leave the droplet at any time, i.e., the solute mass flux
following the receding interface is zero,


nB r (c)

r =R(t )



= nB r (c)

r =R

(c) B (c) (1 ")

dR
dt

=0 .

(3.3.34)

Substituting for the solute mass flux from (3.3.32) and re-arranging gives the solute boundary condition at the droplet surface,

B (c)

r

=
r =R

00
B (c) m
(c) Deff

(3.3.35)

00 is the solvent mass flux from the droplet surface and is related to the rate of
Here, m

55

CSH

3.4. Discrete Phase Description


shrinkage by
dR
dt

00
m
0A

(3.3.36)

00 is calculated is deferred until Section 3.5.


Discussion of how m

3.4 Discrete Phase Description


3.4.1 Population Balance Equation
The discrete solid phase is modelled by a population of particles. These are described by
a particle number density, N , which evolves in time according to a population balance
equation. Ramkrishna (2000) gives the most general form of such a population balance
equation as

+ RN
= D D T + D D T +h .
+ x XN
r
x
x
x
r
r
r
x
r
{z
} |
t |
{z
}

(3.4.1)

In this equation, terms containing an x refer to the internal co-ordinate, and those containing an r to the external co-ordinate. Internal co-ordinates refer to those properties
such as size, shape or even colour which are intrinsic to the particles. In contrast, the
external co-ordinates merely denote the location of the particles in physical space. The
population balance equation allows the particle number density function to evolve in the
multi-dimensional state-space of internal and external co-ordinates via both advective and
random, diffusive, processes. Birth and death terms used to model processes such as
coagulation can be included via the source term, h .
Model Equations
The particles considered in the model are assumed spherical, allowing a single internal coordinate to be used to classify particle size. The particle diameter, L, is chosen as a suitable
internal co-ordinate, with corresponding state space, x = [Lmin , ). Here Lmin is the
minimum stable crystal size a parameter which must be obtained from experimental
measurements.
Assuming spherical symmetry of the drying droplet allows a single external co-ordinate
the radial position of its centroid to completely specify the spatial location of a solid

L
Lmin
particle. This external co-ordinate, r , has the state space r = b + min
,
R

where
2
2
b is the radius of the central bubble and R is the external radius of the droplet. Figure 3.5
illustrates a drying droplet as described by this population balance approach: the solid

56

CSH

3.4. Discrete Phase Description

Solid N (L, r, t )
Particles
L

Continuous
Phase
r
Figure 3.5: Illustration of a droplet with solids represented by a population of spherical particles suspended
in a continuous liquid phase. The number density function shown schematically on the right
characterises the particle population, describing the distributions of particle size and radial location within
the drying droplet.

phase is represented by a population of spherical particles and characterised by the evolving


number density function, N (L, r, t ).
The velocity of particle motion through the state space of internal co-ordinates is denoted by G . The most appropriate form for this linear growth rate depends on the system
being simulated and is obtained from experimental measurements see Section 3.6.2 for
an example. The only assumption made at this stage is that the linear growth rate is independent of crystal size, i.e., G 6= f (L). It is valid to approximate the linear growth rate
as being independent of both particle size and the continuous phase flow if the suspended
particles are smaller than 50 m, (Nagata, 1975).
When describing the evolution of the number density function in the space of internal
co-ordinates, it is assumed that there are no particle sources or sinks, i.e., h = 0 in (3.4.1).
The solid phase may grow through particle growth or nucleation of new particles at the
lower boundary of x , but not by agglomeration or breakage. Allowing such processes is
certainly possible within the framework of the model developed here, but further discussion
of this is beyond the scope of the present thesis.
The solid phase may evolve in physical space through a convective or diffusive process,
or a combination of the two. Thus, combined with the assumptions discussed above, the
general population balance equation, (3.4.1), reduces to the following form used in this
work,

1 2 (d ) 1

N
2
(GN ) + 2
N+
r vr N 2
r D
=0 .
(3.4.2)
t
L
r
r r
r r

57

CSH

3.4. Discrete Phase Description


It is assumed that, prior to shell formation, the solid particles only move by diffusion, i.e.,
v r(d ) = 0. Furthermore it is postulated that once a rigid shell has formed around the droplet,
the solid particles are no longer free to move at all within this shell region. That is, particle
growth is the sole mechanism operating within a solid shell. However, the advective term is
used when modelling shell thickening, as discussed in Section 4.3.
Boundary Conditions
New particles are permitted to form through nucleation. In general the rate of nucleation of
new particles per unit volume per unit time, N0 , may be linked to the boundary condition
at the lower bound of the space of internal co-ordinates by

n
N0 = X
x N (x, r, t ) ,

x x ,

(3.4.3)

is the velocity through the space of internal coordinates and n


where X
x is a local outwards
unit normal to x at the boundary, (Ramkrishna, 2000). Reduced to one dimension for
crystal nucleation and growth the form required for the present model this becomes
N0 = GN (Lmin ) .

(3.4.4)

This nucleation condition is applied at the lower end of x ; it may be understood by


recognising that the nucleation rate must give the flux of particles across this boundary.
The regularity condition is imposed at the upper boundary of x , that is, N 0 as L ,
(Ramkrishna, 2000).
It is assumed that no solids leave the droplet. Instead prior to shell formation
the receding continuous phase menisci between particles drive the solids towards the centre
of the shrinking droplet, (Liang et al., 2001). In the population balance, this is captured
by means of a birth term at the outer boundary, r = R(t ). Again recalling the assumption
that the solids move via a purely diffusive process prior to shell formation, the appropriate
boundary condition is obtained by equating the volume flux from (3.4.2) with the flux
caused by the receding interface, i.e.,
D

N
r

dR
dt

N=

00
m
0A

(3.4.5)

N,

giving rise to the condition



N

r

=
r =R

00
m

D0A

(3.4.6)

N.

00 is the evaporative flux of solvent, A, from the droplet.


Here, m

58

CSH

3.4. Discrete Phase Description


Symmetry implies zero gradient boundary conditions should be employed at the centre
of the droplet until the formation of a central bubble.

3.4.2 Moments of the Population Balance Equation


Full information about the form of the number density at all times is not required; rather
it is sufficient to know how a small number of the moments of the population evolve. The
a th integer moment of the internal co-ordinate is defined by
Z

La N (r, L, t ) d L ,

ma (r, t ) =

(3.4.7)

Lmin

and several quantities of interest are related to the lower such moments, (Hounslow et al.,
1988). The zeroth moment, m0 , gives the total number of particles per unit volume, whilst
the first moment, m1 , is related to the total length per unit volume. The average solid
particle size at a given spatial location is therefore
L=

m1
m0

(3.4.8)

Similarly, the total surface area per unit volume is related to the second moment of the
particle number density, with a scaling factor derived from the assumption that all particles
are spherical,
(3.4.9)

SV = m2 .

The key variable of interest when describing the solid phase is the local solids volume
fraction, " (r ), which appears in the equation describing the continuous phase, (3.3.16).
The solids volume fraction can be calculated from the particle number density by integrating over all particle sizes,
"=

N L3 d L =
Lmin

(3.4.10)

m3 ,

where it is seen that " is directly related to the third moment of the number density function.
As discussed in Section 3.3.1, " is an inherently averaged quantity and so it is appropriate to
use the solids volume fraction from the moments of the number density function directly
in the volume averaged continuous phase equation.
The evolution of the moments can be obtained directly without the need to solve the
whole population balance equation. Substituting from (3.4.2) into (3.4.7) and differentiat-

59

CSH

3.4. Discrete Phase Description


ing gives
ma
t

La

Lmin
Z

N
t

La

=
Lmin
Z

=
Lmin

=L

(3.4.11)

dL

1
r2 r

GN |
Lmin

(GN ) +

1
r2 r

(L GN ) aL

Z
r 2 v r(d )

a1

r 2 v r(d ) N

GN

L N dL r D
1

+ aG ma1

r 2D

r2 r

N
r

dL

dL

Lmin

r2 r

r 2 v r(d ) ma

L N dL
Lmin
2

r D

ma

where it has been assumed that the particle growth rate, advection velocity and diffusion
coefficient are independent of particle size. The first of these assumptions is fairly common, (Rosenblatt et al., 1984), and a relaxation of the later is discussed in Section 3.4.3
below. Applying the nucleation and regularity boundary conditions at the lower and upper
boundaries respectively gives
ma
t

Lamin N0

+ aGma1

1
r2 r

r 2 v r(d ) ma

1
r2 r

ma
r

(3.4.12)

For some drying droplets, particularly those with high initial solids loading, the particle
inception rate, N0 , will be equal to, or close to, zero. Consequently in these systems the
number density of the smallest particles will vanish and the total number of particles in the
system will not change with time.
Recall that an expression for the time evolution of the solids volume fraction is sought.
From (3.4.10) it is clear that the evolution of " is related to that of the third moment by
"
t

m3
6 t

(3.4.13)

The general moment evolution equation, (3.4.12), then shows that a hierarchy of moment
equations is required in order to obtain this. Namely, it is necessary to solve the following

60

CSH

3.4. Discrete Phase Description


closed set of four partial differential equations,
"
t
m2
t
m1
t
m0
t

L3min N0

r 2 v r(d ) "

"

+ 2
D
r
r2 r
r r

1 2 (d ) 1
m2
2
= Lmin N0 + 2Gm1 2
r v r m2 + 2
D
r
r r
r r

m1
1 2 (d ) 1
= Lmin N0 + Gm0 2
r v r m1 + 2
D
r
r r
r r

1 2 (d ) 1
m0
= N0 2
r v r m0 + 2
D
.
r
r r
r r
=

Gm2

(3.4.14a)
(3.4.14b)
(3.4.14c)
(3.4.14d)

In Section 3.3.2, the solute evolution equation, (3.3.7) was shown to contain a source
term describing the rate at which solute crystallises to the solid phase. Having derived the
relevant moment equations, it is now possible to revisit (3.3.6) and express this source term
as

"

00
0
rB = D

t crys

0
3

= D
L N0 + Gm2 .
(3.4.15)
6 min
2
Here, "/ t |crys , the rate at which the solids volume fraction evolves due to crystallisation,
is identified with the first two terms on the right hand side of (3.4.14a).
The general moment evolution equation, (3.4.12), allows the calculation of more than
these first four moments of the population balance indeed, any number of moments
could theoretically be evaluated. Whilst this would give a more detailed representation of
the number density function, N , the information provided by the limited set stated above is
sufficient for the present application. For example, one of the main uses of this information
will be in predicting the structural properties of the shells formed. It is often difficult
to measure the functional dependence of such properties on solids fraction alone and so,
in such cases, further knowledge of the size distribution of particles constituting the shell
would be of limited practical interest.
The boundary conditions for the moments are obtained by inserting (3.4.6) into (3.4.7)
and integrating. At the outer edge of the droplet, the following conditions are then easily
obtained,

00
mi
m
=
mi for i = 0, 1, 2 ,
(3.4.16)

r r =R 0AD

00
00
m
"
m
=
and
m3 = 0 " .
(3.4.17)

r r =R 6 0AD
A D
61

CSH

3.4. Discrete Phase Description


Symmetry implies that the spatial gradients of all the moments are zero in the centre of the
droplet, prior to the formation of a bubble.
It is often useful to express moment related quantities in terms of their value integrated
over the whole droplet, rather than on a volumetric basis. Such quantities are easily obtained
by integrating the moments over the full range of the external co-ordinate,
Z
Mi =

mi d V
Vd
ZR

4r 2 mi (r ) d r ,

(3.4.18)

where M i is the i th moment of the internal co-ordinate summed over the entire droplet
volume, Vd . For example, the integral of the zeroth moment over the entire droplet gives
the total number of solid particles. The rate at which the number of particles changes is obtained by differentiating (3.4.18) and substituting from (3.4.14d). Assuming no advective
motion or bubble growth,
dM0
dt

R(t )

4r 2 m0 (r ) d r
b
2

= 4 R m0 (R)

dR

m0

dr
t


1
m0
2
2
2
r N0 2
= 4 R m0 (R)
+
r D
dr
dt
r
r r
b

R !
ZR

m
dR
0
= 4 R2 m0 (R)
+
r 2 N0 d r r 2 D
,
dt
r b
b
dt
dR

b
ZR

which, on substituting the boundary conditions on m0 from (3.4.16) gives


dM0
dt

= 4

r 2 N0 d r ,

(3.4.19)

that is, the change in the total number of solid particles is as expected the local
volumetric nucleation rate integrated over the entire droplet volume.

3.4.3 Size Dependent Diffusion Coefficient


In some applications, it may not be acceptable to assume the solids diffusion coefficient is
independent of particle size. Instead, the Stokes-Einstein equation might be used to give a

62

CSH

3.4. Discrete Phase Description


relationship between particle diameter, L and diffusion coefficient,
kTd

D=

(3.4.20)

3L

Here, k is Boltzmanns constant, with Td and representing the temperature and viscosity
of the continuous phase respectively. Substituting this expression for D into (3.4.2) gives
N
t

(GN ) +

1
2

r r

r 2 v r(d ) N

1
2

r r

r2

kTd N
3L r

=0 .

(3.4.21)

Taking the discrete phase advection velocity, v r(d ) , to be zero as is assumed to be the case
prior to shell formation gives, on substituting into (3.4.11) and proceeding as in the
previous subsection,
ma
t

Lamin N0

+ aGma1 +

1 kTd
r 2 3 r

ma1
r

(3.4.22)

Again, the linear growth rate has been assumed independent of particle size and the assumption of a uniform droplet temperature allows Td to be taken outside the spatial derivative.
The hierarchy of moment equations required to solve for the solids volume fraction, ", is
therefore

kTd 1
" 3
m2
2

= Lmin N0 + Gm2 +
r
(3.4.23a)
t
6
2
18 r 2 r
r

m2
kTd 1
m1
2
2

= Lmin N0 + 2Gm1 +
r
(3.4.23b)
t
18 r 2 r
r

m1
kTd 1
m0
2

= Lmin N0 + Gm0 +
r
(3.4.23c)
t
18 r 2 r
r

m0
kTd 1
m1
2
= N0 +
r
.
(3.4.23d)
t
18 r 2 r
r
Using the Stokes-Einstein equation to give the diffusion coefficient a dependence on the
internal co-ordinate results in an unclosed equation set (c.f. (3.4.14)a-d). The final term
in (3.4.14d) means that the evolution equation for the zeroth moment depends upon the
unknown value m1 ; it is necessary to approximate m1 to close these equations.
A possible solution to this closure problem is to use an extrapolative closure. One such
scheme is based on simple linear extrapolation of the moments when plotted on a logarithmic scale. This gives
m1 =

m02
m1

(3.4.24)

63

CSH

3.5. Heat and Mass Transfer From The Droplet


which, combined with (3.4.22) for a {0, 1, 2, 3}, gives a closed system of four partial
differential equations for the discrete phase. This interpolative approach has been shown to
be accurate in a number of different applications, (Frenklach, 2002), and it is believed to
be a good approximation in the present case. Nevertheless, a full numerical study would be
needed to prove this conclusively.
The boundary conditions on the moments in the case of a size dependent diffusion
coefficient are

00 3
mi
m
= 0
(3.4.25)
mi +1 for i = 0, 1, 2 ,

r r =R A kTd

00 3 m32 m
00 18 "2
"
m
and
=
=
,
(3.4.26)

r r =R 6 0A kTd m2
0A kTd m2
where the extrapolation m4 = m32 /m2 has been used in the last equation to close the system.
As before, symmetry boundary conditions are applied at the centre of the droplet.
The example presented in Section 3.6.2 uses the closure introduced in this section to
model the drying of a droplet of sodium sulphate solution. As discussed, this should be
done with care. The quality of the closure will likely be system dependent and vary depending on the relative magnitudes of the nucleation, growth and spatial diffusion rates. The
likely quality of the closure for the initially saturated sodium sulphate system is discussed in
Section 3.6.2.

3.5 Heat and Mass Transfer From The Droplet


The previous two sections introduced the equations describing transport of the continuous
phase and suspended solids within a drying droplet. It was seen that, in obtaining suitable
00 , or total rate of evapoboundary conditions for these equations, the solvent mass flux, m
vap , were required. In this section, expressions for these quantities
ration from the droplet, m
are derived.

3.5.1 Evaporation from a Sphere


When considering mass transfer in mixtures, it is traditional to start from the MaxwellStefan equations for multi-component mass transfer,
yi =

n
yi N j y j Ni
X
j =1, j 6=i

C t Di j

(3.5.1)

64

CSH

3.5. Heat and Mass Transfer From The Droplet


Nair

Stagnant
Air

NA

2R

Figure 3.6: Schematic illustrating the fluxes from an evaporating drop of A surrounded by stagnant air.

Here the equation is written in terms of mole fractions, yi , molar fluxes, Ni and the total
molar concentration, C t . Assuming a binary mixture a fair approximations in a spray
drying tower where the air behaves like a single component this equation reduces to
yA =

yANair yair NA

C t DA,air

NA = NA + Nair yA C t DA,air yA ,

(3.5.2)

which is the familiar mass transport equation for a binary mixture. For the present model,
it is more convenient to recast (3.5.2) in terms of mass fractions. Substituting from (3.2.4)
and (3.2.9) gives, after re-arrangement,

nA = nA + nair wA DA,air wA .

(3.5.3)

A number of assumptions are now made concerning the drying droplets. Firstly, it is reemphasised that the droplets are assumed to remain spherical at all times. This is important
as shape appreciably affects the rate of droplet evaporation, (Michaelides, 2006). The drying
system assuming spherical symmetry is shown schematically in Figure 3.6.
The separation between the droplets is assumed large compared with their diameter; the
evaporation characteristics from single droplets differ from those within a spray, (Masters,
1992). Finally, it is assumed that vapour is transported away from the droplet by pure Stefan
flow, i.e., nair = 0. This allows (3.5.3) to be written
nA =

DA,air
1 wA

(3.5.4)

wA .

65

CSH

3.5. Heat and Mass Transfer From The Droplet


The assumption of spherical symmetry leads to the continuity requirement

d
4r 2 nA = 0 ,
dr

(3.5.5)

which, on substituting for the mass flux from the 1-D form of (3.5.4), integrates to give
4r 2

DA,air dwA
1 wA dr

(3.5.6)

vap ,
=m

vap is the total rate of solvent mass lost from the droplet. It is now sought to
where m
integrate this expression from the surface of the droplet out to the bulk.
Under isothermal conditions, the product DA,air is constant and may be taken outside the integral. However, Frank-Kamenetskii (1969) demonstrated that, even when the
temperature varies, it is still possible to write
Z

T
D A,air R

r DA,air

(3.5.7)

dr ,

where D A,air is the averaged diffusion coefficient from


D A,air =

1 T
dr
R r 2 DA,air

(3.5.8)

and T is suitably chosen to represent the average temperature in the gas film surrounding
the evaporating droplet; Yuen and Chen (1976) recommend using
T = Td +

1
3


T T d .

(3.5.9)

The density in (3.5.4) can be similarly averaged and this is denoted . Integrating then
gives
Z

vap
m
R

dr
r2

Z
= 4D A,air

wA,
wA,sur

d wA
1 wA

vap = 4RD A,air log (1 + BM ) ,


m

66

(3.5.10)

CSH

3.5. Heat and Mass Transfer From The Droplet


where BM is the Spalding mass transfer coefficient,12 defined as
BM =

wA,sur wA,
1 wA,sur

(3.5.11)

This coefficient is a dimensionless expression of the driving force for mass transport, (Spalding, 1963), so that it is always possible to write
(3.5.12)

nA = k BM ,

where k is a mass transfer coefficient based upon a mass fraction driving force. Recognising
that
vap =
m

1
4R2



nA

r =R

(3.5.13)

and substituting from (3.5.10), it is seen by inspection that


k =

D A,air log (1 + BM )
R

BM

(3.5.14)

i.e., the mass transfer coefficient depends upon the driving force, BM a well known result
when dealing with high rates of mass transfer, (Bird et al., 1960).
In the limit of a low driving force, i.e., when BM  1, the expression for the total mass
vaporisation rate, (3.5.10), reduces to
(3.5.15)

vap = 4RD A,air BM .


m

Comparison with (3.5.12) using (3.5.13) allows the identification of k , the mass transfer
coefficient in the limit of a low driving force,
k =

D A,air
R

(3.5.16)

which, it is seen, does not depend on BM .


It is often more convenient to express the mass transfer coefficient, k , in terms of the
Sherwood number.13 Using transfer coefficients based on mass fraction driving forces, the

12
13

The Spalding mass transfer coefficient is sometimes referred to as the Blowing coefficient, (Michaelides,
2006).
The Sherwood number is a dimensionless number representing the ratio of convective to diffusive mass
transport rates.

67

CSH

3.5. Heat and Mass Transfer From The Droplet


Sherwood number may be defined
Sh =

(2R) k
D A,air

(3.5.17)

which, in the limit of a low driving force, is


Sh0 =

(2R) k
D A,air

(3.5.18)

=2 .

With a low driving force, the mass vaporisation rate, (3.5.10), may therefore be written as
vap = 2RSh0 D A,air log (1 + BM ) .
m

(3.5.19)

Modifications to account for the effect of bulk gas flow and high mass transfer rates are
discussed in the next section.

3.5.2 Blowing Effects


As is well known, the rate of droplet evaporation is greater when drying in moving air
than when the surrounding gas is stagnant. This observation can be explained by thinking
of mass transfer from the droplet in terms of film theory. Film theory is based upon the
hypothesis that the resistance to mass transfer between a surface and the bulk gas may be
analysed through the introduction of a gas film adjacent to the surface, (Bird et al., 1960;
Slattery, 1999). In stagnant gas, this film thickness, M is constant. However, if the droplet
is moving in the gas stream, the film narrows and the resistance to mass transfer decreases.
There exist a number of common correlations to describe this effect, (e.g., Frssling, 1938;
Ranz and Marshall, 1952), giving Sh = f (Re, Sc). In this work, the correlation of Clift et al.
(1978) is used,
1

(3.5.20)

Sh = 1 + (1 + ReSc) 3 f (Re) ,

where

f (Re) =

Re 1 ,

Re0.077

Re 400 .

(3.5.21)

The thickness of the mass transfer film can also be influenced by the evaporation process
itself. At high rates of mass transfer, the radial velocity field associated with the flow of
vapour from the droplet to the bulk sometimes termed Stefan convection acts to

68

CSH

3.5. Heat and Mass Transfer From The Droplet


thicken the boundary layer, (Schlichting, 1979). This is the so-called blowing effect at
high evaporation rates. Abramzon and Sirignano (1989) propose that this boundary layer
thickening can be described by a correction factor of the form
Fm =

M
M 0

= (1 + BM )0.7

log (1 + BM )
BM

(3.5.22)

A similar correction factor is introduced for the heat transfer film, with BM replaced by BT ,
the Spalding heat transfer coefficient defined as
BT =

C p, A T Td
Q

H vap + m

(3.5.23)

vap

where C p, A is the average vapour specific heat in the film and Q is the total heat penetrating
the droplet. A modified Sherwood number, taking into account blowing effects, is then
defined by
Sh = 2 +

(Sh 2)
FM

(3.5.24)

where (3.5.20) gives the dependence of Sh on the velocity relative to the drying gas. Replacing Sh0 with Sh in (3.5.19) then gives the expression used for the evaporation rate in
this work,
vap = 2D A,air Sh R log (1 + BM ) .
m

(3.5.25)

3.5.3 Thermal Calculation


Although the current model does not seek the temperature distribution within a drying
droplet, (3.2.1), the evolving droplet temperature, Td , is very important. This section
outlines the theory used to evaluate the heat transfer to the droplet from the drying gas and
concludes by giving the algorithm used to obtain both the mass vaporisation rate and the
heat penetrating the drying droplet.
In the previous sections, an expression for the mass vaporisation rate was derived by
considering mass transfer driving forces. This problem could instead be approached by
considering a heat balance on the droplet; the material evaporated has required an quantity
of heat expended per unit mass, (Michaelides, 2006). Following a method similar to that

69

CSH

3.5. Heat and Mass Transfer From The Droplet


detailed in Section 3.5.1, the following expression is obtained,
vap = 2
m

air
C p, A

Nu0 R log (1 + BT ) ,

(3.5.26)

where air is the thermal conductivity of the gas mixture in the film, C p, A is the average
specific heat capacity of the vapour in the film and BT is the Spalding heat transfer coefficient
defined in (3.5.23). Re-arranging this definition, it is seen that the heat penetrating the
droplet, Q , can be obtained from


C
T

T

d
p, A

vap
Q=m
H vap Td .
BT

(3.5.27)

As expected from the heat and mass transfer analogy, the discussion in Section 3.5.2 on
the effect of blowing maps across to the present problem of heat transfer to the droplet. The
effect of movement relative to the drying gas is described by, (c.f. (3.5.20))
1

(3.5.28)

Nu = 1 + (1 + RePr) 3 ,

where f (Re) is again given by (3.5.21). The modified Nussult number is then given by,
(c.f. (3.5.24))
Nu = 2 +

(Nu 2)
FT

(3.5.29)

where FT is given by an equation analogous to (3.5.22). This means (3.5.30) can be rewritten
vap = 2
m

air
C p, A

Nu R log (1 + BT ) ,

(3.5.30)

which, comparing with (3.5.25), shows that the Spalding heat and mass transfer coefficients
are related by
BT = (1 + BM ) 1 ,

(3.5.31)


C
1
p, A Sh
=
.

Nu Le
Cp

(3.5.32)

where

70

CSH

3.5. Heat and Mass Transfer From The Droplet


This relationship between the heat and mass transfer numbers reflects that the two transport processes are coupled in the droplet drying system. The equation system can be solved
iteratively to calculate the evaporation rate, and the heat penetrating the droplet. Algorithm 3.1, based on that given by Abramzon and Sirignano (1989), shows the way in
which these equations are solved in the current model.
To calculate the rate of mass transfer from the droplet, it is necessary to first obtain
the saturated vapour pressure of water at the droplet surface, (line [3.1]-2) and in the bulk,
(line [3.1]-3). Together with the sorption isotherm and the bulk relative humidity, this then
enables calculation of wA,sur and wA, , which are used in the definition of BM , (3.5.11). The
Goff-Gratch equation, (Goff and Gratch, 1946; Goff, 1957) is recommended for use when
determining the saturation vapour pressure of water, (WMO, 1988a,b). Although fitted to
values measured over a flat surface, it is assumed to be a fair approximation for spherical
droplets. The equation is

log10

pAsat
102

=0.78614 + 10.79574 1

Tst
T

5.02800 log10 10


8.2969

T
1
Tst

Tst

(3.5.33)

+ 1.50475 104 1 10



T
4.76955 1 Tst
3
1 ,
+ 0.42873 10
10

where the temperature T must be supplied in Kelvin and Tst = 273.15. This equation is
valid from 50 to 102 C, (Gibbins, 1990); above this temperature range a polynomial is
fitted to data reported by Lide (1992), giving
pAsat
103

=1.410867 0.1017841T + 7.455807 103 T 2 6.734715 105 T 3


7

+ 8.868928 10 T + 1.577971 10 T 1.727391 10

12

(3.5.34)

T ,

where the temperature in this equation must be supplied in degrees Celsius. Figure 3.7
shows both functions plotted with experimental data up to a temperature of 300 C. The
polynomial in (3.5.34) gives a better fit to the data at high temperatures and, consequently,
the routine S_ uses this expression to obtain pAsat (T ) for T 102 C.
Validation
The heat and mass transfer algorithm used for the new model, (3.1) was tested by simulating the drying of a pure water droplet and comparing against data from the literature.
Experimental droplet mass and temperature profiles from Werner et al. (2008c) were used
as the authors give a thorough description of their drying apparatus. In particular, the dimensions of the filament used to suspend the droplet and the thermocouple used to record

71

CSH

3.5. Heat and Mass Transfer From The Droplet


Algorithm 3.1 The procedure used to calculate the heat and mass transfer between a drying
droplet and the bulk, (Abramzon and Sirignano, 1989).
1: procedure Q_(Td , T , P, wA, , HR , BTold )

2:

sat
pA,
S_ Td
sur

3:

sat
pA,
S_ (T )

4:


sat
pA,sur S_ B (c) , " r =R , pA,
sur

5:

sat
pA, HR pA,

6:

yA,sur pA,sur /P

7:

wA,sur

8:

calculate , C p, A, C p , , , D A,air , Le, Pr, Sc



Re air vrel R /air

9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25:
26:

yA,sur WA
yAWA+yair Wair

if Re 1 then
f (Re) = 1
else
f (Re) = Re0.077
end if
1
Nu 1 + (1 + Re Pr) 3 f (Re)
1
Sh 1 + (1 + Re Sc) 3 f (Re)
calculate BM
FM (1 + BM )0.7 log (1 + BM ) /BM
Sh 2 + (Sh 2) /FM
vap
calculate m
BT BTold

. Using (3.5.11)

. Using (3.5.10)

Nu 2 + (Nu 2) /FT


C p, A Sh 1

Nu Le
Cp

28:
29:

BTold BT

30:

. Evaluated at T given by (3.5.9)

repeat
BTold BT

0.7

FT 1 + BTold
log 1 + BTold /BTold

BT (1 + BM ) 1



old
until BT BT < "

27:

. from (2.1.2)


T

T
C

d
p, A
vap
Qm
H vap Td
BT

vap , Q
31:
return m
32: end procedure

72

CSH

3.5. Heat and Mass Transfer From The Droplet

25

6
5
4

P / kPa

Pressure / MPa

20
15
10
5
0
0

10

20

30

40

50

60

T / oC

3
2

GoffGratch
Polynomial

1
0
0

50

100

150

200

250

300

Temperature / C
Figure 3.7: Plot of the saturated vapour pressure of water. Symbols show experimental data from Lide
(1992) and lines are predictions from the Goff-Gratch equation, (3.5.33) and a polynomial fit, (3.5.34).
The inset shows the vapour pressure at lower temperatures on an expanded scale.

the temperature are given. Towards the end of drying, as much as 35% of the total heat
penetrating the droplet is supplied by these items, (Werner et al., 2008c; Neic and Vodnik,
1991). To account for this, the filament and thermocouple were modelled as semi-infinite
rods, for which Incropera and DeWitt (2002) give the expression
Qtherm/fil =


L L Tgas Td ,
2

(3.5.35)

where h and are respectively the heat transfer coefficient to and thermal conductivity of
the rod. For cylindrical rods, the appropriate characteristic dimension, L, is the diameter. In the comparative experiments, the diameters of the filament and thermocouple were
0.31 mm and 70 m respectively; the thermal conductivities of the filament and thermocouple were 1.14 Wm2 K1 and 91.7 Wm2 K1 . The thermocouple comprised two wires
entering the droplet, each of which conducted heat. The heat transfer coefficients to the
filament and thermocouples will depend on the precise geometry and flow which are unknown. Consequently, it is assumed that these heat transfer coefficients will be the same as
calculated for the droplet, i.e., having obtained Nu from line 25 in Algorithm 3.1, the heat

73

CSH

3.6. Applications
transfer coefficient for use in (3.5.35) is given by
h=

Nu gas
2R

(3.5.36)

Figure 3.8 shows a comparison of simulated and measured temperature and mass profiles for a 6 l pure water droplet. The droplet was suspended from a filament for all
experimental runs, but the thermocouple was only in place for those runs used to obtain the temperature profile. The data from Werner et al. (2008c) refers to drying in air
at Tgas = 40 C, flowing with a velocity of 0.3 ms1 and with HR = (3.75 2)%. Since
the temperature profile in particular is quite strongly dependent on the ambient relative
humidity, simulations were performed across the range covered by the experimental uncertainty. Given likely experimental uncertainties in the thermal readings (not reported), the
fits between simulated and experimental temperature and mass profiles are good.
Having validated this part of the code, the remaining sections in this chapter present
simulations of physical systems prior to the formation of a surface shell.

3.6 Applications
Having introduced the new droplet drying model, this section presents two applications
which validate the model formulation and demonstrate some of its core features. The simulations presented in this chapter are run up until the point at which a shell is predicted
to form. This is deemed to happen once the solids volume fraction at the surface of the
droplet exceeds a certain critical value, "crit . The rationale behind this choice of trigger for
shell formation is discussed in detail in Section 4.1.
In general, "crit is a system-dependent parameter, but it is worth considering the range of
values it might take. For uniformly sized spheres, the highest theoretically possible packing
density is 74%, corresponding to close-packing in space, (Hales, 1992). Particles in drying
droplets are unlikely to adopt this most efficient arrangement; rather it is more realistic to
assume random packing of spheres. Random close packing of spheres in three dimensions
is not a well defined arrangement, with possible packing densities covering the range 0.06
to 0.65, (Jaeger and Nagel, 1992; Torquato et al., 2000). However, capillary consolidation
during drying makes figures towards the top of this range are more likely.
In most systems of practical interest, the suspended solids will not be mono-sized spheres.
For non-spherical particles, higher packing fractions are achievable, similarly when there is
a range of particle sizes. In such cases it is necessary to resort to experimental observations
to chose a suitable value for "crit .

74

CSH

3.6. Applications

45

Droplet Temperature / oC

40

Experiment
RH=1.75%
RH=3.75%
RH=5.75%

35
30
25
20
15
10
0

100

200

300

400

500

600

700

800

Time / s

(a)
6
Experiment
RH=1.75%
RH=3.75%
RH=5.75%

Droplet Mass / mg

0
0

100

200

300

400

500

600

700

800

Time / s

(b)
Figure 3.8: Comparison of simulated and measured (a) temperature and (b) mass profiles for the evaporation of a pure water droplet. Experimental data is taken from Werner et al. (2008c) and refers to a 6 l
droplet drying in air at Tgas = 40 C and flowing with a velocity of 0.3 ms1 . The experimental relative
humidity was reported as HR = (3.75 2)% and the plots show results from simulations run across this
range.

75

CSH

3.6. Applications

3.6.1 Drying a Droplet of Colloidal Silica


The first test case simulated is the drying of a droplet of water-borne colloidal silica. Colloidal silica is a dispersed system in which silica constitutes the dispersed phase. The silica
in question is in the colloidal size range, i.e., less than 1 m so that the particles are unaffected by gravitational forces, but larger than 1 nm so that the dispersion differs from a
true solution, (Bergna, 2006). Such a stable dispersion of solid colloid particles in a liquid
is called a sol.
Gelation occurs as the concentration of colloidal silica is increased. A gelling sol first
becomes viscous before developing rigidity, (Bergna, 2006). There is therefore no volume
change on gelation, nor is there an increase in the silica concentration in any macroscopic
region, (Iler, 1979). The mass fraction at which a gel forms is dependent on the size of
the suspended particles, (Roberts, 2006); the smaller the particles, the lower the gelling
concentration. In the system simulated here the droplets contain 16 nm particles which
gel at mass fractions above 40%. Neic and Vodnik (1991) report that the solids diffusion
coefficient in this gel phase is
D = exp

28.1 + 282 A

1 + 15.47 A

(3.6.1)

where A is the mass fraction of water, given by


A = 1 D =

(1 ") 0A
"0D + (1 ") 0A

(3.6.2)

The material density of the suspended silica particles, 0D , is 2250 kgm3 . This value, taken
from Roberts (2006), represents the average density of silica calculated from a number of
commercially available silica sols; the density variation between sols from different manufacturers is found to be small, i.e., 40 kgm3 .
The droplet simulated had an initial silica mass fraction of 30%, corresponding to an
initial solids volume fraction, "0 , of 0.16. Prior to gel formation that is, at solids volume
fractions below 0.23 experimental observations suggest that internal convection currents
keep the drying droplets well mixed, (Neic and Vodnik, 1991). To simulate this, the solids
diffusion coefficient is set to 107 m2 s1 a relatively large value at low solids concentrations. The discontinuity at the point of gel formation was found to cause problems for
the numerical solution routine and so was smoothed using a hyperbolic tangent weighting
function. The expression used for the solids diffusion coefficient is therefore

28.1
+
282
1

D D = f D 107 + 1 f D exp

, (3.6.3)
1 + 15.47 1 D

76

CSH

3.6. Applications

Solid Diffusion Coefficient / m s

2 1

10

10

10

10

12

10

14

10

0.2

0.4

0.6

0.8

Solid Mass Fraction

Figure 3.9: Plot of the solids diffusion coefficient used to simulate the drying of a colloidal silica droplet.

where
1

f D =
1 tanh 20 D 0.35
.
2

(3.6.4)

A plot of the solids diffusion coefficient is shown in Figure 3.9.


To simulate the pure water continuous phase using the current model wherein the
continuous phase is assumed to be an ideal binary solution the solute mass fraction is
initialised to a very small number, 109 . There is no particle nucleation or growth in this
example and so there is only one particle size at all times. This considerably simplifies the
solids description and makes this an ideal initial test case.
The simulation results prior to shell formation are compared with experimental results
from Neic and Vodnik (1991) at two different drying air temperatures. For dispersed
colloidal silica, (i.e., not aggregated using flocculants), Guo and Lewis (1999) conducted
experiments demonstrating that gravity driven settling produced beds with a solids volume
fraction around 0.6. Capillary driven consolidation was found to increase the packing
fraction towards the random packing limit. The consolidated packing fraction was found
to have little dependence on the initial solids concentration. For the present example, the
critical solids volume fraction representing the formation of surface shell is therefore taken
as "crit = 0.65.
The heat conducted to the particle via the suspending filament is modelled as described

77

CSH

3.6. Applications
in Section 3.5.3, using (3.5.35):


Qfil = L L Tgas Td .
2

(3.6.5)

The glass filament balance used by Neic and Vodnik (1991) was 0.3 mm in diameter and
is here assumed to have a thermal conductivity of 1.14 Wm2 K1 .
Effect of the Sorption Isotherm
As described in Section 3.5, the moisture vapour pressure above the surface is required
to calculate the mass vaporisation rate from the droplet. Since suspended insoluble solids
have a negligible vapour pressure lowering effect, (Ranz and Marshall, 1952), it might be
expected that the mass vaporisation rate in the present example could be calculated by assuming pA,sur = pAsat . However, once a gel has formed, it is perhaps more accurate to view
the moisture as dissolved in the network of branched chains formed by the suspended
silica particles, (Keey, 1975; Bergna, 2006). As such, the increasing solids concentration
will be expected to decrease the equilibrium moisture vapour pressure above the surface, as
described by the associated sorption isotherm, (2.1.1). Rokar and Kmetec (2005) measured the water sorption isotherm for one type of colloidal silica,14 with the data reproduced
in Figure 3.10.15 Least squares regression was used to fit an exponential law to the sorption
data, giving

1
pA,sur = pAsat 1 exp 5A2
,

(3.6.6)

where A is the mass fraction of moisture at the droplet surface. This expression is also
shown in Figure 3.10, along with the isotherm that results from assuming insoluble particles. Comparison shows that, as expected, the measured sorption profile exhibits minimal
vapour pressure reduction prior to gel formation, i.e., whilst A > 0.6.
The effect of the sorption isotherm used is illustrated in Figure 3.11. This shows the
predicted temperature and mass profiles for a droplet of a colloidal silica with an initial
diameter of 0.972 mm drying in air at 101 C, with experimental data from Neic and Vodnik (1991) shown as symbols. The choice of sorption isotherm is seen to have virtually no
effect on the simulated mass profile. In contrast, the effect on the temperature profile is relatively pronounced as the surface solids concentration increases. The figure shows the profile
resulting when the particles are assumed insoluble, i.e., no vapour pressure reduction la14

15

The colloidal silica investigated by Rokar and Kmetec (2005) was Aerosil 200F from Degussa, now Evonik Industries. This is a hydrophilic fused silica with a specific surface area of
200000 m2 kg1 , (Aerosil, 2008)
Rokar and Kmetec (2005) originally presented their data using a dry-mass moisture content. This has
been converted to a wet-mass basis for use in Figure 3.10.

78

CSH

3.6. Applications

1
0.9

Relative Humidity / []

0.8
0.7
0.6
0.5
0.4
0.3

(a) Insoluble Particles


(b) Fit to Sorption Data
(c) Fit to Temperature Profile
Sorption Data

0.2
0.1
0
0

0.2

0.4

0.6

0.8

Water Mass Fraction / []

Figure 3.10: Plot of the three sorption isotherms used in the investigation of colloidal silica droplets drying.
The three isotherms are obtained by: (a) assuming insoluble solids; (b) fitting experimental data shown
in the figure from Rokar and Kmetec (2005); and (c) fitting an experimentally determined temperature
profile for a droplet drying in air at 101 C.

belled (a) and that obtained when using (3.6.6), isotherm (b). Whilst a difference
is discernable, neither predicts the measured values as shell formation is approached and
the temperature rises above the wet-bulb. Indeed, it is not clear that the colloidal silica for
which Rokar and Kmetec (2005) measured their sorption isotherm is a good match for that
used by Neic and Vodnik (1991) in their experiments; Gunko et al. (2006) demonstrate
that whilst different types of fumed silica exhibit qualitatively similar sorption behaviour,
the precise sorption isotherms vary appreciably. Consequently, a new sorption isotherm was
developed by adjusting the constants in the exponential law sorption isotherm to give the
best fit to the experimental temperature data at 101 C. The resulting isotherm is

pA,sur = pAsat 1 exp 5A ,

(3.6.7)

which is the isotherm labelled (c) in Figure 3.10.


The new sorption isotherm continues to predict limited vapour pressure reduction before gelation. Once a gel has formed, the partial pressure of moisture above the drying
surface is lower than that predicted by the previous two isotherms. Figure 3.11 shows that
the temperature profile predicted using this isotherm agrees well with the experimental data
as is to be expected given this is how it was derived. This last sorption isotherm is the
one now used when investigating the influence of drying air relative velocity and humidity.

79

CSH

3.6. Applications

100

80

60

Droplet Mass / mg

Droplet Temperature / C

40

(a) Insoluble Particles


20
(b) Fit to Sorption Data
(c): Fit to Temperature Profile

0
0

20

40

60

Isotherm tshell / s
(a)
(b)
(c)

55.8
56.4
58.3

0
100

80

Time / s

Figure 3.11: Simulated drying of a colloidal silica droplet (lines) compared with experimental results
from Neic and Vodnik (1991) (symbols) at Tgas = 101 C with HR = 0% and vrel = 2.5 ms1 . The
figure illustrates how the sorption isotherm influences the predicted temperature and mass profiles.

100

80

60

Droplet Mass / mg

Droplet Temperature / C

40

20
1

0 ms
0
0

20

1 ms
40

tshell / s

0
1
2
3

> 100

81.1
63.3
54.4

2 ms
60

vrel / ms1

3 ms
80

0
100

Time / s

Figure 3.12: Simulated drying of a colloidal silica droplet (lines) compared with experimental results
from Neic and Vodnik (1991) (symbols) at Tgas = 101 C with HR = 0%. The figure illustrates the effect
of relative air velocity, vrel , on the predicted mass and temperature profiles. The time at which a shell
forms, tshell , is given in the table for each air velocity.

80

CSH

3.6. Applications
Effect of Relative Air Velocity
Figure 3.12 illustrates how the velocity of the air relative to the drying droplet, vrel , influences the predicted temperature and mass profiles. The influence on the mass profile is seen
to be considerably greater than that on the predicted droplet temperature; as expected, air
blowing over the droplet accelerates drying considerably. The relative velocity also strongly
affects the timing of shell formation, tshell . Also plotted in the figure is the experimental data
from Neic and Vodnik (1991) relating to colloidal silica drying in air at 101 C. It is claimed
that the relative air velocity in the experiment was 1.78 ms1 , with no error estimate provided. Whilst such a relative velocity is not completely unrealistic, it is not the most likely
value according to the simulation. Table 3.1 shows the values of the l 2 -norm16 of the deviations between the predicted and experimentally measured values of the droplet mass for
a range of drying air relative velocities. From this it is seen that a velocity of 2.5 ms1 gives
the best agreement between experiment and simulation. Given the likely uncertainties in
both experiment and model, this is not an unreasonable value.
Table 3.1: Values of the l 2 -norm of the deviations, "i , between predicted and measured values of the
droplet mass for different drying air velocities, vrel .

vrel
kk2

0.5

1.0

1.5

2.0

2.5

3.0

3.5

3.95 1.84 1.09 0.633 0.443 0.405 0.478 0.600

Effect of Drying Air Humidity


Figure 3.13 illustrates how the drying air humidity influences the simulated temperature
and mass profiles. The profiles in the figure were obtained using (3.6.7) for the sorption
isotherm and with vrel = 2.5 ms1 . In contrast to the effect of the relative air velocity, it is
seen that the bulk humidity has a significant influence on the temperature profile and has
comparatively little effect on the timing of shell formation. Drying in more moist air results
in a greater droplet temperature a result that can be readily deduced from a psychometric
chart. Neic and Vodnik (1991) do not give the air humidity when drying their droplet of
colloidal silica but, from the Figure 3.13, it is clear that the air must have had a relative
humidity close to zero.

16

The l 2 -norm of a vector, Rn , is the Euclidian norm,


kk2 =

n
X

!1
2

|"i |2

(3.6.8)

i =1

81

CSH

100

80

60

40
RH=0%
RH=1%
RH=2%
RH=3%
RH=4%

0
0

20

40

60

80

20

Droplet Temperature / C

Droplet Mass / mg

3.6. Applications

HR / %

tshell / s

0
1
2
3
4

58.2
61.3
64.3
67.2
70.1

0
100

Time / s

Figure 3.13: Simulated drying of a colloidal silica droplet (lines) compared with experimental results
from Neic and Vodnik (1991) (symbols) at Tgas = 101 C with vrel = 2.5 ms1 . The figure illustrates the
effect of air relative humidity, HR , on the predicted mass and temperature profiles. The time at which a
shell forms, tshell , is given in the table for each humidity.

Predicted Solids Profiles


Figure 3.14 shows the predicted evolution of the solids volume fraction within the colloidal
silica droplet drying at Tgas = 101 C. The results are obtained from simulations of a droplet
drying in air with HR = 0% and vrel = 2.5 ms1 . Profiles are plotted at 5 s intervals. Initially
the profiles are flat across the droplet due to internal circulation prior to the formation of the
gel phase. After formation of a gel at " = 0.23, the profiles develop a pronounced curvature
as particles build up at the receding interface. In the current model, shell formation is
predicted when the surface solids volume fraction reaches 0.65, which is seen to occur at
t = 58 s.
Figure 3.15 shows a comparison between experimental data and the simulated temperature and mass profiles for a 0.95 mm droplet drying in air at Tgas = 178 C. The data is
taken from Neic and Vodnik (1991) as before. The simulation results presented assume, as
discussed above, that the relative humidity of the drying air is 0%. The reported air velocity
for these data is 1.4 ms1 but, after investigation, a velocity of 2.5 ms1 was again found to
give the best fit to the experimental data.
The results presented in Figure 3.15 demonstrate good agreement with the data, but the
range over which the comparison can be made is limited since shell formation is predicted
after 24.4 s. Figure 3.16 shows the corresponding solids volume fraction profiles, plotted
at 5 s intervals up until the point of shell formation. A higher drying air temperature clearly
82

CSH

3.6. Applications

Solids Volume Fraction []

0.9
0.8
0.7

Shell formed at solids volume fraction = 0.65

0.6
t=58s

0.5
0.4
0.3
0.2
0.1
0
0

Radial Position / mm

x 10

Figure 3.14: Simulated solids volume fraction profiles during the drying of a colloidal silica droplet in air
at Tgas = 101 C. Profiles plotted at 5 s intervals, with shell formation predicted at t = 58 s, (highlighted).

250

Shell formed at t=24s

200

Droplet Temperature / C

Droplet Mass / mg

150

100

50

0
0

10

20

30

40

50

0
60

Time / s

Figure 3.15: Simulated drying of a colloidal silica droplet (lines) compared with experimental results
from Neic and Vodnik (1991) (symbols) at Tgas = 178 C with HR = 0% and v = 2.5 ms1 .

83

CSH

3.6. Applications

Solids Volume Fraction []

0.9
0.8
0.7

Shell formed at solids volume fraction = 0.65

0.6
t=24s

0.5
0.4
0.3
0.2
0.1
0
0

Radial Position / mm

8
4

x 10

Figure 3.16: Simulated solids volume fraction profiles during the drying of a colloidal silica droplet in air
at Tgas = 178 C. Profiles plotted at 5 s intervals, with shell formation predicted at t = 24 s, (highlighted).

results in faster drying and correspondingly earlier shell formation. Further, the droplets
dried more quickly form a surface shell after having lost less mass than those drying in air
at lower temperatures: the droplet where Tgas = 178 C has lost 70% of its initial moisture
content at the point of shell formation, compared with a 79% loss for the droplet where
Tgas = 101 C. Correspondingly, droplets drying at higher temperatures are larger at the
point of shell formation. This is an important observation if shrinkage is assumed to cease
with the appearance of a surface crust.
Conclusions
Simulating droplets of colloidal silica allowed the new droplet drying model to be tested
on a simple application where solid nucleation and growth can be ignored. The model
has been successfully validated against experimental measurements and the sensitivity to
drying conditions investigated. This example is revisited in Section 4.6.2 where, after having
discussed models for use following shell formation, the drying droplets simulated in this
section are modelled beyond tshell .

3.6.2 Drying a Droplet of Aqueous Sodium Sulphate Solution


The second set of results presented relate to the drying of a droplet of sodium sulphate
solution. The droplet simulated has an initial solute content of 14 wt%, which is just below
saturation at the initial droplet temperature of 20 C, (Rosenblatt et al., 1984). Raoults law
84

CSH

3.6. Applications
is used to obtain the moisture vapour pressure above the surface of the drying droplet, i.e.,
pA,sur = pAsat A(c) .

(3.6.9)

Sodium Sulphate Diffusion Coefficient


The effective diffusion coefficient, Deff , is the key physical parameter in the continuous
phase equation, (3.3.16). This can be estimated using correlations, of which perhaps the
best known is that published by Wilke and Chang (1955),17
D(AB) =

1.173 1013 (AWA)0.5 T


0.6
V m,B

(3.6.10)

Here, A is an association factor to account for hydrogen bonding in the solvent, WA is


the relative molecular mass of the solvent, V m,B is the molar volume of the solute at its
normal boiling point and A is the solvent viscosity. Applied to the sodium sulphatewater
system, it is appropriate to take A = 2.26 for water and estimate V m, B = 0.0667 m3 kmol1
for sodium sulphate.18 Using these values gives, at 25 C, an estimate of D(AB) = 1.1
109 m2 s1 . Without experimental data, using such a correlation is the only way to obtain
the binary diffusion coefficient. However, there are severe limitations to the values obtained;
for example, the Wilke-Chang correlation is only valid for dilute solutions and, even then,
is typically only accurate to within 10%, (Perry and Green, 1997). It is therefore always
preferable to obtain the diffusion coefficient from experimental data, where this is available.
Rard and Miller (1979) report the binary diffusion coefficient for sodium sulphate in
water at 25 C as a function of solute concentration. Data is reported for concentrations
from zero to saturation. After converting the measured molarities to mass fractions, the
measured data is shown in Figure 3.17. It is interesting to note that the diffusion coefficient
measured in the limit of an infinitely dilute solution agrees very well better than could
be reasonably expected with the value obtained from the Wilke-Chang correlation.
1

Least-squares minimisation was used to fit a second-order polynomial in B2 to the data,

17
18

The form of the correlation given here has been converted for use with SI units.
The molar volume of sodium sulphate is estimated using the structural contributions of each atom, taken
from Sinnott et al. (1999):
Element

V m m3 kmol1

Na
S
O (in union with S)

0.00393
0.0256
0.0083

85

CSH

3.6. Applications
9

1.2

x 10

Diffusion Coefficient / m s

2 1

0.8

0.6

0.4

0.2

0
0

0.2

0.4

0.6

0.8

Na SO Mass Fraction / []
2

Figure 3.17: Plot showing the binary diffusion coefficient, D(AB) , of Na2 SO4 in water at 25 C. Symbols
show experimental data from Rard and Miller (1979), covering the range from infinitely dilute through
to saturation; the line shows a polynomial fit to the data, (3.6.11), which is extrapolated to supersaturated
values.

giving

1
2

D(AB) = 0.4444B 1.4497B + 1.1475 109 m2 s1 .

(3.6.11)

This fitted equation is also plotted in Figure 3.17, where it has been extrapolated beyond
the range of the experimental data to cover supersaturated solute mass fractions. Whilst
extrapolation of this kind is dangerous, the general monotonic-decreasing form of the resulting curve agrees with that reported by Horvath (1985) for the system. The experimental
data and associated fitted curve are only applicable at 25 C; the binary diffusion coefficient
at other temperatures may be obtained using
D(AB) (T ) = D(AB) (T = 298K)

T
298

(3.6.12)

To obtain the effective diffusion coefficient, Deff , for use in (3.3.16), it is necessary to
consider the effect of the suspended solids on the diffusion rate. As before, experimental
data is preferable when available but, in its absence, the influence of suspended solids may
be modelled (Silva et al., 2000). Assuming randomly distributed spherical solid particles,
the effective diffusion coefficient through the medium is given by the Clausius-Mossotti

86

CSH

3.6. Applications
equation,19

3D p D s D p "
Deff = D p +

.
D s + 2D p D s D p "

(3.6.13)

Here D p is the solute diffusion coefficient through the pore space, D s is that through the
solids and " is the solids volume fraction. In the present model, it is assumed that the solute
cannot pass through the solids, i.e., D s = 0, giving
Deff =

2 (1 ")
(2 + ")

(3.6.14)

D(AB) .

For the sodium sulphate example, the final functional form for the effective diffusion coefficient is therefore

2 (1 ") T
1
Deff B , ", T =
0.4444B 1.4497B2 + 1.1475 109 . (3.6.15)
(2 + ") 298

Crystallisation Kinetics
The phase diagram for the sodium sulphatewater system, Figure 3.18, shows that an aqueous sodium sulphate solution will crystallise directly to solid sodium sulphate at temperatures above 30 C without forming a hydrated phase. Since the drying droplets in this example rapidly achieve a wet bulb temperature in excess of 30 C, the kinetics used are those
for direct crystallisation to the solid. For such conditions, Rosenblatt et al. (1984) gives the
equation
dL
dt


= Gmax exp

EA
RT

Ci Ceqm

1.5

(3.6.16)

where EA = 57.4 kJ mol1 is the activation energy for the growth process, Gmax = 1.484
103 ms1 is the maximum growth rate20 and R g is the gas constant. Ci and Ceqm are the
local and saturated solute concentrations respectively in kmol m3 . These are related to the
solute mass fraction by
Ci =

(c)
WB

B (c) ,

(3.6.17)

where (c) is the continuous phase density, (3.3.5), and WB = 142 kg kmol1 is the
relative molecular mass of sodium sulphate. The dependence of the continuous phase den19
20

See Barker (1973) for a derivation of this equation.


Note that (3.6.16), the expression for the linear growth rate from Rosenblatt et al. (1984), is an empirical
equation and, consequently Gmax is best thought of as a fitted coefficient with no physical meaning.

87

CSH

3.6. Applications

50

Temperature / C

40

Na2SO4
+
SOLUTION

SOLUTION

30

20
Na2SO4
+
Na2SO4-10H2O

Na2SO4-10H2O
+
SOLUTION
10

10

20

30

40

50

60

Weight Percent Na2SO4

Figure 3.18: Water-Sodium Sulphate phase diagram adapted from Wetmore and LeRoy (1951).

sity on the continuous phase composition, as given by (3.3.23), must be considered when
evaluating (3.6.16). In terms of the density ratio, B , solute mass fraction, B (c) and
equilibrium solute mass fraction, B (c)
, the driving force for crystallisation is
eqm


0A2B B (c) B (c)
eqm
Ci Ceqm =
.


(c)
(c)
WB B + B B eqm + B

(3.6.18)

The equilibrium solute mass fraction, B (c)


, is obtained from the phase diagram shown
eqm
in Figure 3.18. The crystal growth expression given by (3.6.16) corresponds to the linear
growth rate, G/ms1 , in the population balance equation (3.4.2).
The saturation ratio, S , is defined as the ratio of the solute molar concentration and the
equilibrium solubility, i.e.,
S=

Ci
Ceqm

(3.6.19)

Substituting from (3.6.17), and using (3.3.23) to write the continuous phase densities in

88

CSH

3.6. Applications
terms of the density ratio and mass fractions, the saturation ratio may be expressed
B

S=

1 +

(c)
B eqm

1 + B(c)

(3.6.20)

For the sodium sulphate system, the saturation ratio is less than five throughout drying, so
heterogeneous nucleation dominates, (Dirksen and Ring, 1991). The following standard
functional form was used for this heterogeneous nucleation rate,
log10

N0
N

max

!
=

log10 2 (S)

(3.6.21)

where Nmax is a constant representing the maximum nucleation rate. No nucleation kinetics
for the aqueous sodium sulphate system could be found in the literature and values for Nmax
were found to vary considerably between similar systems, (Dirksen and Ring, 1991). The
sensitivity of the results to the nucleation rate and initial seeding conditions are investigated
below. However, it is worth explicitly noting that the lack of experimental data relating
to crystallisation kinetics is the key factor preventing accurate simulation of the sodium
sulphate system.
Effect of Initial Seeding
When solving the droplet drying model, it is necessary to provide initial conditions for all
variables. Assuming a well mixed system implies the initial spatial profiles are uniform across
the droplet. The initial concentration of sodium sulphate is 14 wt%, but the appropriate
initial conditions for the moment system are less clear. Since the solute is at the saturation
concentration, it seems reasonable to assume that there are some crystallites initially present
in droplet, although their number is unknown. This is unfortunate since seeding that
is, the initial number of crystallites is known to have a large effect on the subsequent
crystallisation process, (e.g., Dirksen and Ring, 1991; van Drunen et al., 1996). This effect
can be investigated using the present model.
Figure 3.19 demonstrates the effect of the initial particle number density on the timing
of shell formation. Recall that the particle number density is given by the zeroth moment of
the particle size distribution. Shell formation as plotted in this figure is taken as the point at
which the solids volume fraction at the droplet surface reaches 0.65. As expected, a greater
number of seed particles results in earlier shell formation.
Whilst the initial particle number density gives the initial condition of the zeroth moment of the number density function, m0(0) , it remains necessary to specify the initial conditions for the three remaining moments. The droplet drying model requires that the solid

89

CSH

3.6. Applications

60
L0=2 nm
55

L0=5 nm
L =10 nm

50

tshell / s

45
40
35
30
25
20 5
10

10

15

10

10

20

10

Initial Particle Number Density / # m3

Figure 3.19: The effect of initial particle number density and initial seed particle size on the timing of
shell formation, tshell . The maximum nucleation rate, Nmax , is set equal to 1010 m3 s1 .

particles are assumed spherical. If it is further assumed that the initial crystallites are the
same size then the second and third moments may be calculated once the particle diameter
or the first moment of the particle size distribution has been specified.
It seems sensible to take the initial particle diameter, L0 , equal to the minimum stable crystal size which, for sodium sulphate, is around 10 nm, (Shi and Rousseau, 2001).
However, as seen from Figure 3.19, the choice of initial particle size does not significantly
affect the timing of shell formation. This result is due to the fact that the particle growth
rate, (3.6.16) is assumed independent of particle size. Since all sizes of seed crystal investigated are small compared with the particle sizes at shell formation, the choice of L0 has
minimal effect on tshell .
Effect of Nucleation Rate
No information relating to the nucleation rate for aqueous sodium sulphate could be found
in the literature. Assuming that the dependence of the nucleation rate on the saturation
ratio follows the standard expression in (3.6.21), the new droplet drying model was used to
study the effect of varying Nmax . The results are plotted in Figure 3.20, wherein the timing
of shell formation is shown as a function of initial particle number density for four different
maximum nucleation rates.
The curve relating to the lowest maximum nucleation rate, Nmax = 1010 m3 s1 , reproduces the results already shown in Figure 3.19. From the present plot it is clear that

90

CSH

3.6. Applications

60
10

m s

20

m s

30

m s

40

m s

=10

=10

=10

max

55

max

50

max

tshell / s

Nmax=10

3 1
3 1
3 1
3 1

45
40
35
30
25 4
10

10

10

10

10

12

10

14

10

Initial Particle Number Density / # m3

Figure 3.20: The effect of the nucleation rate, Nmax , on the timing of shell formation, tshell when drying
a droplet initially seeded with 10 nm crystals.

new particle nucleation has little effect on the timing of shell formation in this case. The
situation is quite different when the maximum nucleation rate is increased, as seen by the
form of the corresponding curves in Figure 3.20. Below a certain limit, the timing of shell
formation is seen to become virtually independent of the initial number density. In this
regime, the formation of the shell is dominated by the nucleation of new particles which
fill space and trigger the condition that " > "crit at the droplet surface. The dominance
of the nucleation process renders the initial condition with respect to the particle number
density inconsequential. In contrast, outside this regime, the formation of a surface shell is
dominated by the growth of the seed particles and, consequently, the timing of tshell remains
dependent on m0(0) in this region. The size of the seed particles is found to have no influence
in either regime.
The two regimes outlined above are expected to influence the mean size of the particles
constituting the newly formed surface shell. This is confirmed in Figure 3.21 wherein the
average particle size in the shell, L is plotted against maximum nucleation rate, Nmax for
three different initial particle number densities, m0(0) . Mean particle sizes are obtained from
the first two moments of the particle number density function using (3.4.8). The ability to
predict the average particle size in the shell along with an idea of the width of the particle
size distribution using the second moment is a key feature of the new model.
Looking first at the points corresponding to the highest initial particle number density,
it is seen that the predicted mean particle size in the shell is around 3.5 m and largely
independent of nucleation rate. This corresponds to the second regime discussed above
91

CSH

3.6. Applications

Average Particle Size / m

10

10

10

m(0)=3.39 108 m3
0

m(0)=3.39 1010 m3
0
m(0)=3.39
0

10

10

10

12

10

20

30

10

10

40

10

Maximum Nucleation Rate / m3s1

Figure 3.21: Plot showing the average solid particle size, L, at r = R at the point of shell formation.
The dependence on the maximum nucleation rate is illustrated for three different initial particle number
densities.

where shell formation occurs as a result of seed growth. Since new particle nucleation
plays a limited role in this regime the particle sizes in the shells that result would not be
expected to depend on Nmax . In contrast, a dependence on the maximum nucleation rate
is observed for lower initial particle number densities. Higher values of Nmax result in more
particles being nucleated and, consequently, smaller mean particle sizes are observed. For
low maximum nucleation rates the formation of a shell is again dominated by seed growth
and the mean size of particles in the shell is once more independent of Nmax .
Figure 3.22 shows the influence of the maximum nucleation rate, Nmax on the timing
of shell formation for a droplet with initial particle number density, m0(0) = 3.39 108 m3 .
As expected, tshell is seen to be constant for low values of Nmax , where nucleation plays
almost no role. Interestingly, the initial effect of increasing the maximum nucleation rate
is to delay the onset of shell formation. This effect can be explained by understanding
that nucleation creates small particles at the expense of existing particle growth. These
small particles are able to diffuse away from the droplet interface far more quickly than the,
relatively immobile, large particles. Consequently, the effect of nucleating new particles at
the expense of growing existing ones is to delay the onset of shell formation. At still higher
maximum nucleation rates, the production of new particles is so fast that this initial effect
is negated and a shell is formed increasingly quickly.
Figure 3.23 plots the average particle size within a drying droplet as a function of radial
position and time. The profiles shown relate to a droplet where the initial particle number
92

CSH

3.6. Applications

48
46

tshell / s

44
42
40
38
36
34 10
10

20

30

10

10

40

10

Maximum Nucleation Rate / m3s1

Figure 3.22: Plot showing how the maximum solid particle nucleation rate effects the timing of shell
(0)
formation for a droplet with initial m0 = 3.39 108 m3 s1 .

density is 3.39 108 m3 throughout the droplet and the maximum nucleation rate is
1020 m3 s1 . The timing of shell formation for such a droplet is expected to be dominated
by new particle nucleation.
Solid particle growth adjacent to the receding droplet interface is easily identified from
the figure. New particle nucleation acts to lower the mean particle size, since all newly nucleated particles have the minimum size, Lmin . Close inspection of the profiles in Figure 3.23
shows that this behaviour is observed at all radial locations following the onset of nucleation
around t = 42 s. The mean particle size at the very edge of the droplet is noticeably greater
that that even a short way in and consequently, following the onset of nucleation, a local
minimum develops in the radial profiles. The larger mean size of particles at the edge of the
droplet is understandable this is where the solute concentration is greatest, resulting in
the highest rates of growth. Further, larger solids diffuse into the bulk more slowly, with the
effect that the size distribution at the boundary becomes skewed towards bigger particles.
Solids Diffusion Coefficient and Moment Closure
From the figures and discussion above it is clear that crystal nucleation and growth mean
that the solid particles in the sodium sulphate system vary greatly in size. This is in contrast
to mono-sized particles in the droplet of colloidal silica studied in Section 3.6.1 and, for
this reason, a size dependent solids diffusion coefficient is necessary here.
The modifications required to the moment system to handle a size dependent diffusion

93

CSH

3.6. Applications

10

Particle Size / m

10

10

10

10

15
20
25
30
0

Time / s

35

1
2

40

3
4

45

5
50

6
7

Radial Location / m
4

x 10

Figure 3.23: Plot showing spatially resolved profiles of the mean solid particle size, L, at 2 s intervals
from t = 20 s until the point of shell formation, tshell . The profile at tshell is shown in bold.

coefficient have been discussed in Section 3.4.3. In particular, the closure used is that given
by (3.4.24), based on simple linear extrapolation of the moments when plotted on a logarithmic scale. Frenklach (2002) showed such a closure to be suitable in a number of similar
applications, but a full numerical study would be required to establish its applicability here
beyond doubt. However, preliminary investigations provide encouragement. For example,
the closure approximation may be applied to the first and second moments to approximate
the zeroth moment. Since the zeroth moment is explicitly calculated by the model, the
approximated values using the closure can be compared with the simulated values. Such a
comparison is shown in Figure 3.24 for two different initial particle number densities, m0(0) .
The moment values plotted are those at the outer edge of the receding drying droplet. The
closure is initially very good indeed, it should be exact for a mono-sized distribution of
particle sizes and remains fair up until the point of shell formation.
Experimental Comparison
The validity of any experimental comparison is clearly limited by the uncertainty associated
with the crystallisation kinetics. This uncertainly most obviously manifests itself around the
time of shell formation. Further, as has been demonstrated above, the timing of shell formation is itself strongly dependent on the initial seeding and subsequent nucleation kinetics.

94

CSH

10

Zeroth Moment
First Moment
Second Moment
Approximated Zeroth Moment

Normalised Moments / []

Normalised Moments / []

3.6. Applications

10

10

10

10

Zeroth Moment
First Moment
Second Moment
Approximated Zeroth Moment

10

10

10

20

30

40

10

50

10

20

30

Time / s

Time / s

(a)

(b)

40

50

100

80

Droplet Mass / mg

Droplet Temperature / C

Figure 3.24: Comparison of the simulated zeroth moment at the outer edge of a drying droplet, (thick
line) with that obtained from the first and second moments using a closure approximation of the form
given by (3.4.24), (symbols). The maximum nucleation rate is Nmax = 1020 m3 s1 and the initial
(0)
(0)
particle number density is: (a) m0 = 3.39 105 m3 s1 ; and (b) m0 = 3.39 108 m3 s1 .

Shell formed at t=47s

60

40

20

0
0

20

40

60

80

0
100

Time / s

Figure 3.25: Simulated drying of a 14 wt% sodium sulphate in water droplet at Tgas = 90 C, compared
with experimental results from Neic and Vodnik (1991). For this simulation, the maximum nucleation
(0)
rate was set to Nmax = 1020 m3 s1 and the initial particle number density was m0 = 3.39 108 m3 .

Nevertheless, it is instructive to compare predictions with experimentally determined temperature and mass profiles, if only to validate the model in the initial drying period where
the solids remain relatively unimportant. To this end, comparisons with experimental measurements from Neic and Vodnik (1991) of a sodium sulphate droplet drying in air at 90 C
are presented in Figure 3.25.
The size of the model droplet has been fixed such that the initial mass corresponds

95

CSH

3.6. Applications

Solute Mass Fraction / []

0.9
t=47 s

0.8
0.7
0.6
0.5
0.4

t=18 s

0.3
0.2
0.1
0
0

Radial Position / m

8
4

x 10

Figure 3.26: Simulated solute mass fraction profiles plotted at 5 s intervals throughout the drying of a
droplet of aqueous sodium sulphate solution. In addition, profiles are plotted at the point of shell formation,
(dashed line) and solution at the edge of the droplet first becomes supersaturated, (bold line).

with the experimental observations, thus giving a diameter of 1.78 mm. For this and all
subsequent figures in this section, the simulation was run with Nmax = 1020 m3 s1 and
with an initial solid particle number density of m0(0) = 3.39108 m3 . Taking the trigger for
shell formation to be a surface solids volume fraction of 0.65, it is seen that this choice of
parameters corresponds to the appearance of a rigid crust after 47 s. The match between the
simulated and measured results up until this point is seen to be excellent, demonstrating the
validity of using Raoults law to model the solvent partial pressure above the droplet surface.
Solute Profiles and Moment Evolution
Figure 3.26 plots the solute mass fraction profiles at 5 s intervals through the drying of
the droplet. The dashed curve is the solute profile at the point of shell formation and the
lower bold curve indicates the first profile when the solution becomes supersaturated. The
saturated mass fraction is obtained from the phase diagram, Figure 3.18, where it is seen
that, above 30 C, this is only a weak function of temperature and approximately equal to
33 wt%. Naturally, the first appearance of a supersaturated solution occurs at the outer
boundary of the drying droplet.
The evolution of the moments at the moving external boundary of the droplet are plotted in Figure 3.27. In this figure, the three moments are all normalised against their initial
values. Initially all the normalised moments are coincident, indicating that no growth is taking place. The moment values increase in this period simply because the particle number
96

CSH

Normalised Moments / []

3.6. Applications

Zeroth Moment
First Moment
Second Moment
10

10

10

10

10

20

30

40

50

Time / s

Figure 3.27: Simulated normalised moments at the outer edge of the droplet.

density at the droplet surface rises as a result of shrinkage. Figure 3.26 shows that the continuous phase at the surface of the droplet becomes saturated with the solute at around 18 s.
This corresponds to the point at which growth starts to be observed in the moment system,
as would be expected from inspecting (3.6.16), the expression for linear crystal growth.
The zeroth moment, representing the particle number density, is seen to increase rapidly
after about 42 s. This point marks the time at which new particle nucleation becomes
appreciable and, as has been discussed at length above, is determined by the choice of Nmax .
The decreasing first and second moments reflect the fact that the newly nucleated particles
are smaller than those particles which have grown from the seeds present at the start of
drying.
Figure 3.27 may be contrasted with Figure 3.28 which again shows the normalised
moments, but now integrated over the entire droplet. This figure more clearly indicates the
onset of new particle nucleation around t = 42 s. The change in the rate of increase of
the zeroth moment around t = 18 s in Figure 3.27 is a consequence of the size dependent
diffusion coefficient. As the particles at the droplet edge grow, their mobility decreases
markedly. This reduces the rate at which they diffuse into the droplet bulk and therefore
increases the rate of accumulation at the droplet surface. In contrast, the zeroth moment
integrated over the entire droplet, as shown in Figure 3.28, shows no change in its second
derivative prior to the onset of nucleation. This confirms that the change observed at the
droplet edge is due to a local change in particle mobility rather than a droplet-wide effect.
Finally, Figure 3.29 shows the predicted evolution of the solids volume fraction in the

97

CSH

3.6. Applications

Integrated Normalised Moments / []

10

Zeroth Moment
First Moment
Second Moment
6

10

10

10

10

10

20

30

40

50

Time / s

Figure 3.28: Simulated normalised moments integrated over the entire droplet.

drying droplet. The volume fraction of solids is negligible until just before shell formation.
The relative immobility of the particles forming the shell means that the profiles representing the solids volume fraction are extremely steep. This suggests a narrow region in which
crystals form, surrounding a more dilute core. Qualitative experimental evidence supports
this picture, (Seydel et al., 2006), but no quantitative data could be found in the literature. The shell forms rapidly at a time that, as shown above, is highly dependent on the
nucleation kinetics employed. Good knowledge of these kinetics is required for accurate
prediction of the time to shell formation, however such data is currently unavailable.
Conclusions
The simulation of a saturated droplet of aqueous sodium sulphate demonstrates the full
capabilities of the new model. Drying rates are successfully calculated, as validated against
experimental data, and spatially resolved moisture and dissolved solute profiles are returned.
In addition, the equations describing the evolving moments of the solid particle population
are solved using a size dependent diffusion coefficient, enabling prediction of particle number densities, mean particle sizes and the solids volume fraction as a function of radial
position within the droplet.
The work presented here has demonstrated the number of physical and kinetic parameters required to successfully simulate this drying system. Some of these parameters such
as the sodium sulphate diffusion coefficient and crystal growth rate are readily available,
whereas for other required information such as the particle nucleation rate no rele-

98

CSH

3.7. Conclusions of the Chapter

0.8

Solids Volume Fraction / []

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

Radial Position / m

8
4

x 10

Figure 3.29: Simulated solids volume fraction in a drying droplet of aqueous sodium sulphate solution.

vant data could be found in the literature. This lack of key experimental data hampers the
ability of the model to make quantitative predictions. However, the new model may still be
used to investigate likely sensitivities to these unknown parameters; such studies have been
conducted to investigate the influence of the maximum particle nucleation rate and initial
particle number density.
The population balance describing the solid phase enables structural quantities of interest to be predicted, for example, the average size of the solid particles forming the initial
surface shell. As discussed in the next chapter, this type of information is essential if droplet
drying and morphological development following shell formation are to be simulated successfully.

3.7 Conclusions of the Chapter


The aim of this chapter was to develop a new droplet drying model capable of simulating droplets containing both dissolved and suspended solids. The continuous phase was
modelled as an ideal binary solution and, after reviewing various approaches for describing
transport in the presence of solids, volume averaging was identified as the most promising
method for deriving appropriate transport equations. The suspended solids were described
by means of a population balance equation, allowing for particle nucleation, growth and
transport in physical space. Solving for the moments of this equation allowed properties
of interest relating to the solids such as particle number density, mean particle size and

99

CSH

3.7. Conclusions of the Chapter


solids volume fraction to be spatially resolved. Equations for the discrete and continuous
phases were derived, along with a detailed discussion of the appropriate boundary conditions. Taken together, this represents a comprehensive framework for describing drying
droplets.
The later sections of this chapter dealt with model validation. Simulations of a colloidal
silica droplet reproduced experimentally observed mass and temperature profiles. The ability of the new model to investigate sensitivity to drying conditions was also demonstrated.
The more complicated aqueous sodium sulphate system was then investigated, illustrating
the full capabilities of the model. The influence of particle nucleation rate and initial seeding on the timing of shell formation and mean solid particle size was investigated, producing
several interesting qualitative results: for example, increasing the maximum nucleation rate
was initially found to delay shell formation. Experimental data relating to droplet mass and
temperature histories were reproduced, although uncertainty associated with key kinetic
parameters prevented further quantitative validation.
In summary, this chapter contained the following novel elements:
a framework to describe drying droplets based on volume averaged transport equations and a population balance describing suspended solids was developed;
the core droplet drying model was used to simulate systems of practical interest and
validated against experimental data.

100

CSH

Chapter 4
Drying Droplets With an Outer Shell
In which the novel droplet drying model introduced in Chapter 3 is extended to
droplets drying in the presence of a surface shell and a centrally located bubble.
Following the initial appearance of an outer shell, a number of sub-models may be
used to described subsequent drying and structural development. These sub-models
include a novel shell thickening description, as well as the wet and dry shell
models from the literature. Physically motivated criteria are developed that allow
the appropriate sub-model to be chosen at each stage in the droplet drying history.
Comparisons between model predictions and experimental data from the literature
are presented for two systems of interest: detergent crutcher mix and colloidal silica.
These simulations demonstrate the extended model is capable of simulating multiple
dried-particle morphologies together with other properties of interest such as moisture
profiles and moments of the particle size distribution. The material presented in this
chapter includes work published in Handscomb et al. (2008b).

4.1 Background
Spray drying removes moisture and simultaneously produces particulate products. Chapter 3 introduced the core features of a new model capable of simulating the removal of
moisture from droplets. However, the discussion in that chapter was limited to droplet drying prior to shell formation; the simulations in Section 3.6 were explicitly terminated on
the appearance of an external crust. Whilst such a model provides some important information, a drying model incapable of simulating dried-particle formation is clearly limited.
The extended model developed in this chapter allows prediction of moisture removal rates
following shell formation but, equally importantly, simulates the evolving morphology of
the drying droplet.

101

4.1. Background

4.1.1 Droplet Morphology


The morphologies observed when droplets are spray dried were discussed at length in Section 2.4.2, with the major particle types illustrated in Figure 2.9. In principle, a drying
model could be formulated to simulate a droplet drying to any single one of these morphologies. Of more interest would be a model which could simulate the production of
multiple dried-particle structures and, crucially, be capable of deciding which of these structures will result when a given initial droplet is subjected to different drying histories. The
model developed in this chapter seeks to do this, making use of the information about the
solid phase given by the moments of the population balance (3.4) to simulate different
structural developments. However, before proceeding to develop such a model, it is necessary to narrow the field of potential dried-particle morphologies.
A core assumption of the model developed in the previous chapter was spherical symmetry of the drying droplets. This immediately limits the range of dried-particle morphologies
capable of being simulated by the extended model. Three possibilities remain: solid particles; particles with a centrally located void; and inflated shells. All three of these particle
types are observed at the outlet of real spray drying installations, (Marshall and Seltzer,
1950a). Processes that are symmetry breaking notably particle bursting or collapse
cannot be explicitly simulated within such a framework. However, if the conditions leading
to these processes are identified, the new model might be used to flag when the associated
dried-particle morphologies are likely to be observed.
The formation of any dried-particle commences with the appearance of a surface shell.
The present discussion proceeds by considering what is meant by this term and discussing
the physical processes associated with the appearance of a shell.

4.1.2 Formation of a Surface Shell


There is no unique meaning of the term surface shell. Rather, the definition depends on
the context, i.e., the particular model or experiment. Whilst a structural feature associated
with the term might be clearly identifiable in the later stages of drying, the precise point at
which such a shell forms is not obvious, leading to the various definitions.
First consider the processes preceding shell formation however that may be defined.
Depending on the droplet composition, suspended solids may be present initially or may
crystallise out of solution as moisture is evaporated. If present initially, the solids are assumed uniformly distributed within the droplet as illustrated in Figure 4.1a. As drying
proceeds, the solid particles are drawn inwards by surface tension forces, causing the solids
volume fraction at the receding surface to increase. Where the solids are initially in solution, crystallisation will first occur in the regions of highest solute supersaturation, i.e., at
the droplet surface, (Farber et al., 2003). Once formed, these crystals are drawn inwards

102

CSH

4.1. Background
behind the receding interface as described previously.
The speed with which the surface solids concentration increases depends on the relative
rates of moisture evaporation and internal solids mixing. That is, the ratio between the
interface recession rate and the rate at which solids can diffuse back into the bulk. Tsapis
et al. (2005) introduced characteristic mixing and drying times to illustrate this idea; the
characteristic mixing time in a droplet of radius R is related to the solids diffusion coefficient
by
mix

R2

(4.1.1)

so that a shell forms when mix  dry . One way to identify the point at which a surface
shell forms is to specify some associated critical solids concentration at the interface. This
is a common approach in droplet drying models, (e.g., Elperin and Krasovitov, 1995; Kadja
and Bergeles, 2003; Seydel et al., 2006), and is indirectly used in those experiments which
take a visual change as the criterion for shell formation. For example, Walton and Mumford
(1999a) noted changes in droplet opacity and Eslamian and Ashgriz (2007) used changes
in the amount of reflected light from droplets viewed under an optical microscope.
In the new droplet drying model, a shell is deemed to have formed once the volume fraction of particles at the droplet surface rises above a certain critical value, "crit , (Figure 4.1b).
A number of researchers (e.g., Brouwers, 2006; Masuda et al., 2006) have discussed how the
solid particle size distribution may be used to predict the maximum packing fraction and it
is imagined that the information from the population balance could be used to inform the
point at which shell formation is deemed to occur. However, for the systems simulated in
this thesis, "crit is taken to be a pre-defined parameter.
Using a critical surface solids concentration to identify the timing of shell formation
says nothing about the subsequent drying behaviour, nor about the nature of the growing
shell. Drying models make numerous assumptions on these two issues, but it is important
to remember that these are features of the model rather than a reflection of the physics.
For example, most models in the literature assume the droplet ceases shrinking as soon as
a shell is formed, (e.g., Mezhericher et al., 2007; Dalmaz et al., 2007; Kadja and Bergeles,
2003), but this need not be the case. Of the three morphological categories identified by
Walton and Mumford (1999a), shell formation arrested shrinkage only for droplets forming dried-particles with a crystalline structure. Skin-forming droplets and those forming
dried-particles with an agglomerated structure both continued to decrease in size after the
appearance of a surface shell. It is this observation which motivates the development of
the shell thickening model, introduced in Section 4.3. This thickening regime allows for
continued shrinkage of the droplet whilst the shell grows to a critical shell thickness, Tcrit .
At this point, the shell is is capable of supporting itself and prevents the droplet shrinking

103

CSH

4.1. Background

"

"

"

"crit

"crit

"crit

R0

(a)

Tcrit

Tmin

r
S R R0

SR

(b)

(c)

Figure 4.1: Schematic illustrating the process of shell thickening: (a) initially the droplet is homogenous
and drying proceeds ideally; (b) a shell forms when the maximum packing fraction, "crit , is reached at the
surface of the droplet; (c) the solid particles continue to redistribute and the droplet continues to shrink
until a critical shell thickness, Tcrit , is attained.

further, (Figure 4.1c).

4.1.3 Drying With a Rigid Shell


Most drying droplets stop shrinking at some point.21 Some researchers define this to be the
point of shell formation, but in the current work the stabilisation of the droplet radius is
associated with the formation of a rigid shell. Such a rigid shell fixes the volume of the drying droplet; further moisture removal must therefore lead to an expanding vapour-saturated
space somewhere within the droplet. This vapour space may be dispersed throughout the
droplet, in which case detailed methods for simulating moisture transport through porous
media containing a gas phase, as discussed in Section 3.1.1, are required. However, for the
reasons explained in that section, such detailed simulation is not attempted here. Rather,
following Lee and Law (1991), two limiting cases are identified upon which shell drying
models are developed.
In the dry shell model, the vapour space is located in the shell itself and an evaporative
front recedes through the droplet. This is a variant of the classic shrinking core approach
used by many researchers to model droplets drying and discussed in detail in Section 2.4.4.
This approach is a fair approximation for several drying systems and an implementation for
21

Exceptions to this statement are droplets with very low initial solid (suspended or dissolved) concentrations, or those droplets which inflate before shrinkage is arrested.

104

CSH

4.2. Deciding Which Model to Use


use with the new model developed in this thesis is introduced in Section 4.4. However, a
number of problems are evident. Firstly, most implementations assume that the porosity
and other properties of the shell are uniform, (e.g.,Cheong et al., 1986; Dalmaz et al., 2007).
This is not necessarily the case, especially if solute is crystallising from the continuous phase
within the still-wet core. The dry shell model described in Section 4.4 overcomes this
objection by allowing the shell properties to vary. A more difficult objection to overcome
is that, on its own, the dry shell model can only predict solid dried-particles. Furthermore,
these dried-particles will have a solids volume fraction similar to that of the droplet when it
stops shrinking. The simulated dried-particle densities are, consequently, often too low.
A further challenge to the dry shell model comes from Walton and Mumford (1999a),
who observed that droplets which form agglomerated dried-particles that is, droplets
containing suspended rather than dissolved solids exhibit saturated surface drying throughout the shrinking period and for a short time thereafter. Similar behaviour is found with
skin forming droplets, which remain wetted at the surface during skin thickening even
though the surface film can present a significant obstacle to mass transfer. Such behaviour
is in stark contrast to that predicted by the dry shell model, based as it is upon the idea of
evaporation occurring at an interface receding into the droplet bulk. These problems are
addressed by the wet shell model.
The wet shell model assumes that the shell region is, at all times, wetted by the continuous phase; the vapour now lies in a single, centrally-located, bubble. Evaporation proceeds
from the wetted external surface, as observed in those experiments discussed above. Furthermore, the model predicts the formation of hollow dried-particles and, with suitable
modifications, can also be used to simulate inflating droplets. The development of the wet
shell model, along with details of its implementation, is given in Section 4.5. The wet and
dry shell models in combination are capable of simulating the three droplet morphologies
identified at the start of this section. After giving details of each model in isolation, it is
shown how the evolving droplet properties can be used to select the appropriate model to
use. The chapter closes with two applications of the new model showing comparison with
experimental results from the literature.

4.2 Deciding Which Model to Use


Before proceeding to discuss the structure and implementation of the thickening, wet and
dry shell models in more detail, this section considers which of these models should be
used following shell formation. As mentioned above, the dried-particle formation model is
traditionally specified a priori. In contrast, the new droplet drying model developed in this
thesis allows the choice of drying model and consequently the generic morphology of
the final particle to be decided during the course of the simulation. In order to explain

105

CSH

4.2. Deciding Which Model to Use


the manner in which this is achieved, it is first necessary to take a closer look at the physics
of moisture removal following shell formation.

4.2.1 Physics of Drying Following Shell Formation


Pressure Drop Across Surface Menisci
Initially, all the solid particles are completely wetted by the continuous phase as illustrated
in Figure 4.2a. Once a shell has formed, the continuous phase recedes to the pore mouths
and menisci form between the particles at the surface. These menisci support a pressure
drop of
P =

2
rM

(4.2.1)

where is the surface tension of the liquid phase and rM is the radius of curvature of the
meniscus. The pressure gradient within the droplet resulting from this capillary pressure
leads to transport of the continuous phase towards the outer surface. At the same time,
the tension from these menisci drives the solid particles past each other towards the droplet
centre, (Tsapis et al., 2005). These two processes occur simultaneously and provided
the capillary forces are strong enough result in continued droplet shrinkage after the
formation of a surface shell. This continued, capillary driven shrinkage is termed shell
thickening.
During this thickening period the shell remains quite weak and the pore network shrinks
in response to the capillary tension, (Brinker, 2006). As a result, the radius of curvature,
rM , remains large, as shown in Figure 4.2b. In some types of gel, it is possible for the pore
structure to shrink to as little as one tenth of the original volume during the thickening
period, (Scherer, 1990). However, the network will stiffen as it shrinks, increasing the
capillary pressure. At some point, the radii of curvature of the surface menisci will become
so small that they can fit into the pores, as shown in Figure 4.2c. This point corresponds
with the start of the first falling-rate period discussed in Section 2.1.3 and also represents
the point with the maximum capillary pressure in the liquid.
Simple geometry shows that, if the pores are assumed cylindrical, the critical radius of
curvature for a meniscus entering a pore is
rM =

a
cos

(4.2.2)

where a is the pore radius and is the contact angle, (Scherer, 1990). Clearly there will
be some characteristic pore radius, rc , corresponding to the peak capillary pressure and
this will likely be related to the size of the narrowest pores. Several workers have shown

106

CSH

4.2. Deciding Which Model to Use

rM

(a)

(b)

rH

(c)

Figure 4.2: Illustration showing the drying process in the shell thickening period. (a) Prior to shell
formation, the suspended solids are completely wetted by the liquid phase. (b) On shell formation, tension
develops in the liquid as surface menisci form with radii of curvature, rM . Initially, the solid network
yields easily requiring little stress to deform and, as a result, the radii of the menisci are large. (c) As the
porous network stiffens, the stress required for deformation increases and the menisci radii decrease. The
limiting radius corresponding to the maximum capillary pressure occurs when the menisci recede
into the pores.

that this characteristic pore size is closely tracked by the hydraulic radius of the porous
medium, (Scherer, 1994; Smith et al., 1995). If the total surface area per unit volume is
given by SV , the hydraulic radius is
rH =

2 (1 ")
SV

(4.2.3)

Taking a in (4.2.2) as rH and substituting into (4.2.1) gives the maximum pressure drop
across the surface menisci as
Pmax =

cos
(1 ")

(4.2.4)

SV ,

where the need to assume cylindrical pores has been removed. Instead, the present model
allows the volumetric solids surface area to be obtained form the second moment of the
particle number density using (3.4.9), that is
(4.2.5)

SV = m2 .

If the suspended particles are assumed to be mono-disperse with diameter L, then (4.2.3)
becomes
rH =

(1 ")
3"

(4.2.6)

L,

and it is seen that the characteristic pore size is inversely proportional to the solids volume

107

CSH

4.2. Deciding Which Model to Use


fraction.
Equation (4.2.4) shows that it is theoretically possible to develop substantial pressure
drops across surface menisci. As an example, consider the drying of a water based gel where
the solids volume fraction, ", is 0.65 and SV = 2 108 m2 m3 . Water has a surface tension
of 0.072 Nm1 and the contact angle is taken as = 0 rad for convenience. Under such,
fairly typical, circumstances the maximum pressure drop is greater than 40 MPa.
In general, the pressure drops predicted by (4.2.4) cause a negative absolute pressure
within the fluid. This implies that the drying fluid is under tension, (Imre, 2000). Water the fluid of interest in the present thesis is capable of withstanding very large
negative pressures; that is, water has a high tensile strength. Theoretical values as high
as 3000 bar have been derived for the tensile strength of water at room temperature, (Tas
et al., 2003), with negative pressures as high as 1400 bar at 42 C having been observed
experimentally, (Zheng et al., 1991). Cavitation occurs when the negative pressure exceeds
the tensile strength of the fluid, or earlier in the presence of impurities capable of acting
as cavitation nuclei, (Briggs, 1950; Tas et al., 2003). Cavitation is the result of the energy
barrier to bubble nucleation being overcome; the result is the spontaneous appearance of a
bubble within the stretched liquid, (Maris and Balibar, 2000).
Pressure Drop Across a Thickening Shell
During most of the thickening period the capillary pressure is significantly less than the
maximum value given by (4.2.4) and, consequently, the meniscus radius is greater than the
hydraulic radius, rH . It is therefore not possible to use (4.2.1) to directly calculate P in
this regime. Instead, the capillary pressure, Pcap , is obtained by likening the movement of
the continuous phase through the shell to filtration through a porous filter with the same
thickness as the growing crust, (Minoshima et al., 2001). The pressure drop across the shell
is then assumed equal to the pressure drop across the aircontinuous phase interface. This
analogy is illustrated schematically in Figure 4.3.
Given knowledge of the liquid flow rate, Darcys law may be used to obtain the corresponding pressure drop across a porous medium. Darcys Law may be written
K
q = P ,

(4.2.7)

where K is the permeability tensor for the porous body and is the viscosity of the flowing
fluid. Making use of the assumed spherical symmetry in the present problem, this reduces
to
q =

P
r

(4.2.8)

108

CSH

4.2. Deciding Which Model to Use


00
m
0A

00
m
0A

Pe

00
m
0A

Pe

Pe

Tmin
T
r

(a)

(b)

(c)

Figure 4.3: Schematic demonstrating how the process of shell thickening can be likened to the growth of a
filter cake: (a) Prior to shell formation; (b) the point of shell formation corresponds with the appearance of
a filter cake one particle thick; (c) the filter cake thickness, T continues to increase as the shell grows. The
00 , and results
volumetric flow through the filter cake is simply related to the mass flux from the droplet, m
in a pseudo-pressure acting on the shell, Pe .

where q is the volume flux of liquid through the shell. At the external surface, the volume
00 , by
flux is simply related to the mass vaporisation flux, m
q=

00
m
0A

dR
dt

(4.2.9)

Assuming that the problem may be analysed as a pseudo-steady state and, further, that the
total volumetric fluid flow through the shell is not a function of radius, allows (4.2.8) to be
integrated, giving
P = PR PS =
=

vap
m

R
S

4r 2 0A

00 R
m
0A S

dr

(4.2.10)

T ,

where T = R S is the shell thickness. In deriving this expression, it has also been assumed
that the permeability and viscosity are constant across the thickening shell. This assumption
seems fairly reasonable and is used again in the next section to calculate the shell growth
rate.
The permeability, , in (4.2.10) is estimated using the Carmen-Kozeny relation, (Coulson et al., 1996),
=

1 (1 ")3
180

"

L2 .

(4.2.11)

In the above expression, L is the diameter of the solid particles assumed spherical and
monodisperse making up the porous medium. To account for a distribution of particle

109

CSH

4.2. Deciding Which Model to Use


sizes, (4.2.11) may be replaced by
=

1 (1 ")3
5

SV2

(4.2.12)

where SV is the solids surface area per unit volume, discussed above, (Scherer, 1990).
Strength of a Surface Shell
So far, this section has demonstrated a method for calculating the pressure drop that exists
across the surface menisci of a droplet drying in the presence of a surface shell. This pressure
drop is important as it can cause continued shrinkage of the droplet following shell formation. In order to say when the droplet stops shrinking, it is necessary to determine when the
thickening shell becomes structurally capable of supporting itself. The continued deformation of the shell is hypothesised to occur through a series of mini-buckling events driven by
the capillary pressure of the receding continuous phase. Timoshenko (1936) showed that a
spherical shell of radius R and thickness T will buckle when subjected to a uniform external
pressure, Pbuck , given by

Pbuck =

T
RT

2E
3 1 2

(4.2.13)

,

where E is the Youngs modulus of the material and is its Poissons ratio. It is hypothesised
that the newly formed shell will continue to experience buckling events and therefore
continue to thicken so long as the capillary pressure is greater than the buckling pressure,
Pbuck .

4.2.2 Criteria for the Different Models


Having introduced the necessary theory, the criteria for applying the different sub-models
following shell formation are now presented. These criteria are based on the relative magnitudes of the capillary pressure in the surface pores, the strength of the growing shell and
the pressure drop across it quantities which are continuously tracked in the course of
the simulation. The decision process applied to decide the appropriate model to use is
summarised in Figure 4.4; this section explores that decision process in more detail.
Thickening Regime
Upon formation of a surface shell, all drying droplets are assumed to enter the thickening regime. The initial shell is set to have a certain minimum thickness. Whilst this
is an essentially free parameter, the diameter of the smallest solid particle sets a physical

110

CSH

4.2. Deciding Which Model to Use

Pshell > Pbuck

Pbuck < Pmax

Pbuck Pshell
> Pcrit

D S

W S

Figure 4.4: Illustration of the decision process used to select the appropriate drying model to use following
shell formation.

lower bound on possible shell thicknesses. The pressure drop across this shell is calculated using (4.2.10) and this will be equal to the pressure drop across the aircontinuous
phase interface. Further, the pressure required to buckle the shell, Pbuck , is calculated using (4.2.13). As explained above, the capillary pressure in the surface pores can be thought
of as exerting an external pseudo-pressure, Pe , on the thickening shell. The magnitude of
this pseudo-pressure is equal to the capillary pressure at the surface which, in turn, is equal
to the pressure drop across the shell. The shell will continue to thicken so long as this
pseudo-pressure is greater than the pressure required to cause buckling of the shell. That is,
thickening continues whilst
(4.2.14)

Pshell > Pbuck ,

or, on substituting from (4.2.10) and (4.2.13) and re-arranging, whilst


R (R T )
T

00 >
m

2E0A

(4.2.15)

.
2
3 1

The structural properties of the growing shell are seen to directly influence the duration of
00 .
the thickening regime, as is the drying rate, m
The thickening period extends whilst the inequality in (4.2.15) is satisfied. However,
this condition is sufficient but not necessary for thickening, i.e., thickening does not necessarily cease once (4.2.15) is no longer satisfied. The reason for this is that the surface
menisci between suspended solid particles are potentially capable of supporting a greater

111

CSH

4.2. Deciding Which Model to Use

R
S

S
P
0

P
Pshell

0
Pshell

(a)

Pbuck

(b)

Figure 4.5: Illustration of the pressure profiles within a drying droplet following formation of a surface
shell: (a) the pressure drop across the liquidair interface is equal to the pressure drop across the shell,
Pshell when this is greater than the buckling pressure, Pbuck ; (b) when Pbuck > Pshell , the pressure
drop across the interface is equal to the buckling pressure and the pressure in the droplet core falls below
ambient.

pressure drop than that across the thickening shell. Once (4.2.15) is no longer satisfied,
the menisci will retreat into the surface pores, reducing their radius of curvature and thus
increasing the associated capillary pressure. This retreat will continue until the pressure is
sufficient to cause further buckling, or until the capillary pressure exceeds the maximum
possible value as given by (4.2.4).
Thickening will cease once the capillary pressure exceeds the maximum possible value,
Pmax . At this point, the menisci retreat beneath the surface and the dry shell regime
commences. However, it is possible for the thickening regime to end earlier by progression
to the wet shell regime.
Wet Shell Regime
When drying droplets with a surface shell, the surface menisci support a pressure drop
across the air-continuous phase interface. As explained above, it is initially assumed that
this pressure drop is equal in magnitude to the pressure drop across the thickening shell
and, consequently, the pressure profile in the drying droplet is similar to that sketched in
Figure 4.5a. Once the menisci begin retreating into the surface pores, the pressure drop
across the menisci grows larger than that across the shell. As a result, the fluid pressure in
the core region falls below atmospheric, as illustrated in Figure 4.5b.
The wet shell model, as described above, postulates the growth of a centrally located
vapour-saturated bubble within the drying droplet. This approach is clearly a gross simplification of the physics occurring within a real drying droplet, but it is possible to motivate
the idea of a centrally located bubble in a number of ways. Firstly, dissolved gases present in

112

CSH

4.2. Deciding Which Model to Use


the initial droplet may come out of solution to form bubbles, or small bubbles may even be
present in the initial feed. As the pressure in the droplet falls, these bubbles will expand. In
the absence of such features, cavitation will occur when the pressure falls below the vapour
pressure of the liquid by an amount dictated by the tensile strength, (4.2.1). However
the bubble forms in a particular system, the present drying model postulates that a droplet
will cease thickening and enter the wet shell regime once Pbuck Pshell < Pcrit , where Pcrit
is some critical pressure. The complete process for deciding the relevant drying regime is
illustrated in Figure 4.4.
Wet Shell to Dry Shell Switch
As mentioned above, the dry shell model can directly follow the thickening regime when
Pbuck > Pmax . This is likely to occur for droplets containing large suspended solids, e.g.,
the detergent crutcher mix droplet simulated in Section 4.6.1. However, it is also possible
that the dry shell regime might be entered following a period of wet shell drying.
The wet shell model includes an expanding central vapour space with a slowly growing
outer shell. It is clear that if this regime persist for a sufficiently long time, the growing
bubble will meet the retreating inner shell surface. At this point, the bubble can expand no
more and further moisture removal necessitates drying of the shell itself. The new model
therefore executes a switch from wet shell to dry shell drying when b = S . An example of
this behaviour is presented in Section 4.6.2.
In general, it is possible that the switch between the wet and dry shell drying regimes
might occur before b = S . Such an eventuality leaves three spatial domains: a region
adjacent to the bubble, formerly the wet core; a region corresponding to the previous wet
shell; and the new, growing, dry shell region. Such a situation may also be handled in the
current model formulation and an example is shown in Section 4.6.2.

4.2.3 Internal Boiling When Drying in Hot Air


When the drying air has a temperature greater than the boiling point of the continuous
phase, Tboil , the possibility exists that the drying droplet will boil. Since the new drying
model developed in this thesis assumes a uniform temperature across the droplet, (3.2.1)
the condition for boiling of the continuous phase is simply Td > Tboil . The behaviour
following the start of boiling depends on the morphology of the droplet at that point.
Whilst the details of this dependence are beyond the scope of the current thesis, it is possible
for a number of observations to be made concerning likely drying behaviour.

113

CSH

4.2. Deciding Which Model to Use


Boiling with a Dry Shell
The assumption that there are no internal temperature gradients is approximately true, but
only holds for regions wetted by the continuous phase; in the dry shell model, the dried
shell is be expected to be hotter than the droplet average. Consequently, it seems reasonable
to assume boiling will commence at the evaporative front marking the edge of the hotter dry
shell region when the droplet average temperature, Td , rises above the boiling temperature,
Tboil . Making this assumption allows boiling in the presence of a dry shell to be simply
incorporated into the existing model framework provided that the rate of boiling is slow
enough to enable the vaporised moisture to escape through the porous shell. Assuming
that this condition is satisfied, the slow boiling model may be applied, as discussed in
Section 4.4.
Boiling causes the partial pressure of water vapour at the evaporative interface to rise.
The initial effect of this rise is to enhance the vapour flow away from the interface towards
the external surface. Such conditions can be modelled using the slow boiling model. However, if the pressure rise is sufficiently high then the droplet is expected to crack or shatter.
Obtaining a qualitative measure corresponding to this sufficiently high criterion requires
that the vapour flow through the shell be modelled. Such a calculation could, in principle, be incorporated within the new model framework. Structural information about the
shell provided by the moments of the solid particle population balance could then be used
to predict cracking or shattering of the droplet. However, this is beyond the scope of the
present thesis.
Boiling with No Shell or a Wet Shell
If the boiling condition is satisfied prior to shell formation, or during the wet shell or
thickening regimes, vaporisation is expected to commence within the drying droplet. This
results in bubble formation and droplet inflation or puffing. Such behaviour has been
observed by Walton and Mumford (1999b) who report that, at high drying air temperatures,
nearly all skin forming particles undergo inflation as a result of internal bubble nucleation.
Although the temperature is assumed uniform throughout the droplet, the temperature
at which the continuous phase boils is not necessarily constant. Rather, Tboil will be a
function of the continuous phase composition, with more dilute regions boiling at lower
temperatures. Consequently boiling is expected to commence in the most dilute regions
which, for a drying droplet, will be at the centre.22 This motivates the expansion of a
centrally located bubble in those drying models which attempt to simulate puffing, (e.g.,
Sano and Keey, 1982; Hecht and King, 2000b).
22

Whilst the continuous phase is most dilute at the droplet centre in most systems, this is not necessarily
the case for droplets which are already saturated at the start of drying.

114

CSH

4.3. Shell Thickening


Implementation of a sub-model to simulate drying during inflation is beyond the scope
of this thesis. However, the model presented here can indicate when such inflation is likely
to occur. Furthermore, it is noted that the new model framework developed in the present
work lends itself to future incorporation of such a sub-model; the infrastructure to handle
an expanding, centrally located bubble has been incorporated during implementation of the
wet shell model.

4.3 Shell Thickening


The previous section demonstrated the criteria for deciding which of the sub-models
thickening, dry shell, wet shell or slow boiling to use following formation of a surface
shell. Here and in the following two sections, each of these drying regimes is discussed in
more detail. Particular attention is paid to specifying how each sub-models is implemented
within the overall framework of the new droplet drying description introduced in this thesis.
A complete listing of all the equations in the present model in the precise form in which
they are solved is given in the following chapter.
This section focuses on the thickening regime which all droplets are hypothesized to
enter upon shell formation. As discussed in Section 4.1.2, there is experimental evidence
showing that some drying droplets continue to shrink following the appearance of a surface shell; contraction only ceases when the structural strength of the shell is sufficient to
withstand those collapsing forces acting on the droplet arising from continued evaporation, (4.2.2). The thickening model describes the droplet drying behaviour between the
initial appearance of a shell and the formation of rigid crust.

4.3.1 Model Description


The continuous phase continues to wet the shell during the thickening regime and consequently the equation which describes the behaviour of this phase, (3.3.16), is unchanged:


(1 ") (c) B (c)
t

1

(c)
2
(c)
(c)
(c)
2
(c)
(1 ") B
+ 2
r (1 ") v r B r Deff
r
r r
"
+ 0D
=0 .
(4.3.1)
t

115

CSH

4.3. Shell Thickening


Under ordinary circumstances it is assumed that there is no central bubble present during
the thickening regime and, consequently, f (t ) = 0 in (3.3.31), giving
v r (c) = Deff

B (c)

B + B (c)

"

1" r

!
.

(4.3.2)

Similarly, the equation describing the solids behaviour in the core region, (3.4.2) is unchanged, with v r(d ) = 0, giving

N+

(GN )

1
r2 r

r 2D

N
r

=0 .

(4.3.3)

In contrast, the solids within the thickening shell now behave differently. During the thickening regime, the particles in the shell are re-arranging themselves as the crust deforms. The
model allows for continued solids growth during this period, but does not seek to model
the precise nature of the spatial re-arrangement. Rather it is assumed that the solid particles
are no longer free to diffuse but are subjected to an imposed velocity,
v r(d )

 2
R dR
r

(4.3.4)

dt

arising from the deforming shell. This expression is derived from a volume balance on the
solids over an element in the shell, assuming that there is no accumulation as a result of
buckling. Equation (3.4.2) therefore reduces to

N+

(GN ) +

 2
R dR N
r

dt r

=0 ,

(4.3.5)

in the thickening shell.

4.3.2 Shell Growth


The continuous phase is assumed to continue wetting the surface particles although not
necessarily completely during the thickening regime and so evaporation proceeds at the
external surface of the droplet. Consequently, the external radius of the particle continues
to decrease according to (3.3.36), i.e.,
dR
dt

vap
m
0A

(4.3.6)

vap in this expression, is modAlgorithm 3.1, used to calculate the mass vaporisation rate m
ified to account for the vapour pressure reduction as the menisci recede between the solid

116

CSH

4.3. Shell Thickening


particles. The reduced partial pressure of moisture at the droplet surface, pA , is related to
that above a solids-free surface at the same temperature, T , by the Kelvin equation,
pA = pA exp

2V m,L cos

rM R g T

(4.3.7)

Here, rM is the radius of curvature of the surface menisci, as introduced in the previous
section, V m,L is the molar volume of the continuous phase and is the contact angle. This
correction becomes significant as the radii of the surface menisci approaches the hydraulic
radius of the porous shell, rH .
The rate at which the thickening shell grows is calculated by considering a number
balance on the solid particles in the region [S (t ) , R (t )]. Assuming that nucleation of new
particles does not occur within the shell region, the total number of particles may only
change as a result of flux across the interface at r = S (t ), i.e.,
S

dS
dt

m0 |S =

1 d

m0 d V
4 dt Vshell
Z
d R(t ) 2
=
r m0 d r
dt S(t )
=R

dR
dt

m0 |R S

dS
dt

r2

m0 |S + +
|

m0
t
{z

dr ,
}

(4.3.8)

=0

where the Leibniz integral rule has been used to write the last line. In these equations,
S = S r represents a radial location just inside the thickening shell, and S + = S + r
a location just outside. Making the further assumption that the number density of particles
within the growing shell is spatially uniform and unchanging with time allows the final term
in (4.3.8) to be set equal to zero. The shell growth rate is then given by
dS
dt

m0 | R
m0 |S + m0 |S

 2
R dR
S

dt

m0 |S 6= m0 |S + .

(4.3.9)

It is also possible for the thickening shell to grow through the process of sink diffusion,
where the inner boundary of the shell is viewed as a sink for randomly diffusing particles.
However, the contribution made by this process will be significantly less than the term on
the right hand side of (4.3.9), and further discussion is deferred until Section 4.5.3.
Assuming that the solid particle number density is temporally constant within the shell
implies that it remains equal to its initial value. The zeroth moment of the particle number
density function throughout the shell therefore equals the value at droplet edge at the point

117

CSH

4.3. Shell Thickening


of shell formation, i.e.,


m0 t tshell , r [S (t ) , R (t )] = m0 r = R t = tshell .

(4.3.10)

The rational behind assuming a constant particle number density to define the shell growth
rate as opposed to, say, a constant solids volume fraction, is that no new particles are allowed
to nucleate in the shell region. In contrast, growth of existing particles is still permitted and
the solids volume fraction in the shell region is expected to continue increasing. In situations
where only one particle size is present, such as the colloidal silica example discussed in
Section 4.6.2, all moments are constant in the thickening shell region.

4.3.3 Boundary Conditions


At the point of shell formation, that is, when " = "crit at r = R, the shell thickness, Tmin , is
set equal to the diameter of the smallest of the suspended particles, (Figure 4.1b). This sets
up two spatial domains, termed the central core region, [b , S], and the wet- or thickeningshell, [S, R]. The five partial differential equations describing the time evolution of the
solute mass fraction and moments of the solid particle number density are solved in both of
these domains. In effect, following shell formation, there are therefore ten coupled partial
differential equations to be solved, each requiring two boundary conditions.
Continuous Phase Equation
In the thickening regime, the solute boundary condition at the external droplet surface,
(3.3.35), remains unchanged:

B (c)

r

=
r =R

00
B (c) m
(c) Deff

(4.3.11)

However, the appropriate boundary conditions to be implemented at the internal surface of


the thickening shell are more complicated. Considering the superficial solute flux at r = S
gives

(c)
nB r

r =S



= (1 ") nB r (c)

r =S

= (1 ") v r (c) (c) B (c)


(1 ") B (c)
Deff (c)
r

= (1 ") nB r (c)

(4.3.12a)

r =S(t )

+ (c) (1 ") B (c)

118

dS
dt

(4.3.12b)

CSH

4.3. Shell Thickening


which, on substituting for v r (c) from (4.3.2) may be re-arranged to give
(c)

B
r


(c)
n

dS
Br
r =S(t )

(c)
=
+
B
.
(c)
(c)
dt
Deff

0A

(4.3.13)

Note that the intrinsic solute mass fraction and, consequently, the continuous phase density
must be the same either side of the inner shell wall at r = S . Further, the solute mass flux
across the growing shell boundary must be continuous, that is, it is required that


B (c)

r =S


(c)
=


B
+

(4.3.14a)

r =S

and

(c)
nB r

r =S (t )


(c)
= nB r

r =S + (t )


nB r
(c)

r =S + (t )

1 "

1 "+


nB r
(c)

r =S (t )

(4.3.14b)

In these last two equations, S = S r represents a radial location just inside the thickening shell, and S + = S + r a location just outside; " and "+ denote the solids volume
fraction at S and S + respectively.
Equation (4.3.13) may now be evaluated at both S and S + which, on substitution
into (4.3.14b), can then be re-arranged to give a relationship between the solute mass fraction gradients on either side of the growing shell interface,

B (c)

r

=
r =S

+
Deff

Deff

1 "+
1 "


B (c)

r

r =S

0AB (c)

Deff
(c)

" "+
1 "

dS
dt

. (4.3.15)

This, along with (4.3.14a), gives the two boundary conditions required for the solute equation at the shell interface. The boundary condition at the centre of the droplet, assuming no
bubble growth in the thickening regime, remains unchanged, i.e., a zero-gradient symmetry
condition is applied.
Discrete Phase Equations
The appropriate boundary conditions for the moment system in the shell region are those
of zero spatial gradient at both ends. This arises from the assumption that spatial transport
of solids in the shell is purely advective, (4.3.5). The moment boundary conditions in the
central core are also those of zero spatial gradient at both ends, although inclusion of sink
diffusion at the shell interface modifies this. As mentioned previously, a thorough discussion
of the moment boundary conditions in the central region is deferred until Section 4.5.3.

119

CSH

4.4. Dry Shell and Slow Boiling Models

4.4 Dry Shell and Slow Boiling Models


4.4.1 Model Description
The dry shell model is, in its simplest form, a classical shrinking core type analysis whereby
a central core of wetted material contracts as moisture evaporates from a receding interface, (Audu and Jeffreys, 1975; Cheong et al., 1986). The dry shell is defined to be the
region beyond this wetted core. As suggested by the schematic of dried-particle morphologies in Figure 2.9, droplets drying via the dry shell route might be expected to form solid
particles or, at high temperatures, they might shatter. Although the initiation of shell formation may be signalled by some critical solids volume fraction at the droplet surface, it is
noted that the dry shell model does not explicitly place a requirement on the solids volume
fraction within the growing shell region. The porosity of dry shell particles might therefore
be expected to vary perhaps considerably with position. In contrast to several previous implementations which assumed constant dry crust properties, (e.g., Cheong et al.,
1986; Dalmaz et al., 2007), porosity variations in the dry shell are tracked in the present
model.
In the present implementation of the dry shell model, the shell is assumed to be completely dry, that is, there is no solvent or solute remaining in the crust region. Further, it is
assumed that there is no vapour in the wet core and, consequently, the continuous phase
remains funicular throughout drying, as illustrated in Figure 4.6. Within the wet core,
the equations to be solved for the solute mass fraction and the moments of the population
balance are the same as outlined in Chapter 3. As demonstrated in the Section 4.6.1, continuing to solve these equations in the wet region allows the properties of the growing dry
crust to be predicted.
If the temperature of the droplet rises above the boiling point of the continuous phase
during the dry shell regime then, as discussed in Section 4.2.3, the droplet is assumed to
undergo slow boiling. Two key assumptions underpin this sub-model: first, the pressure
increase arising from continuous phase boiling is assumed to be lower than that which
would cause the droplet to shatter; secondly, the boiling rate is assumed to be heat transfer
limited, that is, all energy supplied to the droplet is assumed to go towards vaporising
the continuous phase. Improving upon these assumptions would require vapour transport
through the dry shell to be modelled and is beyond the scope of this thesis.

4.4.2 Shell Growth


The dry shell is assumed to grow according to
dS
dt

 2
R

00
m

0A 1 "|S

(4.4.1)

,

120

CSH

4.4. Dry Shell and Slow Boiling Models

wet
core

r
wet core

dry shell
region

S
R

Figure 4.6: Illustration of a droplet of radius R drying through a dry shell with internal radius, S . The
shell region is completely dry, containing no solvent or solute. The wet core remains free of vapour and thus
the continuous phase remains funicular throughout drying.

vap
m

S
w
wA,S A
wA,sur
wA,
r
S R
Figure 4.7: Evaporation from a droplet of radius R in the presence of a dry shell. Evaporation proceeds
from a front located at r = S(t ), which is receding towards the centre of the droplet. The graph below the
diagram shows illustrative radial profiles of the water mass fraction, wA, in the vapour-filled pores and
surrounding gas.

where the solids volume fraction is evaluated at the receding interface, r = S (t ). The
00 still refers to the evaporative solvent flux at the
ratio of radii accounts for the fact that m
external surface whereas the moisture actually evaporates from a front within the drying
droplet. Under such circumstances it is necessary to consider the increased resistance to
mass transfer resulting from the presence of a dry shell, (Incropera and DeWitt, 2002).
This situation is illustrated in Figure 4.7, where it is seen that the moisture mass fraction
adjacent to the evaporative interface, wA,S , falls to wA,sur at the surface of the droplet before
dropping to wA, in the bulk gas.
The additional mass transfer resistance from an internal evaporative front may be char-

121

CSH

4.4. Dry Shell and Slow Boiling Models


acterised by the effective diffusion coefficient of solvent vapour through the dried shell
region, Deff . This may be related to the binary diffusion coefficient of water in air, D A,air ,
via
Deff =

(1 ") D A,air

(4.4.2)

where is the tortuosity of the porous shell, (Cussler, 1997). Assuming a pseudo-steady
shell thickness, (3.5.4) can be integrated over the dry region to give

vap = 4 Deff
m

1
R

1

log

1 wA,sur

1 wA,S

(4.4.3)

where wA,sur is unknown. The transport from the droplet surface to the bulk can be analysed
as in Section 3.5.1 to obtain
!
1 wA,
vap = 2D A,air Sh R log
m
,
1 wA,sur
which is the same as (3.5.10). Using this to substitute for wA,sur in (4.4.3) gives, after
re-arrangement,

1
D A,air S R
vap = 4R2 m
00 = 2D A,air R
m
+ log (1 + BM ) .
Deff
2S
Sh

(4.4.4)

As expected, this expression reduces to (3.5.10) when S = R, i.e., when there is no dry shell
surrounding the droplet.
During the slow boiling regime, all energy supplied to the droplet is assumed to go towards vaporising the continuous phase. The expression for the mass vaporisation rate (4.4.4),
is therefore modified to read
vap = 4R2 m
00 =
m

Q
H vap

(4.4.5)

where Q is the heat penetrating into the droplet and H vap is the latent heat of vaporisation
of the continuous phase. The dry shell growth rate in the boiling regime is still obtained
00 now given by (4.4.5). An example of this behaviour is presented in
from (4.4.1) with m
Section 4.6.2.

122

CSH

4.5. Wet Shell Model

4.4.3 Boundary Conditions


The symmetry boundary conditions in the centre of the droplet remain unchanged once a
dry shell has formed. However, the external boundary conditions, now at r = S (t ), require
modification. The solute boundary condition at this outer edge of the wet central core
becomes

00
B (c) m
B (c)
=
(4.4.6)

,
r
(c) Deff 1 "|
r =S

00 is given by (4.4.4) or (4.4.5) in the dry shell and slow boiling regimes respecwhere m
tively. Comparison with (3.3.35) shows that the only difference between this expression
and the external solute boundary condition prior to shell formation is the inclusion of the
1
1 "|S
term. The outer boundary condition on the moment system is now

N

r

(4.4.7)

=0 ,
r =R

as no solids are allowed to enter or leave the dried region. This completes the description of
the dry shell and slow boiling models.

4.5 Wet Shell Model


4.5.1 Model Description
The key assumption of the wet shell model is that the solids in the shell region remain
wetted by the continuous phase and, consequently, the evaporative front remains at the
droplet surface. Continued solvent evaporation from an unshrinking droplet requires the
presence of a growing vapour space. In the wet shell model, this takes the form of a single,
centrally located bubble, as illustrated in Figure 4.8.
As the continuous phase wets all the solid particles in this model, the shell itself cannot
be identified in terms of a dry region as was done in the dry shell approach. Instead the
shell is defined as the region which has a solids volume fraction higher than some critical
value. This value is normally taken to be the same as the critical solids volume fraction
triggering shell formation, "crit , although it is noted that this need not be the case. Figure 4.8
demonstrates how the droplet is divided into the core and wet shell regions, separated by
the internal edge of the growing shell at r = S (t ).
Volume conservation requires that the rate of growth of the central bubble must be

123

CSH

4.5. Wet Shell Model


central
bubble

wet shell
region

core

r
S

S+

r
S

Figure 4.8: Illustration of a droplet of radius, R drying through a wet shell with internal radius, S .
The solid particles in the wet shell remain wetted by the continuous phase and a central bubble grows as
evaporation proceeds.

related to the evaporative moisture flux according to


db
dt

00
R2 m
b 2 0A

(4.5.1)

b >0 .

From this expression, it is clear that a seed bubble of finite size is required to avoid an infinite
initial growth rate. Physically, this is likely to be an air bubble present in the feed material
or introduced during spraying. The central bubble is filled with vapour-saturated drying
air, not pure solvent vapour. That is, the partial pressure of solvent in the bubble will be in
equilibrium with that in the adjacent droplet. As the bubble grows, it is assumed that the
mass of solvent in the bubble remains small compared with the amount in the surrounding
droplet. Consequently, the moisture flux to the bubble is negligible and has no effect on the
surrounding concentration gradients.
The growing bubble is modelled as imposing an outward advective velocity on the
droplet. The bulk continuous-phase velocity, (3.3.31), is modified to read
v r(c) = Deff

B (c)

B + B (c)

"

1" r

!
+

b 2 db
r 2 dt

(4.5.2)

where the final term is the additional advective component due to the bubble. With this
modification to v r(c) , the continuous phase equation, (3.3.16), can be applied without further alteration in both the shell and core regions.
The evolution equation for the population of solid particles, (3.4.2), has the same additional advective component as a result of the growing bubble. Therefore, the equation

124

CSH

4.5. Wet Shell Model


describing the solids in the core region now reads

N+

(GN ) +

b 2 db N
r 2 dt r

1
r2 r

r D

N
r

=0 .

(4.5.3)

Within the shell region, the solid particles are no longer free to move at all; the particle
population evolves purely as a result of crystallisation from the continuous phase and (3.4.2)
reduces to

N+

(4.5.4)

(GN ) = 0 .

Nucleation of new particles within the wet shell is possible, but this is thermodynamically
unlikely in a region where, by definition, the volume fraction of solids is already high.
Nucleation is therefore neglected.

4.5.2 Shell Growth


The shell grows as a result of solid particles depositing on the inside wall at r = S .
From (4.5.3) it is seen that the particle number density flux in the space of the external
co-ordinate is given by
Nr =

b 2 db
2

r dt

N D

N
r

(4.5.5)

If the advective flux of solids is assumed to dominate the deposition process, then
Nr

b 2 db
r 2 dt

(4.5.6)

N,

and so, following the growing shell interface at r = S(t ), it is possible to write
N

b 2 db
S 2 dt

= N N+

 dS
dt

(4.5.7)

Here N and N + represent the particle number density at S and S + respectively. Rearranged, this gives the shell growth rate as
dS
dt

b 2 db

S 2 dt N + N

(4.5.8)

.

125

CSH

4.5. Wet Shell Model


Note that this expression implies that the gradient of the particle number density is zero at
the shell interface, i.e.,

N
=0 .
(4.5.9)

r
r =S

This might be viewed as unrealistic as the growing boundary also acts as a sink for diffusing
solid particles. This effect can be captured through the incorporation of a sink diffusion
term (Hansson, 2003; Won et al., 2001). It is assumed that the rate of sink diffusion may
be described by the simple model
Nsink = ksink N ,

(4.5.10)

which, included in (4.5.6), gives a new solids flux to the growing wall,
Nr =

b 2 db
S 2 dt

(4.5.11)

N + ksink N .

The wet shell growth rate is then modified to read


dS
dt

N+ N

b 2 db
S 2 dt

+ ksink

(4.5.12)

which, in terms of the moments, (3.4.22), is


dS
dt

ma
ma+ ma

b 2 db
S 2 dt

+ ksink

(4.5.13)

4.5.3 Boundary Conditions


The existence of two physical domains following the formation of a wet shell the central
core and the wet shell region requires that four boundary conditions be specified for each
of the equations, one at either end of both domains.
Continuous Phase Equation
The solute boundary condition at the external droplet surface remains unchanged from the
base case considered in Chapter 3, i.e., (3.3.35) continues to be used once the moisture
vap has been suitably modified using (4.3.7) to account for the vapour
evaporation rate, m
pressure reduction caused by the wet shell. The appropriate boundary conditions to be
implemented at the internal surface of the wet shell are similar to those in the shell thickening sub-model, as discussed in Section 4.3.3. Substituting the revised expression for the

126

CSH

4.5. Wet Shell Model


continuous phase velocity, (4.5.2), in to the superficial solute flux balance at r = S (t ) gives

(1 ") nB r
(c)


(1 ") B (c)
r =S
r
(4.5.14)

dS
(c)
(c)
(c)
,
= (1 ") nB r
+ (1 ") B
r =S(t )
dt
= (1 ") v r(c) (c) B (c) Deff (c)

and so (4.3.13) becomes


B (c)
r

nB r (c) r =S(t )
0A

dS b 2 db
(c)

.
=

+
B

dt
S 2 dt
Deff (c)
(c)

|
{z
}

(4.5.15)

vs

Imposing continuity of the solute mass flux across the growing shell and re-arranging as in
Section 4.3.3 gives the boundary condition

B (c)

r

=
r =S

+
Deff

Deff

1 "+
1 "


B (c)

r

r =S

0AB (c)

(c)
Deff

" "+
1 "

v s , (4.5.16a)

where
vs =

dS
dt

b 2 db
S 2 dt

(4.5.16b)

Solute mass fraction continuity across the shell interface gives the second boundary condition required at r = S ,

(c)

(c)+
= B
.

(4.5.16c)

The fourth boundary condition is that applied at the bubble interface at r = b . The
moisture content of the gas in the bubble is assumed to be negligible compared with that
in the adjacent wet core, (Sano and Keey, 1982) The flux of moisture from the droplet to
the bubble is negligible and, consequently, the presence of the expanding bubble is assumed
to have no effect on the concentration gradients at r = b . That is, the solute boundary
condition at this interface is unchanged from the symmetry condition used previously,

B (c)

r

(4.5.17)

=0 .
r =b

127

CSH

4.6. Applications
Discrete Phase Equations
The solid particles in the wet-centre of the droplet remain free to move once the outer shell
has formed. When these particles contact the inner wall of the shell through either
advection or the sink diffusion process discussed above they aggregate and contribute
to shell growth. This is modelled by introducing a death-rate boundary condition at r =
S , which is obtained directly from the sink diffusion condition for the number density
function, (4.5.10). The relevant equations for the moments therefore read

"
= ksink " ,

r r =S

ma
= ksink ma

r

(4.5.18a)
a {0, 1, 2} .

(4.5.18b)

r =S

From the definition of the wet shell region the region where the solids volume fraction exceeds the critical value, "crit it is clear that the appropriate boundary condition for
the solids volume fraction at r = S + is
"+ = "crit .

(4.5.19a)

Re-arranging (4.5.13) then gives the boundary conditions for the remaining moments at
r = S + as

2
1
db
b
dS
, a {0, 1, 2} ,
ma+ = ma 1
+ ksink
(4.5.19b)
2
dt
S dt
where the shell growth rate is
dS
dt

"
"crit "

b 2 db
S 2 dt

+ ksink

(4.5.20)

The specification of the wet shell model is completed by applying zero gradient boundary conditions to all moments of the solids number density at the both the droplet external
radius and at the bubble interface.

4.6 Applications
Two physical systems are simulated using the new model and the results obtained compared
with experimental data from the literature. Both examples presented focus on aspects of the
model relating to shell formation and the subsequent drying behaviour.

128

CSH

4.6. Applications

4.6.1 Drying a Detergent Droplet


The first test case is the simulation of a droplet of detergent slurry drying. Such crutcher
mix is typically a complex mixture of around 10 components, with precise formulations
being closely guarded commercial secrets, (de Groot et al., 1995). Griffith et al. (2008) have
recently conducted a series of experiments investigating the drying of a generic crutcher mix
formulation, the composition of which is shown in Table 4.1. Griffith et al. report that the
droplets they observed showed no shrinkage during drying which indicates that this system
might be a suitable choice to demonstrate the dry shell model.
Table 4.1: The composition of the detergent slurry investigated by Griffith et al. (2008).

Component
Initial Mass Fraction
LAS
0.09
Water
0.29
Acusol Polymer
0.03
Sodium Sulphate
0.35
Sodium Aluminosilicate
0.24
The methodology described in Section 4.2.2 can be applied to further motivate the use
of the dry shell approach for this example. Assume that the solid particle population in
the crutcher mix is mono-disperse with a diameter of 10 m. This represents the base case
described below. Further, assume that shell formation is triggered at a critical solids volume
fraction, "crit = 0.65. The hydraulic radius of the capillaries in this freshly formed shell
can be calculated from (4.2.6), giving rH = 1.79 m. According to (4.2.4), the maximum
pressure drop capable of being supported by the menisci in such capillaries is Pmax =
80 kPa.
Bortolotti et al. (1992) reports the Youngs moduli for a number of different zeolite
containing detergent powders; these are all of the order of 10 MPa. Using this value along
with = 0.3 in (4.2.13) gives, on setting Pbuck = Pmax and assuming that T is small
compared with the droplet radius,
(4.6.1)

Tcrit 0.0752 R .

The result suggests that the assumption regarding the relative magnitudes of droplet radius
and critical shell thickness is justified. For a 1.5 mm droplet, the critical shell thickness
is less than six solid particle diameters. Therefore, following the methodology outlined in
Section 4.2.2, the thickening regime is predicted to be negligibly long. In such situations,
there are computational advantages associated with dictating that the simulation initiates
the dry shell model immediately on shell formation namely, this avoids the need for
creating the second co-ordinate system. This is the approach taken for the simulations
presented in this section.
129

CSH

4.6. Applications

CONTINUOUS PHASE
neat

1.0

SOLID PHASE

CONTINUOUS PHASE

lye

1.0

0.9

SOLID PHASE

lye

0.9

0.8

0.8

zeolite

0.7
0.6
0.5
0.4

zeolite

0.7

water

mass fraction

mass fraction

neat

LAS

0.3

water

0.6
0.5
0.4

LAS
LAS

0.3
Na2SO4

polymer

Na2SO4

0.2

0.2
Na2SO4

Na2SO4

0.1

0.1

0
0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

mass fraction

mass fraction

(a)

(b)

Figure 4.9: (a) Initial composition of the crutcher mix droplets as measured by Griffith et al. (2007);
and (b) simplified description of the crutcher mix used to model the system in the base case.

The description of the crutcher mix system needs to be further simplified before the
present model can be used. Griffith et al. (2007) report that the continuous phase actually
comprised individual neat and lye phases which were rich in LAS and sodium sulphate
respectively. This is illustrated in Figure 4.9a. To simulate this system within the present
framework, the lye phase was modelled as an aqueous sodium sulphate solution. Similarly,
the small amount of sodium sulphate in the neat phase is ignored and this is considered
to be a LASwater binary. This further simplified system is illustrated Figure 4.9b, and
represents the initial composition used for the simulations presented in this section.
In terms of the present model, the solid phase, D , comprises the zeolite sodium
aluminosilicate and some crystallised sodium sulphate and the water in the system is the
solvent, A. The LAS and dissolved sodium sulphate must be described as component B . For
the remainder of this section, these two components will collectively be referred to as the
solute. However, since only sodium sulphate crystallises to form new solid, it is important
to track the mass fraction of the combined solute that is Na2 SO4 , i.e.,
=

mass of Sodium Sulphate


mass of Sodium Sulphate + mass of LAS

(4.6.2)

Because of the simplifying assumptions made above, also represents the mass fraction of
the solute in the lye phase. Assuming that this fraction is uniform throughout the droplet
allows to be simply calculated from the initial masses and knowledge of the mass of
Na2 SO4 that has crystallised during the drying period. This assumption is justifiable as the
rate of moisture removal is slow when drying at Tgas = 60C. For the droplets observed

130

CSH

4.6. Applications
by Griffith et al. (2008), (t = 0) = 0.56.
NMR results obtained by Griffith et al. (2007) demonstrate that when drying crutcher
mix, water is initially lost from the lye phase. Assuming that the ratio of LAS to water in the
neat phase remains unchanged until the lye phase has dried allows the mass concentration
(c)
of sodium sulphate in the lye phase, Na
, to be determined from :
B (c)

(c)

Na =

1 1.43B (c) (1 )

(4.6.3)

(c)
Initially, Na
= 0.31, which is the saturated mass fraction at 60 C, (Wetmore and LeRoy,
1951). This is to be expected as the aqueous and crystallised sodium sulphate are initially
in equilibrium. Sodium sulphate crystallises out of solution as the drying proceeds. In the
absence of any data relating to the model crutcher mix system, the kinetics used to describe
this process are those reported by Rosenblatt et al. (1984) for direct crystallisation to the
solid in a sodium sulphatewater mixture. First introduced in Section 3.6.2, the relevant
expression is

dL
dt


3

= 1.484 10 exp


57.4
RT

Ci Ceqm

1.5

(4.6.4)

which corresponds to the linear growth rate, G , in the population balance equation (3.4.2).
Ci and Ceqm are the local and saturated sodium sulphate concentrations respectively in
kmol m3 .
As discussed in Section 3.4, the moment system describing the solid phase allows nucleation of new particles to be modelled. However, the presence of a considerable amount
of crystallised sodium sulphate at the outset of drying in the detergent system makes nucleation of new crystals thermodynamically unlikely. Therefore, the nucleation rate, N0 ,
is set to zero in this simulation. In turn, this avoids the difficulties discussed at length in
Section 3.6.2 associated with determining a suitable value for Nmax .
The solids contained in the crutcher mix droplets are relatively large typically between
2 and 50 m in diameter, (Bayly, 2008) and, consequently, are relatively immobile. For
simplicity, the solid particles are here assumed initially mono-disperse with a size of 10 m.
A single, non-size dependent, solids diffusion coefficient is then used, D = 1015 m2 s1 .
Griffith et al. (2008) determined an effective diffusion coefficient appropriate for modelling the movement of water in crutcher mix,

Deff = exp

27.5 + 174.5u
1 + 8.5u

(4.6.5)

which, as expected, is a strong function of the dry mass basis moisture content, u .

131

The

CSH

4.6. Applications
sorption isotherm is required to obtain the rate of moisture evaporation from the drying
droplet.23 The isotherm was obtained by fitting a standard sorption isotherm to measurements on the crutcher mix obtained by Bayly (2007). The resulting equation is
HR =

pA
p sat

= f A

0.8 A
1 + 11A

+ 1 f
A

0.05 + exp

A 3.92
0.15

, (4.6.6)

where
1

f A =
1 tanh 100 A 0.25
2

(4.6.7)

is a blending function required to join the different functional forms fitted at high and
low moisture contents. The isotherm is plotted in Figure 4.10 along with the associated
experimental data.
A crutcher mix droplet with an initial diameter of 1.5 mm was simulated drying in
dehumidified air at 60 C with a relative velocity of 1.68 ms1 . Figure 4.11 shows the simulated moisture mass fraction compared with experimental measurements of a droplet drying
under these conditions from Griffith (2008). The simulated moisture content matches the
experimental results very well, although the errors associated with the measured data were

1
0.9

Relative Humidity / []

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Water Mass Fraction / []

Figure 4.10: Water sorption isotherm for crutcher mix. The line shows the sorption isotherm, (4.6.6),
obtained by fitting to the experimental points obtained by Bayly (2007).
23

The sorption isotherm is evaluated by the function S_ which is called by Algorithm 3.1 in
the course of determining the moisture evaporation rate.

132

CSH

4.6. Applications

30

Moisture Mass Fraction / []

Simulated Moisture Mass Fraction


Experimental Moisture Mass Fraction
25

20

15

10

0
0

500

1000

1500

2000

2500

3000

3500

Time / s

Figure 4.11: Simulated evolution of the moisture mass fraction in a crutcher mix droplet (line) compared
with experimentally measured values from Griffith (2008), (symbols).

not reported.
The use of NMR techniques to follow the droplet drying by Griffith (2008) allows for
further validation of the current model. Figure 4.12 shows a comparison between experimentally measured moisture profiles and those extracted from the simulation. To render the
model results suitable for comparison with the experimental data, it is necessary to apply
two Abel transforms, (Bracewell, 1990). Details of the Abel transformation, the routine
used to perform the projection and its associated accuracy are contained in Appendix A.
The NMR data returns intensity readings in arbitrary units. Therefore, assuming an initially homogeneous spherical particle, an appropriate scaling factor was determined using
least squares minimisation and applied to all the simulated profiles. It is clear that the experimental droplet was not a perfect sphere, but the error associated with this approximation
is small.
The model fit to the measured profiles is fair. The intensity maxima at the centre of the
droplet are very well predicted for the 10 and 20 minute profiles, but the predicted profile at
30 minutes does not match the data so well. Between 20 and 40 minutes, the experimental
data shows faster drying than predicted by the model as is demonstrated by the observation
that the 30 minutes data coincides with the 40 minutes predicted profile. It is possible that
this discrepancy is a result of the experimental data which appears to show particularly fast
drying between 20 and 40 minutes. However it could also be that the dry shell assumption
used in the model is not entirely compatible with this system, or that water associated with
the polymer not considered in the simplified model becomes important in the later
133

CSH

4.6. Applications

1
0 min
10 min
20 min
30 min
40 min
50 min
0 min
10 min
20 min
30 min
40 min
50 min

0.9

Water Signal / [a.u.]

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1

1.5

2.5

Radial Position / m

3.5

4
3

x 10

Figure 4.12: Simulated moisture profiles in a drying detergent droplet (lines) compared with experimental
observations (symbols) from Griffith (2008). Measured and simulated profiles are displayed at 10 minute
intervals.

stages of drying.
Figure 4.12 clearly shows the influence of the dry shell assumption on the model results.
A dry shell contains no moisture and consequently no NMR signal would be observed
from such a region. The simulated profiles reflect this, showing the diameter of the wet
core shrinking with time. Such an effect is, however, not clear in the experimental data,
suggesting that an extension to the dry shell approach may be required for this system. One
idea is to allow for a damp shell, whereby a certain fraction of the pore-volume in the shell
remains filled by the continuous phase.
The new droplet drying model gives additional information about the drying droplet
not immediately apparent from the experimental data. For example, Figure 4.13 shows
the time evolution of the individual component masses. The mass of solids increases as
sodium sulphate crystalises out of solution. This is also reflected in the decreasing solute
mass, which tends to a constant value which equals the mass of LAS in the system.
Figure 4.14 shows the position of the receding dry shell interface as drying proceeds.
As demonstrated by (4.4.4), the growing dry shell acts as a resistance to further moisture
evaporation. In the simulation, the tortuosity of the shell is taken as = 10, which is
representative of the values measured by Griffith et al. (2007). This value is used in (4.4.2)
to give the effective diffusion coefficient of moisture through the dried shell. The shell solids
volume fraction used in this expression is the average value of " in the shell region.
Figure 4.15 shows the solids volume fraction profiles through the droplet drying history.

134

CSH

4.6. Applications

4
Solvent Mass
Solute Mass
Solids Mass
Total Mass

3.5

Mass / mg

3
2.5
2
1.5
1
0.5
0
0

500

1000

1500

2000

2500

3000

3500

Time / s

Figure 4.13: Simulated evolution of total droplet mass together with the mass of each of the three components in the model droplet.

x 10

0.9

Radial Distance / m

0.8
0.7
0.6
0.5
0.4
Dry Shell Interface
Outer Droplet Radius

0.3
0.2
0.1
0
0

500

1000

1500

2000

2500

3000

3500

Time / s

Figure 4.14: Simulated time evolution of the dry shell interface during the drying of a droplet of crutcher
mix.

135

CSH

4.6. Applications

0 mins

10 mins
0.6

/ []

/ []

0.6

0.5

S (t = 0 mins)

0.4

0.3

S (t = 10 mins)
0.5

0.4

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.3

0.8

0.1

0.2

Radial Position / mm

0.6

/ []

/ []

S (t = 20 mins)

0.5

0.6

0.7

0.8

0.6

0.7

0.8

0.6

0.7

0.8

S (t = 30 mins)

0.5

0.3
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.1

0.2

0.3

0.4

0.5

Radial Position / mm

Radial Position / mm
0.7

0.7

50 mins

40 mins
0.6

/ []

0.6

/ []

0.5

0.4

0.4

0.5

S (t = 40 mins)

0.4

0.3

0.4

30 mins

20 mins
0.6

0.3

0.3

Radial Position / mm

0.1

0.2

0.3

0.4

0.5

S (t = 50 mins)

0.4

0.5

0.6

0.7

0.8

0.3

Radial Position / mm

0.1

0.2

0.3

0.4

0.5

Radial Position / mm

Figure 4.15: Simulated solids volume fraction, ", profiles at 10 minute intervals during the drying of a
crutcher mix droplet. The diamond marker in each subplot shows the position of the inner edge of the dry
shell at the corresponding time.

Initially, the crutcher mix droplet is homogeneous with a uniform solids volume fraction
of 0.36. As the solid particles are relatively immobile, the surface solids volume fraction
quickly reaches 0.65 and shell formation is triggered almost immediately. The rigid shell
prevents further shrinkage of the droplet and, following the discussion at the start of this
section, the dry shell model is immediately applied. The solids volume fraction behind this
initial shell is still quite low. The dry shell is therefore predicted to have a loosely packed
region immediately behind the external skin; recall that the dry shell model in contrast to
the wet shell model does not place an explicit requirement on the solids volume fraction
within the growing shell region, (4.4.1).
As more water is removed, crystallisation causes the solids volume fraction to rise, as
is reflected in later profiles. This in turn causes the packing of the dry shell to increase
once more. Each subplot in Figure 4.15 contains a diamond indicating the position of the
evaporative front at the relevant time. The solids volume fraction is, as expected, continuous
across this dry shell interface. This is a further difference in expected behaviour between the
wet and dry shell models.

136

CSH

4.6. Applications

Figure 4.16: SEM images showing detergent droplets dried in a spray dryer. The dried-particles are seen
to have a thin surface skin covering a core region in which the solids are loosely packed. Courtesy of Cheyne
et al. (2002)

By the plot at 40 minutes, all of the sodium sulphate has crystallised out of solution.
This corresponds with the removal of all the water from the lye phase. At this point, there
is no further mechanism within the model to increase the solids volume fraction and the
dried-particle is therefore predicted to have a loosely packed centre. The pore space in
this central region is completely filled by the neat phase, i.e., LAS and its associated water.
Continued evaporation removes the water from this neat phase and, in practice, some of
the LAS will solidify as it concentrates. Nevertheless, it still seems likely based on the model
predictions that the centre of the dried-particles will be less dense than their outer regions.
There is some evidence that the particle morphology predicted above is indeed observed
in real dried-particles. Figure 4.16 shows SEM images of two detergent droplets similar
to those simulated in the present study. The imaged droplets were dried at a higher temperature and therefore more quickly than those simulated, but they did not undergo
inflation. The particles have a fairly smooth surface skin. However, this skin is thin and
encloses a large central region where the solids are far less densely packed. Qualitatively, this
agrees well with the predictions from the simulation.
Tracking the sodium sulphate mass fraction in the solute, , allows profiles of the concentration of sodium sulphate in the lye phase to be reconstructed. Figure 4.17 shows some
such profiles plotted at 10 minute intervals through the drying of a crutcher mix droplet.
It is seen that the concentration stays relatively constant around 30wt% until around 40
minutes. At this point it falls quickly to zero across the entire droplet. As discussed above,
this leads to the cessation of solid crystallisation at the same point in the drying history. The
observation of a relatively constant sodium sulphate concentration agrees with experimental
results for similar systems. Griffith (2008) used 23 Na and 1 H NMR to observe the drying of
a vial of crutcher mix over an extended period. 23 Na NMR can track the amount of sodium
sulphate in solution as the vial dried. This was found to decrease in step with the free water

137

CSH

4.6. Applications

0 min
10 min
20 min
30 min
40 min
50 min

0.6
0.5
0.4
0.3
0.2

Na SO Lye Phase Mass Fraction / []

0.7

0.1
0
0

Radial Position / mm

8
4

x 10

Figure 4.17: Simulated lye phase sodium sulphate concentration profiles at 10 minute intervals during
the drying of a crutcher mix droplet.

content, indicating its concentration was approximately constant.


Conclusions
The methodology outlined in Section 4.2.2 was applied to motivate the use of the dry
shell model to simulate a drying droplet of detergent slurry. This system is of considerable
industrial interest and, although modelling a simplified system, several interesting results
are obtained. The continuous phase in this system is actually comprised of individual neat
and lye phases and the core droplet drying model is extended to simulate this. The resulting
prediction of the lye phase sodium sulphate mass fraction is in good qualitative agreement
with experimental observations and demonstrates the wide predictive power of the new
model.
For the first time, predictions from a mechanistic droplet drying model are compared
with spatially resolved experimental data. Comparisons of predicted and measured moisture
profiles across the droplets show fair agreement. However, the experimental data suggests
some moisture remains in all regions of the droplet throughout drying, raising questions
about the validity of using the dry shell sub-model.

4.6.2 Drying a Droplet of Colloidal Silica


The second test case simulated is the drying of a colloidal silica droplet containing 16 nm
particles suspended in water at an initial mass fraction of 30%. This system was introduced
138

CSH

4.6. Applications
in Section 3.6.1, where the drying was simulated prior to shell formation and compared
with experimental results from Neic and Vodnik (1991) at two different drying air temperatures. This comparison demonstrated that the results from the core drying model agreed
well with experimental data; this section extends that work to simulate moisture removal
following the formation of a surface shell and the associated formation of a dried-particle.
Walton and Mumford (1999a) identified colloidal silica as a material where droplets
formed dried-particles with an agglomerate structure. In general, such materials contain
suspended rather than dissolved solids. As no crystallisation occurs in such droplets, they
are found to continue shrinking upon drying, even after the volume fraction of solids rises
above the supposed critical value, i.e., after the formation of a surface shell as defined in Section 4.1.2. Furthermore, it is observed that moisture is removed by saturated surface drying
throughout this shrinking period and this continues for a short while after the droplet size
has stabilised. This behaviour suggests that this system is a good candidate for demonstrating the shell thickening and wet shell models.
Physical Properties and Adjustable Parameters
Many aspects of this system have already been discussed at length; the reader is referred to
Section 3.6.1 for details of the core simulation prior to shell formation. To model moisture removal and dried-particle formation following the appearance of a surface shell, the
Youngs modulus and Poissons ratio of the colloidal silica shell are required. These physical
properties describe the strength of the shell and, as discussed in Section 4.2.1, are required
to determine the buckling pressure, Pbuck . Smith et al. (1995) measured the shear modulus of wet silica gels, from which the Youngs modulus can be calculated, given knowledge
of the Poissons ratio, .24 Using = 0.2, as was measured by Scherer (1992) for similar
silica gels, gives the estimated Youngs modulus for the system as 1 MPa. This is relatively
low, suggesting that the shell should be expected to buckle repeatedly as the droplet dries
leading to considerable thickening. This agrees with separate observations that such gels
can compress to as little as one tenth of their original volume on drying, (Scherer, 1990).
However, it is noted that the films measured to obtain the figure of 1 MPa were composed
of 540 nm particles considerably larger than those in the present system. The sensitivity
of the results to this quantity is therefore investigated below.
The method described in Section 4.2 is used to select the appropriate model to use as the
droplet dries. To apply this methodology, two parameters are needed in addition to those
describing the structural properties of the shell. The first of these is the liquid-solid contact
24

The Youngs modulus, shear modulus and Poissons ratio are related by
(4.6.8)

E = 2G (1 ) .

139

CSH

4.6. Applications
angle, , for use in (4.2.4). Asmatulu (2002) reports a range of possible values, depending
on the hydrophobicity of the silica particles used. Hunter et al. (2007) give a value of 37 ,
obtained by averaging the results obtained from a number of experimental methods. This
is in broad agreement with the contact angles given by Asmatulu for silica particles with no
added cleaning agents and so this is the value taken for the present set of simulations.
The second remaining parameter is Pcrit , the critical pressure which determines when
a droplet will cease thickening and enter the wet shell regime, (see Figure 4.4). By coincidence, it transpires that Pcrit is not required when simulating these particular sets of
experimental data. This is because the drying droplet is predicted to enter the dry shell
regime directly following shell thickening, thus avoiding a wet shell drying period. However, in general, this parameter must be assigned a value and a method for doing this is
discussed below.
When the growing bubble meets the internal boundary of the wet shell region, the
model switches from the wet to dry shell mode as discussed in Section 4.2.2. The dry
shell tortuosity, , used in (4.4.4) when calculating the mass vaporisation rate is taken as
10 m m1 . This is based on the crust diffusion coefficient reported by Neic and Vodnik
(1991) for the system simulated here. When the temperature of the droplet reaches 100C,
the remaining water will boil. Given long enough in the dryer, all the remaining water will
be vaporised and the temperature of the dried-particles will rise to that of the drying air.
Comparing Model Predictions with Experimental Data
Direct comparisons with the experimental data of Neic and Vodnik (1991) at Tgas = 101 C
and Tgas = 178 C are presented in Figure 4.18 (a) and (b) respectively. This is the same
experimental data used in Section 3.6.1, although the simulations are now run until the
droplet is dry. Following the findings of the investigations reported in that section, the
drying air is here assumed to be dehumidified and moving with a velocity of 2.5 ms1
relative to the droplet. A shell is deemed to have formed once the solids volume fraction
at the droplet surface exceeds 0.65. With these assumptions, the simulated profiles prior to
shell formation are identical to those presented in Figures 3.13 and 3.15.
At both temperatures, the agreement between the predicted droplet mass and that measured experimentally is very good throughout the drying history. For the droplet drying in
air at 101 C, the modelled temperature under-predicts the experimentally measured value
following the formation of a shell at t = 58.2 s. It is likely that this discrepancy is caused
by the assumption of a uniform droplet temperature, even after the formation of a dry shell
at t = 70.2 s. Indeed, it is within this dry shell period that the temperature discrepancy is
most pronounced. In reality, it is expected that a dried shell region will be hotter than the
still-wet core; Farid (2003) goes so far as to suggest the temperature in this region will equal
that of the drying gas, although this is strongly disputed by Schlnder (2004). Nevertheless,

140

CSH

100

80

60

40

20

0
0

Droplet Temperature / C

Droplet Mass / mg

4.6. Applications

20

40

60

80

0
100

Time / s

250

200

150

100

50

0
0

10

20

30

40

50

Droplet Temperature / C

Droplet Mass / mg

(a)

0
60

Time / s

(b)
Figure 4.18: Simulated drying of a colloidal silica droplet (lines) compared with experimental results from
Neic and Vodnik (1991) (symbols) at: (a) Tgas = 101C and R0 = 0.972 mm; and (b) Tgas = 178C
and R0 = 0.95 mm.

141

CSH

4.6. Applications
it seems likely a dry shell region will be somewhat warmer than the core and, furthermore,
that the experimental thermocouple readings will reflect the presence of this hotter region.
Consequently, it is perhaps not surprising that the measured droplet temperatures are higher
than those predicted by the model as a result of the uniform temperature assumption.
Figure 4.18 (b) shows the results for a colloidal silica droplet drying in air at 178 C.
As the air temperature is above the boiling point of water, it is expected that this droplet
will undergo boiling; this is immediately clear from the figure. The switch to dry shell
drying, occurs at t = 33.2 s, at which point the droplet temperature begins to rise rapidly.
At t = 35.5 s, when the temperature reaches 100 C the boiling point of water the
slow boiling regime is said to commence, (4.4.2). The remaining water in the droplet is
removed by boiling, during which time the temperature remains constant. Once all the
moisture has been removed, at t = 41.4 s, the droplet mass ceases to decrease and the
temperature rises to that of the drying air.
The prediction of the droplet temperature for this example is seen to be in very good
general agreement with the experimentally observed values. It is noted that there is a fair
degree of scatter in the measured values following the predicted formation of a dried shell.
As discussed in relation to the results at the lower air temperature, this could reflect the
thermocouple variously returning readings reflecting the temperature in the still-wet core or
the hotter, dried shell. However, the general timescale over which the droplet temperature
rises to that of the drying gas is correct and, given the nature of the model, the predicted
profiles make good sense.
The sorption isotherm used to produce the results in Figure 4.18 is that given by (3.6.7).
However, as discussed in Section 4.3.2, this refers to evaporation from a fully wetted surface;
the Kelvin equation, (4.3.7), is applied to account for the vapour pressure reduction that
results as the surface menisci begin to retreat into the pores of the newly formed shell. For
most situations the resulting correction is small, but the small size of the colloidal particles in
this example means it may be significant here. Figure 4.19 shows the predicted temperature
history of a colloidal silica droplet drying in air at 178 C, both with and without the Kelvin
correction applied. Clearly, the effect of applying this correction is minimal in all but the
shell thickening region. It is not surprising that the vapour pressure reduction effect should
be greatest in the thickening region shown expanded in the sub-plot of Figure 4.19
as surface evaporation persists whilst the menisci start to withdraw into the pores.
Structural Development
As discussed in Section 4.2, the various pressure drops within the drying droplet are key
to determining its evolving structure. Figure 4.20 shows the evolution of the buckling
pressure, Pbuck , the pressure drop across the shell, Pshell and the maximum pressure drop
that can be supported across the surface menisci, Pmax . For an array of 16 nm particles, the
142

CSH

4.6. Applications

180
Droplet Temperature / C

100

160

Droplet Temperature / C

140
120
100

90
80
70
60
50
40
24

26

28

30

32

34

Time / s

80
60
40
Experiment
Prediction with Kelvin Correction
Prediction without Kelvin Correction

20
0

10

20

30

40

50

60

Time / s
Figure 4.19: Simulated temperature profiles for a 1.9 mm colloidal silica droplet drying in air at 178 C
(lines) compared with experimental results from Neic and Vodnik (1991), (symbols). The plot illustrates
the influence of vapour pressure reduction as modelled by the Kelvin equation, (4.3.7). The effect is greatest
in the shell thickening regime, expanded in the subplot.

hydraulic radius approximating the minimum pore size as given by (4.2.6) is 2.87 nm,
where the solids volume fraction is assumed to be 0.65. Using (4.2.1) with = 37 shows
the surface menisci in this system can support a pressure drop of over 40 MPa. Also shown
in the figure is the pressure in the wet core although, in this example, this does not change
during the thickening period. This figure can be used in conjunction with Figure 4.4 to
determine the morphological progression of the drying droplet. The pressure drop across
the shell is significantly greater than that required for buckling at all times. Thickening
therefore continues until the pressure drop across the shell equals Pmax , which is seen to
occur at t = 33.0 s. At this point the menisci retreat into the surface pores and, consequently,
the droplet is predicted to pass directly from the thickening to the dry shell regime.
Figure 4.21 shows the simulated structural evolution of the droplet drying in air with
Tgas = 178C. The different drying regimes are clearly visible and it is possible to see the
morphology of the droplet at a given time in the drying history. Thickening commences at
t = 24.4 s, after which the wet shell thickens considerably before, as just discussed, passing

143

CSH

4.6. Applications

50
45
End of Thickening Regime, t = 33 s

Maximum Pressure = 40 MPa

Pressure / MPa

40
35
30

Buckling Pressure
Pressure Drop Across the Shell
Core Pressure

25
20
15
10
5
0
24

26

28

30

32

34

Time / s

Figure 4.20: Plot showing various pressures relating to a droplet of colloidal silica drying in air at
Tgas = 178C during the shell thickening regime. The suspended silica particles have a diameter of
16 nm.

directly into the dry shell regime at t = 33.0 s. At this point of transition the droplet
has a region of still-wet core of radius 0.227 mm surrounded by a thickened shell region
extending out to the external radius at R = 0.599 mm. Once the dry shell regime has
commenced there are, in effect, three domains: the wet core; the thickened shell; and the
dry shell region. Rather than explicitly modelling three spatial domains, this situation is
handled here by setting the solids diffusion coefficient equal to zero in the thickened shell
that is, particles in the thickened shell are not free to move during the dry shell regime.
A dried-particle is said to have formed once all the moisture has been removed from the
droplet this occurs after 41.7 s.
Figure 4.22 shows the simulated evolution of the solids volume fraction within the
droplet of colloidal silica, drying with Tgas = 178C. Profiles are plotted at 5 s intervals.
Initially these profiles are flat across the droplet due to internal circulation prior to gelation.
After formation of a gel at " = 0.233, the profiles develop a pronounced curvature as particles build up at the receding interface. The profile shown in bold is that at t = 24.4 s when
the solids volume fraction at the surface reaches 0.65 and the shell forms. After this, the
droplet continues to shrink whilst the shell thickens. The solids volume fraction in this shell
region is 0.65 whilst the fraction in the core region is considerably less. This is clear from the
two profiles with discontinuities taken during the thickening regime. Close examination of
these two curves also shows that the solids volume fraction in the core immediately adjacent
to the thickening shell is slightly less than that a short distance further in. This is because

144

CSH

4.6. Applications

Thickening
Dry Shell
Boiling

0.9

Radial Distance / mm

0.8
0.7
0.6
0.5
0.4
0.3

Wet
Droplet Core

Dried
Particle

0.2

Thickening
Shell

0.1
0
0

10

20

30

40

50

60

Time / s

Figure 4.21: Simulated morphological evolution of a droplet of colloidal silica containing 16 nm particles
drying in air at 178C.

the growing shell boundary is modelled as a solids sink, as discussed in Section 4.5.2.
The interface at r = 0.227 mm between the wet core and thickened shell regions is
clear in the later profiles plotted in Figure 4.22. This is a result of the direct transition
between the thickening and dry shell regimes, discussed above. Bypassing a wet shell drying
period means that the resulting dried-particle is predicted to be solid but, as is clear from
Figure 4.22, the central region is expected to have a considerably lower solids content.
Although not presented here, the droplet drying in air at Tgas = 101C is predicted to
have a very similar morphology to that outlined above. The main reason for this is the small
size of the suspended solid particles, leading to high pressure drops across the thickening
shell as well as large maximum pressure drops across the surface menisci.
Effect of Suspended Particle Size
It is interesting to investigate the morphologies that might result from similar suspensions
of larger particles. To do this, it is assumed that other properties of the droplet remain
constant in particular, the solids diffusion coefficient is unchanged. This is likely to be
un-physical, but does not prevent qualitative conclusions being drawn.
The pressure profiles during the thickening regime for a droplet containing 500 nm particles are shown in Figure 4.23. Comparison with Figure 4.20 immediately demonstrates
some major differences. Firstly, the pressures involved are considerably lower; the maximum
pressure drop across the surface menisci for a droplet with 500 nm particles is 1.3 MPa com-

145

CSH

4.6. Applications

Solids Volume Fraction / []

0.9
Final
Particle Size

0.8
0.7

t=24.4 s

0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Radial Position / mm

Figure 4.22: Simulated solids volume fraction profiles during the drying of a colloidal silica droplet
containing 16 nm particles drying in air at Tgas = 178C. Profiles are plotted at 5 s intervals, with the
t = 24.4 s profile at the point of shell formation highlighted in bold. The external radius of the droplet is
shown on each profile as a small dot; a larger dot represents the final particle size. The diamonds on the
later profiles show the position of the evaporative front during the dry shell regime.

pared with the figure of 40 MPa for the previous droplet. Further, the pressure drop across
the shell is now seen to be less than that required for buckling. However, thickening continues as Pbuck is itself initially less than the maximum pressure drop across the air-liquid
interface. The interesting line on this plot is that indicating the pressure in the droplet
core, i.e., inside the thickening shell. This is seen to fall during the thickening regime with
negative values indicating a negative absolute pressure within the droplet.
Reference to Figure 4.4 shows that the thickening regime can end if the core pressure
falls below a certain critical value, Pcrit . This is indicated on Figure 4.23, where it is seen that
this is indeed the manner in which shell thickening is halted for this droplet. The choice of
the critical value is far from clear and, essentially, this may be thought of as a free parameter.
In the present example, the value Pcrit = 0.7 MPa was chosen for reasons explained below.
Figure 4.24 shows the structural evolution of the hypothetical droplet containing 500 nm
silica particles. It is instructive to compare this with Figure 4.21, the analogous plot for the
droplet containing 16 nm particles. The droplet with larger particles is seen to undergo a
period of wet shell drying and, consequently, the resulting dried-particle is hollow. This is
clearly seen in Figure 4.25, which shows the simulated evolution of the solids volume fraction within the droplet containing 500 nm particles. The central bubble is clearly visible
and close comparison with Figure 4.22 shows that a larger dried-particle is predicted.

146

CSH

4.6. Applications

1.5

Pressure / MPa

End of Thickening Regime, t = 32 s

Maximum Pressure = 1.3 MPa

Buckling Pressure
Pressure Drop Across the Shell
Core Pressure

0.5

0.5
Critical Pressure = 0.7 MPa
1
24

25

26

27

28

29

30

31

32

33

Time / s

Figure 4.23: Plot showing various pressures relating to a droplet of colloidal silica drying in air at
Tgas = 178C during the shell thickening regime. The suspended silica particles have a diameter of
500 nm.

Thickening

0.9

Wet Shell

0.8

Radial Distance / mm

Dry Shell
Boiling

0.7
0.6
0.5
0.4
0.3

Wet
Droplet Core

Wet
Shell

Dried
Particle

0.2

Bubble

0.1
0
0

10

20

30

40

50

60

Time / s

Figure 4.24: Simulated morphological evolution of a droplet of colloidal silica containing 500 nm particles drying in air at 178C.

147

CSH

4.6. Applications

Solids Volume Fraction / []

0.9
Final
Particle Size

0.8
0.7

t=24.4 s

0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Radial Position / mm

Figure 4.25: Simulated solids volume fraction profiles during the drying of a colloidal silica droplet
containing 500 nm particles drying in air at Tgas = 178C. Profiles are plotted at 5 s intervals, with the
t = 24.4 s profile at the point of shell formation highlighted in bold. The external radius of the droplet is
shown on each profile as a small dot; a larger dot represents the final particle size. The diamonds on the
later profiles show the position of the evaporative front during the dry shell regime. The solids profile in
the dried-particle is shown by the heavy dashed line.

Clearly the size of the suspended particles influences the drying modes and, consequently, the morphology of the final dried-particle. Figure 4.26 illustrates this link more
clearly, combining the results of many simulations to produce a map showing the drying
routes of droplets containing various sizes of suspended particles. For droplets containing
small particles like the first droplet simulated in this section solid dried-particles result with no wet shell drying period. Droplets containing solid particles between 50 and
1000 nm in diameter pass through a wet shell region in their drying histories and, as a result, produce hollow dried-particles. However, for droplets with suspended particles larger
than 1 m, the wet shell region is again bypassed and solid dried-particles are produced.
The existence of a transition from hollow back to solid dried-particles as suspended particle size increases is the key to choosing the critical pressure that determines the start of the
wet shell drying regime, Pcrit . Walton and Mumford (1999a) observed a range of colloidal
silica droplets drying and reported that solid dried-particles tended to be formed when the
suspended solids were larger than 1 m in size, with hollow dried-particles resulting from
droplets containing smaller suspended solids. Baring this result in mind, Pcrit was set equal
to 0.7 MPa in these simulations to ensure the transition from hollow to solid particles
occurs at the correct suspended particle size.
Figure 4.27 shows the pressure profiles in the thickening regime for a drying droplet
148

CSH

4.6. Applications

44

Dried Particle

42
40

Boiling

Time / s

38
36
34
32

Dry Shell

Wet Shell

30

Thickening

28
26
24
0

500

1000

1500

2000

Silica Particle Size / nm

Figure 4.26: Schematic showing the effect silica particle size on the predicted drying process.

containing suspended particles 1500 nm in diameter. Such droplets are predicted to form
solid dried-particles that is, they bypass the wet shell drying period. The reason for
this is clear from the pressure profiles. With the pressure drop across the shell less than
the pressure required for buckling, it is the later which determines the pressure drop across
the airliquid interface, (4.2.1). For this droplet, the constraint that this pressure must
be less than Pmax is seen to be binding, i.e., the buckling pressure exceeds the maximum
pressure capable of being supported by the surface menisci before the core pressure falls
below the critical value, Pcrit . Reference to Figure 4.4 shows that this sequencing results in
direct transition to the dry shell drying regime.
Effect of the Youngs Modulus
The Youngs modulus of the shell used to produce all the figures in this section was E =
1 MPa. However, as mentioned above, the films measured by Smith et al. (1995) to obtain
this figure were composed of 540 nm particles considerably larger than the 16 nm particles suspended in the droplets simulated thus far. The sensitivity of the results to the choice
of E was therefore studied using a droplet containing 500 nm particles which, as discussed
above, is predicted to form a hollow dried-particle.
The droplet mass and temperature profiles were found to be relatively insensitive to the
choice of Youngs modulus, with the timing of the wet shell to dry shell switch varying by
less than 1 s as E was varied over three orders of magnitude. However, as is clear from
Figure 4.28, the magnitude of the Youngs modulus is of fundamental importance in deter-

149

CSH

4.6. Applications

1.5

Pressure / MPa

End of Thickening Regime, t = 31.2 s

Buckling Pressure
Pressure Drop Across the Shell
Core Pressure

Maximum Pressure = 0.43 MPa

0.5

0.5
Critical Pressure = 0.7 MPa
1
24

25

26

27

28

29

30

31

32

Time / s

Figure 4.27: Plot showing various pressures relating to a droplet of colloidal silica drying in air at
Tgas = 178C during the shell thickening regime. The suspended silica particles have a diameter of
1500 nm.

mining the dimensions of the final dried-particle. Droplets which form shells with a larger
Youngs Modulus produce larger particles with narrower shells, as might be expected.
Conclusions
The new droplet drying model is once again used to simulate a droplet of colloidal silica,
but now the focus is on drying following shell formation. Droplet mass and temperature
histories are produced and these are in good agreement with experimentally measured values. In producing these predictions, use is made of all the key sub-models introduced in
this chapter.
When running the simulations presented in this section, no particular drying route is
specified in advance. Rather, the choice of sub-model used at each stage in the drying
history is determined on the fly using the criteria developed in Section 4.2.2. This allows
the influence of drying conditions, droplet composition and structural properties of the shell
on evolving particle structure to be investigated. For example, it is demonstrated that the
size of the suspended particles influences whether solid or hollow dried-particles are formed
and that the Youngs modulus of the growing shell has a major influence on dried-particle
shell thickness. For the first time it has been demonstrated how a single mechanistic droplet
drying model is capable of simulating multiple dried-particle morphologies.

150

CSH

4.7. Conclusions of the Chapter

0.8

Radial Possition / mm

0.7

Bubble Radius
Outer Radius

0.6
0.5
0.4
0.3
0.2
0.1 1
10

10

10

10

Youngs Modulus / MPa

Figure 4.28: Plot showing how the Youngs modulus of the shell affects the size and shell thickness of a
dried-particle. The data refers to a droplet of colloidal silica with initial radius, R0 = 0.95 mm containing
suspended particles with diameter L = 500 nm. The droplet was dried in air at Tgas = 178C.

4.7 Conclusions of the Chapter


The aim of this chapter was to extend the core droplet drying framework developed in
Chapter 3 to produce a model capable of simulating droplets following formation of a
surface shell. This was achieved by augmenting the core model with a number of submodels, each describing different structural developments following shell formation.
The choice of sub-models included was motivated by a review of experimentally observed dried-particle morphologies. Where appropriate, existing sub-models from the literature the wet and dry shell models were formulated for incorporation within the
new framework. Discussion of the physics of drying following shell formation allowed these
existing models to be related to real physical processes and motivated the development of
the new shell thickening and slow boiling sub-models. Most importantly, understanding
how all the sub-models relate to real physical processes occurring during drying allowed the
formulation of criteria for deciding which sub-model to use at any point throughout the
drying history. Taken together, this enables for the first time simulation of multiple
dried-particle morphologies by a single mechanistic droplet drying model without requiring
explicit a priori specification of the drying route.
The later sections of this chapter dealt with model validation and demonstrating the features of the new, extended model. A droplet of detergent slurry was simulated by successfully
simplifying the complex multi-component system and applying the dry shell model. The

151

CSH

4.7. Conclusions of the Chapter


predictions were compared with NMR observations which meant that, for the first time,
a mechanistic model has been validated against spatially resolved data. These comparisons
raise some questions about the applicability of the dry shell model in this case, but the general qualitative fit between predictions and experiments is good. Predicted solids profiles
suggest dried-particle morphologies similar to those observed in SEM images and simulated
solute concentrations agree with those obtained from NMR data. The new droplet drying model has therefore been shown to yield useful results for a system of great industrial
interest.
Simulations of a colloidal silica droplet utilise all features of the new droplet drying
model and demonstrate the ability of the extended model to simulate multiple dried-particle
morphologies. Temperature and mass histories are compared with experimentally measured
values, with which they show good agreement. Parameter studies are used to illustrate how
the structural evolution of the droplet is influenced by varying the initial droplet composition, the structural properties of the shell or the drying air conditions. Such studies are only
possible as a result of the physically motivated criteria developed for choosing the appropriate sub-model at each point in the drying history.
In summary, this chapter contained the following novel elements:
the shell thickening sub-model was developed;
the wet and dry shell sub-models were extended, providing physical motivation and
demonstrating compatibility with the new model framework;
physically motivated criteria to determine the appropriate sub-model to use following shell formation were developed, thus allowing simulation of structural evolution
influenced by droplet composition and drying conditions;
the extended droplet drying model was used to simulate systems of practical interest
and validated against experimental data.
The simulations presented in this chapter demonstrate the ability of the new model to
simulate complex systems of commercial interest and, for the first time, demonstrate a
method of simulating multiple dried-particle morphologies based on droplet composition
and drying conditions.

152

CSH

Chapter 5
Solution Methodology
In which the method employed to solve the droplet drying model developed in the
previous two chapters is discussed. Methods to solve systems of advectiondiffusion
reaction equations are reviewed, along with approaches for handling moving boundary problems. The model equations are stated in full along with all boundary conditions and expressions giving the evolution of the solution domains. These equations
are subsequently reformulated for solution using the NAG D03PLF routine. The
chapter ends with a detailed discussion of the numerical issues encountered when
solving the equation system. In particular, the solution grid, numerical flux function, error tolerances and integration time step are investigated.

5.1 Background
The preceding chapters have presented the motivation for and derivation of a new model
for the drying of droplets containing dissolved or suspended solids. This model has been
used to simulate a number of different drying systems and these results have been compared
with experimental data from the literature. What remains to be discussed is the means by
which the model has been solved; this is the subject of the present chapter.

153

5.1. Background

5.1.1 Equation Classification


The droplet drying model consists of a set of coupled partial and ordinary differential equations. The partial differential equations are of the form25
P (u, x; t )

u (x; t ) +

F (u, x; t ) = C (u, x; t )
D u,
, x; t + S (u, x; t ) ,
| {z }

x
|
{z
} |
{z
} S T
C F

(5.1.1)
which is the convectiondiffusionreaction equation, (LeVeque, 2002). Physically, the
terms F , C D 0 and S represent convective flux, diffusion and source terms respectively. This
equation may be classified as hyperbolicparabolic: it combines the first-order hyperbolic
advection equation,
u
t

(5.1.2)

F (u, x; t ) = 0 ,

with the parabolic diffusion equation,


u
t

u
x

, x; t

(5.1.3)

The geometric classification of an equation that is, whether it is hyperbolic, parabolic or


elliptic has important implications for the numerical methods suitable for its solution.
For hyperbolic problems, the domain of dependence the set of points in spacetime
on which the solution, u (x; t ), depends is a bounded set, (Schatzman, 2002). This
arises from the fact that information can only travel at a finite speed in hyperbolic systems,
reflecting their underlying wave or advective nature. As a consequence of this fact, explicit
schemes may often be used successfully to solve hyperbolic problems. In contrast, the
domain of dependence in parabolic systems is the entire real line, although distant points
may have little effect on the solution as their contributions die away exponentially fast.
Nevertheless, this means that implicit methods are often required for parabolic equations.
The present problem has assumed spherical symmetry, thereby reducing all partial differential equations in the model to a single spatial dimension. The general approach to
solving such 1-D problems is to subdivide the spatial domain into intervals. These intervals are marked by grid points in finite difference approaches or are termed grid cells in
finite volume formulations. The finite difference and finite volume approaches are closely
25

Note that the u dependence of D permits advective that is, hyperbolic contributions to the
diffusive term. This form is used here in anticipation of the structure permitted by the NAG D03PLF
routine, described in Section 5.3.

154

CSH

5.1. Background
related; it is often possible to directly interpret a finite volume discretisation as a finite difference approximation, (Chung, 2002). However, finite volume methods are derived from
the integral form of the conservation law which turns out to have several advantages.

5.1.2 Solution Methods


This subsection illustrates the construction of numerical schemes for the solution of partial differential equations using the finite volume approach. The material presented is the
minimum necessary to understand the method used to solve the droplet drying model; the
reader is referred to the many standard texts for more details, (e.g., LeVeque, 2002; Hirsch,
1992a,b).
Consider solving the conservative system of 1-D equations,

u (x; t ) +

u,

u
x

, x; t

=0 ,

(5.1.4)

where f represents the flux of u. The advection and diffusion equations, (5.1.2) and (5.1.3)
respectively, are seen as special cases of this general equation. Centring each cell in the
solution grid on an associated grid point and denoting these


Ci = xi 1 , xi + 1
2

(5.1.5)

as illustrated in Figure 5.1, allows Ui (t ) to be defined as an approximation to the average


value of u over Ci at time t ,
Ui (t )

xi + 12

u(x, t )d x

xi 12

1
x

Z
u(x, t )d x ,

(5.1.6)

Ci

where x = xi + 1 xi 1 is the width of the cell.26 The integral conservation law states that
2
2
Ui can only change as a result of fluxes across the cell boundaries, i.e.,
dUi
dt

1 d
x dt


 
i
1 h  
(5.1.7a)
f u xi 1 , t f u xi + 1 , t
2
2
x

Fi 1 Fi+ 1 .
(5.1.7b)
2
2
x

u(x, tn ) d x =
Ci

The terms on the right, Fi 1 , are called flux vectors and represent the boundary fluxes at
2

26

Here it is assumed that the grid cells are all of the same size, i.e., x is a constant. However, this is
for clarity of presentation rather than a requirement of the method. When using uneven spacings, it is
important to specify whether the grid points or grid cells are defined first, (Versteeg and Malalasekera,
2007). Whenever such uneven grids are used in this work, the grid points are considered primitive, with
cell boundaries being placed equidistant from adjacent nodes.

155

CSH

5.1. Background
x
xi 1

Ci
xi 1

xi

xi+ 1

xi +1

Figure 5.1: Grid points, cells and cell boundaries on a one dimensional domain with uniform grid
spacing.

x = xi 1 . Information travels at a finite speed in hyperbolic problems and, as a result, it is


2

reasonable to suppose that these boundary fluxes may be approximated using only the cell
averages on either side of the interface, i.e.,

Fi 1 = F Ui 1 , Ui ,

(5.1.8)

where F is some numerical flux function. The general finite volume method is therefore
dUi
dt

1
x

F Ui , Ui +1 F Ui 1 , Ui
.

(5.1.9)

The specific method depends on the choice of formula for F but, in general, any method
of this type exploits a three-point stencil. That is, the evolution of Ui depends on the
three values Ui1 , Ui and Ui +1 . For parabolic problems, restricting the functional form of
the flux function in this way corresponds to the use of finite difference methods similarly
characterised by stencils with, at most, three spatial points at a single time.
An appropriate choice of numerical flux function is essential for accuracy and stability
of the solution algorithm. The simplest way of approximating the flux is via an arithmetic
average,


f Ui 1 + f Ui
,
Fi 1 = F Ui 1 , Ui =
2
2

(5.1.10)

but, in general, this approach is unstable for hyperbolic problems, (LeVeque, 2002). For
such systems, upwind methods give better results. As an example, consider the simple
advection equation
u
t

+v

u
x

(5.1.11)

=0 ,

156

CSH

5.1. Background
where the upwind numerical flux function is given by

 v Ui1
Fi 1 = F Ui 1 , Ui =
2
v U
i

v >0

(5.1.12)

v <0 .

The direction of upwinding is chosen to reflect the flow of information in the underlying
partial differential equation.
Systems of differential equations require more complicated flux functions which, for
hyperbolic systems, are generally based on upwinding along the characteristic directions.
Second-order accurate methods, e.g., Lax-Wendroff or Beam-Warming, add an additional
diffusive term to the simple, first-order numerical flux functions. Such methods give better accuracy for smooth solutions but introduce oscillations near discontinuities and with
certain boundary conditions. High resolution methods using slope- or flux-limiters seek
second-order accuracy whilst avoiding the oscillations introduced by pure second-order
methods. There exists a large body of literature covering the relative merits of these methods, (e.g., Roe, 1981; LeVeque and Yee, 1990; LeVeque, 2002; Hirsch, 1992a,b; Chung,
2002), and the reader is referred to these works for more detail. The form of the numerical
flux function used in the present work is discussed in Section 5.5.
The finite volume method represented by (5.1.9) is an example of the method of lines
for solving systems of partial differential equations, (Ames, 1977). This is a semi-discrete
approach in which the partial differential equation system is discretised in all but one of the
independent variables, yielding a coupled system of ordinary differential equations. Fully
discrete methods, in contrast, discretise all the independent variables, including time. A
simple time-marching algorithm is used to obtain Uin+1 an approximation to the solution
in Ci at time tn+1 from values at the current time, tn . The finite volume method written
for such an approach reads
Uin+1 = Uin

t
x

n
Fin+ 1 Fi
1
2

(5.1.13)

n
where the flux vectors, Fi
1 are approximations to the fluxes across x = x i 1 averaged over
2

the time step, t . Here the Euler method has been used for the time integration, yielding
an explicit method, (Hirsch, 1992a); time discretisation using the backward or implicit
Euler method would give an implicit scheme with

n+1
, Uin+1 ,
Fin 1 = F Uin1 , Uin , Ui1

(5.1.14)

where the numerical flux function is now dependent on the solution values at two time steps.
As discussed above, such implicit approaches are often required for parabolic problems.

157

CSH

5.1. Background
It is clear that (5.1.13) may be viewed as a direct finite-difference approximation to the
conservation law, (5.1.4), since re-arrangement gives
Uin+1

Uin

n
F Uin , Uin+1 F Ui1
, Uin
x

=0 .

(5.1.15)

The method of lines clearly reduces to the fully discrete approach if the Euler method is
used to integrate the resulting system of coupled ordinary differential equations. However,
the advantage of the method of lines lies in the flexibility afforded to take advantage of more
sophisticated methods for solving such systems. Cash (2003) reviews a number of wellestablished software packages for numerically integrating systems of ordinary differential
equations. The method of lines is therefore the natural method for Cauchy problems
that is, transient problems where a specified initial state is evolved in time,27 (Ames, 1977).

5.1.3 Moving Boundary Problems


Methods for Dealing with Moving Boundaries
The present problem is one that involves moving boundaries. During the first drying period
the droplet physically shrinks as moisture is removed. The situation gets further complicated once a shell has formed. As discussed in Chapter 4, the partial differential equations
describing behaviour in the shell region may be different from those in the adjoining core.
It therefore makes sense to define two adjoining solution domains representing the inner
and shell regions, as illustrated in Figure 5.2. There may then be three boundaries moving
simultaneously: the external radius, R, might continue to decrease, e.g., during the thickening regime, or may increase due to puffing; the inner shell radius, S marking the
boundary between the two domains will move as the shell grows; and a centrally located
bubble may alter the position of the left hand end of the inner domain. It is therefore necessary to give consideration both to the manner in which this moving boundary problem is
handled and to the way in which these two domains are coupled.
There are a number of different methods for dealing with free and moving boundaries
and the reader is referred to Crank (1996) for a detailed discussion. Whilst analytical solutions are available for some specific cases, they are of little practical help beyond use as
reference solutions for testing numerical approaches.
Three principle numerical approaches to solving moving boundary problems may be
identified. It is possible to solve the equations on a fixed grid. This avoids complications
associated with evolving the solution grid in time, but has the disadvantage that the boundary may lie between grid points at some time steps. It is then necessary to interpolate the
solution near the boundary, (e.g., Crank, 1957) or adapt the integration time step, t so
27

This type of problem is sometimes termed a propagation problem, (Ames, 1977).

158

CSH

5.1. Background
Central
Bubble

Inner
Domain

b (t ) S (t )

R (t )

Outer
Domain

Figure 5.2: Following shell formation, the drying droplet may be divided into two adjoining solution
domains, both of whose extents vary with time.

that the boundary always lies at a grid point, (e.g., Douglas and Gallie, 1955). A further
problem is the loss of resolution as the droplet shrinks the number of grid points in the
solution domain inevitably decreases under such circumstances.
The second approach is to transform the problem into a new co-ordinate system in
which the boundary and the solution grid are fixed, (Crank, 1957, 1996). Applied to
droplet drying problems, this idea is the basis of the solids- and solute-fixed methods introduced by Van der Lijn et al. (1972) and since used successfully to solve several drying
models, (e.g., Hecht and King, 2000b; Adhikari et al., 2007). The computational grid
points remain fixed at uniform intervals across the space of solids or solute mass. This
fixed grid reduces the computational resources required and the transformation means no
interpolation is required to determine the position of the moving interfaces. Further, whilst
evenly distributed in the space of solids or solute mass, the grid points are physically closer
together near the outer edge of the droplet. This coincides with the location of the steepest
gradients in the evolving profiles and so the approach helps give better resolution in these
key areas. Despite these attractive features, the solids-fixed approach is not suitable for the
present application where the mass of solid material present may change as a result of solute
crystallisation.
The final approach, developed by Murray and Landis (1959), is to allow the solution
grid to move and rewrite the partial differential equations in terms of the total derivative following the moving nodes. Sometimes called the variable space network, (VSN) method,
this approach has been used by many researchers to model droplet drying, (e.g., Farid, 2003;
Shabde et al., 2005; Seydel et al., 2006). The total number of mesh points is kept constant
and the time derivatives following a grid point at r = Ri (t ) are related to those at constant

159

CSH

5.1. Background

(t ) R1 (t ) R2 (t )
b (t ) R0

R3 (t ) R4 (t )

R5 (t ) R6 (t )

R (t )

r
Figure 5.3: A spacetime illustration of a uniformly expanding solution domain showing the positions of
the grid points, Ri (t ). The particles shown remain at a single location in physical space. To satisfy this
condition, a virtual flux must be introduced to the equations written following the moving grid.

r by



t



dRi
=
.

t r
dt Ri
Ri (t )
| {z }

(5.1.16)

V F

The relationship between the moving grid and stationary physical frame is illustrated in
Figure 5.3, which shows an expanding solution domain, spanned at all times by six spatial
intervals. Whilst the particles shown are at rest in physical space, it is clear that they are in
motion from the perspective of the moving grid. This explains the virtual flux term seen
in (5.1.16).
The VSN method for handling moving boundary problems has the advantage that the
boundaries of the domain lie on a grid node at all times. The numerical implementation is
further facilitated by the fact that the number of grid points remains constant but, unlike
the solids-fixed approach, the location of the grid points is unconstrained.28 The primary
disadvantage of this method in general is the introduction of the virtual flux. For problems
which are initially purely parabolic, the introduction of a hyperbolic term can mean that an
entirely different possibly considerably less efficient solution method is required. As
28

Whilst Figure 5.3 suggests the grid points are equally spaced, this is not a requirement of the method.
Indeed, as discussed in Section 5.6, it is found that a non-uniform grid is required to resolve the evolving
profiles adjacent to the outer droplet boundary.

160

CSH

5.1. Background
is seen below, this is an issue for the system of equations comprising the new droplet drying
model. However, the added difficulty is not severe and the problems relating to the virtual
flux term are found to be manageable. For all these reasons, this is the approach used to
solve the present problem.
Application of the Moving Grid Approach
The moving grid approach is used to handle the moving boundary conditions in both of
the domains inner and shell introduced in the previous section. To use the method
it is necessary to know how these domains evolve in time, i.e., an expression is needed for
the dRi /dt in (5.1.16). It is assumed that, whilst the initial choice of node locations is free,
the relative positions of these grid points in the evolving co-ordinate system do not change
with time. That is, a grid point initially located in the centre of the domain will remain at
the mid-point for all time although, of course, the physical location of the point and the
entire domain may vary freely. This condition is expressed by the transformation
z=

r rL
rR rL

rR rL z

(5.1.17)

which is applied to all the model equations. In (5.1.17), [rL , rR ] = [b , S] for the internal
co-ordinate system and [rL , rR ] = [S, R] for the shell region. This transformation has the
effect of non-dimensionalising the spatial co-ordinate and fixing both internal and shell
domains on the interval z [0, 1], as illustrated in Figure 5.4. Having both the inner
and shell domains represented by the same interval in the transformed zspace allows all
the partial differential equations in the model to be solved on the same grid, considerably
simplifying the numerical implementation.
Time derivatives in the z co-ordinate systems are related to those in the r co-ordinate
system by



1
drL
drR drL

+z

,
=

t z t r rR rL dt
dt
dt
z

(5.1.18)

which, comparing with (5.1.16), shows that the grid points move in physical space according to
dR
dt

1
rR rL

drL
dt

+z

drR
dt

drL
dt

(5.1.19)

A full listing of the model equations following application of this transformation is given in
the following section.

161

CSH

5.2. Model Equations and Boundary Conditions

r
S (t )

b (t )

R (t )

Figure 5.4: Illustration of the co-ordinate transformation applied to a drying droplet with a shell region.
After transforming, the core and shell regions are both fixed on the interval z [0, 1].

5.2 Model Equations and Boundary Conditions


This section gives the complete model formulation. The equations are presented in the
form in which they are solved, i.e., following the co-ordinate transformation discussed in
Section 5.1.3. In addition, normalised moments are introduced,
i (z; t ) =
m

mi (z; t )
(0)

mi

(5.2.1)



where mi(0) = max mi (z; t = 0) , the maximum initial value of mi . All the droplets simulated in this work are initially well mixed meaning mi(0) is simply the initial, spatially uniform
moment value. The moment equations are rewritten in terms of these new variables.
For each drying regime, the number of spatial domains considered is stated, followed
by the equations solved in each. The boundary conditions are then listed and, finally, the
ordinary differential equations describing the evolution of the spatial domains are given.

5.2.1 Prior to Shell Formation


Before the formation of a shell, there is one co-ordinate system spanning the entire physical
domain, [b , R]. This domain is mapped on to the interval [0, 1] through application of the
transform discussed in Section 5.1.3.

162

CSH

5.2. Model Equations and Boundary Conditions


The continuous phase equation to be solved is


1
(1 ") (c) B (c) +
t
(rR rL ) [rL + z (rR rL )]2

Deff
2
(c)
(c)
(c)
(c)
[rL + z (rR rL )]
(1 ") v r B
(1 ") B

z
rR rL z

drL
drR drL

"
1

(1 ") (c) B (c) + 0D
+z

=0 .
rR rL dt
dt
dt
z
t crys

(5.2.2)
where, rL = b and rR = R when applied prior to shell formation. The mass-averaged
continuous phase velocity is given by
v r (c) =

Deff

B (c)

rR rL

B + B (c)

"

1" z

!
,

(5.2.3)

where B is a dimensionless ratio of material densities,


B =

0B
0A 0B

(5.2.4)

and the continuous phase density, (c) , is


(c)

0AB
B (c) + B

(5.2.5)

The solids sink term in (5.2.2), i.e., the rate of solute crystallisation, is obtained from the
evolution equation for the solids volume fraction. In full, this equation reads
"

(0)
2
= L3min N0 + Gm2 m
t 6
2
!
[rL + z (rR rL )]2 "

1
+
D
rR rL
z
(rR rL ) [rL + z (rR rL )]2 z

drR drL
1
drL
"
+z

,
+
rR rL dt
dt
dt
z

(5.2.6)

and the first two terms on the right hand side represent the contribution made by crystallisation. The evolution of the other moments of the solid particle number density is given by

163

CSH

5.2. Model Equations and Boundary Conditions

a
m
t

Lamin N0

(0)

+ aG

ma(0)
+

ma1
ma(0)

a1
m

[rL + z (rR rL )]2

rR rL
(rR rL ) [rL + z (rR rL )] z

a
1
drL
drR drL
m
+
+z

,
rR rL dt
dt
dt
z

a
m

(5.2.7)

a {0, 1, 2} .

Zero gradient boundary conditions are applied to all variables at z = 0. At z = 1, the


continuous phase boundary condition is

B (c)

z

= (rR rL )
z=1

00
B (c) m
(c) Deff

(5.2.8)

and the boundary conditions for the moments and solids volume fraction are

a
00
m
m
a ,
m
= (rR rL )
z z=1
D0A

00
"
m
and
".
= (rR rL )
z
D 0

a {0, 1, 2}

(5.2.9)
(5.2.10)

z=1

where, as before, rL = b (t ) and rR = R (t ). The extent of the domain changes according to


db
and

dt
dR
dt

(5.2.11)

=0 ,
=

00
m
0A

(5.2.12)

where it is assumed that there is no bubble growth prior to shell formation.

5.2.2 Thickening Regime


In the thickening regime, the physical domain is divided into two regions such that [rL , rR ] =
[b , S] for the internal co-ordinate system and [rL , rR ] = [S, R] for the shell region. Both of
these domains are mapped on to the interval [0, 1] when the equations are transformed into
the z co-ordinate system, (5.1.3). The solute mass fraction, solids volume fraction and
a, shell respectively.
normalised moments in the shell region are denoted B (c)
, "shell and m
shell
Having made the appropriate substitutions for the shell region variables, (5.2.2) is applied
unaltered in both regions. Equations (5.2.6) and (5.2.7) are unaltered in the core, but in

164

CSH

5.2. Model Equations and Boundary Conditions


the shell the appropriate equations are now
a, shell
m
t

(0)

=aG

ma1, shell

R2

a1, shell +
m

a, shell
dR m

(R S) [S + z (R S)]2 dt

a, shell
m
1
dS
dR dS
+
+z

,
R S dt
dt
dt
z

(0)

ma, shell

(5.2.13)

a {0, 1, 2}

and
"shell
t

(0)
2, shell
Gm2, shell m

1
RS

R2

dR "shell

(R S) [S + z (R S)]2 dt z

dS
dR dS
"shell
+z

.
dt
dt
dt
z

(5.2.14)

The boundary conditions at the centre of the droplet are, again, those of zero gradient
for all the variables. For the internal co-ordinate system, the boundary conditions at z = 1
are

(c)
+
(c)
Deff 1 "shell | z=0 B shell
B

=

z z=1 Deff
1"
z

z=0
(5.2.15)

0
(c)
AB
" "shell | z=0 dS
(S b )
.
1"
dt
Deff (c)
+

represent the effective diffusion coefficient for


for the continuous phase, where Deff
and Deff
+

the continuous phase calculated at S and S respectively. The boundary conditions on the
continuous phase equation in the shell region are

(c)
B shell
z=0


= B
(c)

(5.2.16)

z=1

at z = 0 and

(c)
B shell


z

(c)

= (R S)
z=1

00
B shell m
(c) Deff

(5.2.17)

at the external droplet boundary.


The discrete phase boundary conditions for the inner region at z = 1 are

a
a
1 ksink m
m
,
=
z z=1
Sb D

a {0, 1, 2} ,

(5.2.18)

165

CSH

5.2. Model Equations and Boundary Conditions


for the moments, and

"
1 ksink "
,
=
z z=1
Sb D

(5.2.19)

for the solids volume fraction. Zero gradient boundary conditions are applied at both ends
of the domain for the discrete phase equations in the shell region.
The bubble, shell and external interfaces evolve according to the equations
db

(5.2.20)

=0 ,

dS
=
dt m

dt

and


 
0, shell
m
R 2 dR
z=1


,
S

dt

m

0 z=1
0, shell

dR

dt

00
m
0A

m0, shell

z=0

0 | z=1 , (5.2.21)
6= m

z=0

(5.2.22)

5.2.3 Wet Shell Regime


In the wet shell regime, there are again two co-ordinate systems representing the core and
shell regions. Variables in the shell region are notated in the way described in the previous
section. Having made the appropriate substitutions for the shell region variables, Equation (5.2.2) can be applied in both regions, with the mass-averaged radial velocity given by
v r (c) =

Deff

B (c)

rR rL

B + B (c)

"

1" r

b 2 db
r 2 dt

(5.2.23)

In the inner region, the moment equations read


a
m
t

Lamin N0
ma(0)
+

(0)

+ aG

ma1
ma(0)

a1
m

[b + z (S b )]2

(S b ) [b + z (S b )]2 z

b2

Sb

a
m

(5.2.24)

a
db m

(S b ) [b + z (S b )]2 dt z

a
1
dS db
db
m
+
+z

,
S b dt
dt
dt
z

166

a {0, 1, 2}

CSH

5.2. Model Equations and Boundary Conditions


and the solids volume fraction, ", is given by
"

(0)
2
= L3min N0 + Gm2 m
t 6
2
!
[b + z (S b )]2 "

1
D
+
Sb
z
(S b ) [b + z (S b )]2 z
2
b
db "

(S b ) [b + z (S b )]2 dt z

1
db
dS db
"
+
+z

.
S b dt
dt
dt
z

(5.2.25)

In the shell region, the corresponding equations are


a, shell
m
t

(0)

=aG

ma1
ma(0)

ma1,shell

1
RS

dS
dt

+z

dR
dt

dS

a, shell
m

dt

(5.2.26)
a {0, 1, 2}

and
"shell
t

Gm2,shell +

1
RS

dS
dt

+z

dR
dt

dS
dt

"shell
z

(5.2.27)

Zero gradient conditions are imposed at the bubble interface for all variables in the inner
region. At z = 1, for the inner region, the continuous phase boundary condition is

B (c)

z

=
z=1

+
Deff

Deff


(c)
B shell


1"
z

z=0

(5.2.28)
0AB (c) " "shell | z=0
dS b 2 db
(S b )
2
.
1"
dt
S dt
Deff (c)

1 "shell | z=0

The boundary conditions on the continuous phase equation in the shell region are given
by (5.2.16) at z = 0 and (5.2.17) at z = 1.
The boundary conditions on the moments in the central region at z = 1 are

a
m

z

=
z=1

a
ksink m

Sb

a {0, 1, 2} ,

(5.2.29)

167

CSH

5.2. Model Equations and Boundary Conditions


and

"

z

=
z=1

ksink "

Sb D

(5.2.30)

for the solids volume fraction. For the shell region, the boundary conditions for the moments at z = 0 are

1
2

ma(0)
dS
b db

, a {0, 1, 2} , (5.2.31)
a, shell =
a | z=1 1
m
+ ksink
m
2
(0)
z=0
dt
S dt
ma, shell
and
(5.2.32)

"shell | z=0 = "crit .

for the solids volume fraction. Zero gradient boundary conditions are applied to all the
moments at the outer boundary.
The bubble, shell and external interfaces evolve according to the equations
db
dt
dS

and

dt
dR
dt

00
R2 m
b 2 0A

(5.2.33)

b >0 ,

"| z=1
"shell | z=0 "| z=1

b 2 db
S 2 dt

+ ksink

(5.2.34)
(5.2.35)

=0 .

5.2.4 Dry Shell Regime


In the dry shell regime, the drying droplet is split into a core and shell region, but spatial
evolution equations are only solved in the core region spanning [b , S]. The continuous
and discrete phase equations in this region are given by (5.2.2), (5.2.7) and (5.2.6). Zero
gradient boundary conditions are applied to all variables at both ends of the domain, except
for the continuous phase equation at z = 1. For this variable, the boundary condition is

B (c)

z

= (S b )
z=1

00
B (c) m
(c) Deff 1 "| z=1

.

(5.2.36)

The bubble and external particle radii do not change in the dry shell regime. The position
of the internal shell interface is described by
dS
dt

 2
R

00
m

0A 1 "|S

(5.2.37)

.

168

CSH

5.3. Solving the Model Equations


Droplet drying in the presence of a dry shell does not require a second co-ordinate
system. The dry shell is free of the continuous phase and the particle number density may
no longer evolve in either its internal or external co-ordinates. Rather, following formation
of a dry shell at t = t DS , the spatial location of the shell interface is stored at regular time
intervals, t . This information is held in an expanding vector,




T
Sstore = S t = t DS , S t = t DS + t , S t = t DS + 2t , . . . .

(5.2.38)

The interfacial moments of the solids number density are stored at each time step in similarly expanding vectors; for example, the solids volume fraction at the inner edge of the
growing dry shell is stored in




T
store = " r = S t DS , " r = S t DS + t , " r = S t DS + 2t , . . . . (5.2.39)

At any given time, spline interpolation may then be used to reconstruct the spatially resolved
moment values within the dry shell region. Within the drying code, this spline interpolation
is performed using the NAG E02BEF and E02BBF routines,29 (NAG).
This completes the statement of the equations in the new droplet drying model. The
remainder of this chapter demonstrates how these equations are solved and investigates the
numerical behaviour of the solution.

5.3 Solving the Model Equations


Solving the new droplet drying model comes down to as shown in the preceding section integrating a coupled system of partial and ordinary differential equations in time.
For this purpose, the functions whose evolution is described by partial differential equations
are grouped together in a vector

T
u (z; t ) = uin , ushell

(5.3.1)

where uin contains those functions relating to the wet inner region of the drying droplet,

T
2, m
1, m
0
,
uin (z; t ) = B (c) , ", m

(5.3.2a)

and ushell contains the corresponding functions in the shell region,

T
(c)
2, shell , m
1, shell , m
0, shell
.
ushell (z; t ) = B shell , "shell , m
29

(5.3.2b)

Refer to Section 5.3.1 for more information regarding numerical routines from the Numerical Algorithms Group (NAG).

169

CSH

5.3. Solving the Model Equations


Prior to shell formation, the functions contained in ushell are not of interest they refer to
a non-existent domain and so (5.3.1) reduces to u = uin .
The functions contained in u evolve through time in the transformed spatial domain
z [0, 1]. The discretisation of this domain is described by the vector of grid points
T

z = z1 , z2 , . . . , znpts , where npts is the number of spatial grid points. The numerical
approximation to u (t ) at the point xi is denoted Ui (t ). Alternatively,

U (z; t ) = U1 , U2 , . . . , Unpts ,

(5.3.3)

may be used to denote the array of all spatial approximations to u at time, t . Similarly, the
variables whose evolution is described by an ordinary differential equation may be grouped
into the vector v (t ), with V (t ) denoting the numerical approximation to these values.
The droplet drying model is a 1-D Cauchy problem describing the time evolution of u
from an initial state and subject to boundary conditions. An introduction to the numerical
methods available for solving such problems was presented in Section 5.1.2; this section
describes specifically how the solution to the new droplet drying model is obtained.

5.3.1 The NAG Library Routine D03PLF


Within the structure of the code, the coupled partial and ordinary differential equations
outlined in Section 5.2 are solved using the D03PLF routine from the NAG Fortran Library,
Mark 20.30 This off-the-shelf routine was chosen as it is flexible, well documented and
capable of solving coupled systems of convectiondiffusionreaction equations such as those
constituting the present model.
The NAG D03PLF routine implements an algorithm developed by Pennington and
Berzins (1994). This algorithm uses the method of lines introduced in Section 5.1.2
to discretise the equation system in space, yielding a system of ordinary differential equations. The parabolic terms in the equation system are discretised automatically using a
centred differencing scheme, whilst the hyperbolic elements require a user-supplied numerical flux function, (5.1.2). This numerical flux function is augmented with a Van Leer
slope-limiter to produce a sophisticated high resolution scheme for handling the hyperbolic
aspects of the problem, (NAG; LeVeque and Yee, 1990; LeVeque, 2002). The form of the
numerical flux function used is discussed in Section 5.5.
The system of ordinary differential equations obtained from the spatial discretisation is
integrated in time using a backward differentiation formula (BDF) method, (NAG). BDF
methods are a type of implicit multistep scheme constructed directly from the differential
30

The Numerical Algorithms Group (NAG) began life in the early 1970s when researchers from various
UK universities began to develop a library of routines for numerical computation, (Phillips, 1986).
Since then, the library has grown to become the largest commercially available collection of numerical
algorithms, (NAG, retrieved 2008).

170

CSH

5.3. Solving the Model Equations


equations; the updating formulae are based on a polynomial constructed to satisfy the ordinary differential equations at t n+1 and to interpolate the values calculated at k previous
steps, (Schatzman, 2002). Coupled with Newton iteration for solving the linear system at
each step, BDF approaches are the integrators of choice for stiff systems. A full discussion
of such methods is given by Hairer and Wanner (1996).
The D03PLF routine is the core element of the program written to simulate droplet drying using the new model developed in this thesis. After initialising all appropriate variables,
the program passes U (z; t ) and V (t ) to the D03PLF routine. The routine integrates the
profiles over a pre-specified time step, t , returning the evolved values of these arrays at
t + t . The time step, t , is chosen to allow data to be written to file at regular intervals; under most circumstances, the choice of t should not affect the solution in any way.
Since no changes are made to U and V between successive calls to D03PLF, the routine can
normally be resumed rather than restarted at successive calls. This avoids computationally
costly re-initialisation of the numerical integration. The exceptions to this general rule are
at the points of regime change, e.g., the onset of the thickening regime or the switch from
the wet to dry shell models. At these points, the equations to be solved, boundary conditions and even the number of domains change and, consequently, the D03PLF routine must
be restarted.
The time step, t should not be confused with t , the time step used within the D03PLF
routine when performing the temporal integration. The later is an important numerical
parameter whose effect on the solutions obtained is investigated further in Section 5.6.3.
As illustrated in Figure 5.5, the D03PLF routine must be supplied with user-defined
subroutines specifying the partial differential equation system (pdedef), ordinary differential equation system (odedef), boundary conditions (bndary) and numerical flux function
(numflx). These in turn call upon functions to calculate physical properties of the droplet
and drying air listed in Table 5.1 and the moisture evaporation rate, (Algorithm 3.1).
The remainder of this section demonstrates how the model equations summarised in Section 5.2 are manipulated into the form required for these user supplied routines. The
implementation of and numerical issues relating to the boundary conditions and numerical
flux function are then discussed in Sections 5.4 and 5.5 respectively.

171

CSH

Equation

172

Temperature in degrees Celsius


Temperature in Kelvin
(1)
Rahman (1995)
(2)
Werner et al. (2008c)
(3)
White (2005)
(4)
Rogers and Mayhew (1994)

Viscosity / Pa s
Heat Capacity / J kmol1 K1
Thermal Conductivity / W m1 K1

Air

T
291.15
2

23

b = 4 104 T 2 + 2.38 10 T + 1004.5


C
p
= 1.97 104 T 0.858

= 18.27 106 T411.15


+120

Water

Viscosity / Pa s
= 2.58 106 T 3 + 5.8 104 T 2 4.72 102 T + 1.7584 103
Latent Heat / J kg1
H vap = 1.3 103 T 2 2.29618T + 2500 103
b = 8.76 103 T 2 0.6042T + 4190
Heat Capacity (liquid) / J kmol1 K1
C
p
b = 1.67 102 T 2 2.61 102 T + 1866.4
Heat Capacity (vapour) / J kmol1 K1
C
p

Property

Table 5.1: Expressions for the various physical properties of water and air used in droplet drying simulations.

(2)

(4)

(3)

(2)

(1)

(1)

(1)

Reference

5.3. Solving the Model Equations

CSH

5.3. Solving the Model Equations

D03PLF
U (z; t + t )
V (; t + t )

U (z; t )
V (; t )

USER
SUPPLIED
ROUTINES

pdedef

numflx

bndary

odedef

FUNCTIONS TO CALCULATE
PHYSICAL PROPERTIES AND
DROPLET EVAPORATION RATE

EXTERNAL
DATA

Figure 5.5: Illustration of the routines called by the time integrator, D03PLF.

5.3.2 Definition of the System of Partial Differential Equations


pdedef
The D03PLF routine is intended to solve systems of convectiondiffusionreaction equations
that can be written in the general conservative form
npde
X
j =1

Pi , j (x, u, v; t )

u j (x; t ) +

Ci (x, u, v; t )

Fi (x, u, v; t ) =

Di

x, u,

u
x

, v; t

+ Si

x, u, v,

dv
dt

;t

(5.3.4)
for i = 1, 2, . . . , npde and where u is the vector of functions whose evolution is described by
the partial differential equations. This equation has the same general form as (5.1.1) and
all the partial differential equations in the new model, with additional dependence permitted on v , the vector of variables described by the coupled ordinary differential equations.
The system of coupled ordinary differential equations is discussed further in Section 5.3.3.
The terms F , C Dx and S represent convective flux, diffusion and sources respectively
although, as is shown below, these do not correspond exactly with the advective and diffu 173

CSH

5.3. Solving the Model Equations


sive terms in the model equations. As indicated in (5.3.4), the functions P, F and C may
depend on x , t , u and v ; D may additionally depend on u x whilst S may have additional
linear dependence on the time derivative of v .
As discussed in Section 5.1.2, the hyperbolic portion of (5.3.4) that is, the F term
requires special consideration. This is handled using a user-supplied numerical flux function
in the numflx subroutine. The remaining terms that, together with F define the entire
system of partial differential equations, are computed by the pdedef subroutine. When
passed values of x , t , u, u x , v and v , pdedef evaluates the functions P, C , D and S . The
remainder of this section is devoted to manipulating the drying model equations into the
appropriate forms for use in this subroutine. An expression is also obtained for F , although
discussion of the numerical flux function used is deferred until Section 5.5.
P term

The coefficients of the time derivatives of u are grouped into the matrix P. This differs from
the identity matrix as a consequence of the solute equation, (5.2.2). Here it is necessary to
expand the time derivative, i.e.,

(c)
"

(1 ") (c) B (c) = (1 ")
B (c) (c) B (c)
,
t
t
t

(5.3.5)

which, on substituting for (c) from (3.3.23), gives


(c) 2 (1 ") B (c)
"

(c)
(c)
(c)
(c)
(1 ") B
=

.
B
t
t
t
0A

(5.3.6)

Equation (5.2.2) applies in both the wet core and shell region, when present, giving the
general form of P as

P=

a
0
0
0
0

b
1
0
0
0

0
0
1
0
0

0
0
0
1
0

0
0
0
0
1

0
0
0
0
0

0
0
0
0
0

0
0
0
0
0

0
0
0
0
0

0
0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

0
0
0
0

a
0
0
0

b
1
0
0

0
0
1
0

0
0
0
1

0
0
0
0

0 0 0 0 0 0 0 0 0 1

174

(5.3.7)

CSH

5.3. Solving the Model Equations


where
(c) 2 (1 ")

(5.3.8a)

b = (c) B (c) .

(5.3.8b)

a=

0A

and

Whilst there is only one co-ordinate system i.e., prior to shell formation or in the dry
shell regime only the upper left quadrant of this matrix is needed.
Terms associated with the Diffusive Flux

T
The diffusive coefficient term in (5.3.4) may be expressed C = Cin , Cshell , and allows
the solution of problems in non-Cartesian co-ordinates. The present problem is assumed to
have spherical symmetry for which,
Cin (U , z; t ) =

1
[b + z (S b )]

(1, 1, 1, 1, 1)T ,

(5.3.9a)

and
Cshell (U , z; t ) =

1
[S + z (R S)]

(1, 1, 1, 1, 1)T .

(5.3.9b)

Inspection of (5.2.2) shows that the remaining part of the diffusive term in the solute equation is
D1 =

[b + z (S b )]2
(S b )

(c)

(c)

(c)

(1 ") v r B

Deff
Sb z

(c)

(1 ") B

(5.3.10)
which, on substituting for v r (c) from (3.3.31) gives, after re-arrangement,
D1 =

(1 ") (c)
Sb

B (c) f (t )

[b + z (S b )]2 (c)
(S b )

0A

Deff

B (c)
z

. (5.3.11)

A similar equation applies for the diffusive contribution to solute transport in the shell
region. As explained in Section 4.5.1, the term f (t ) accounts for bulk movement of the

175

CSH

5.3. Solving the Model Equations


continuous phase due to an expanding bubble:
f (t ) = b 2

db
dt

(5.3.12)

where db
= 0 for a void-less droplet. When f (t ) = 0, the advective term in the solute evodt
lution equation, (3.3.16), becomes a diffusionlike component on substituting for v r (c) .
This is expected since the mass-averaged velocity arises naturally from the assumption of
Fickian diffusion rather than as a consequence of any bulk flow, (3.3.6). Clearly the situation is different when there is a bulk movement, i.e., when f (t ) 6= 0, in which case a true
advective term is introduced.
The diffusion terms for the moments are

[b + z (S b )]2
D

m
a
v r(d ) ma
, a {0, 1, 2} ,
(5.3.13)
D5a =
Sb
Sb z
with a similar expression being obtained for the solids volume fraction and the moments in
the shell region. The solids velocity, v r(d ) , may include a contribution from an expanding
bubble,
v r(d )

b2

db

(rL + z (rR rL ))2 dt

(5.3.14)

where [rL , rR ] = [b , S] in the core region or [rL , rR ] = [S, R] in the shell. Alternatively, the
solids velocity in the shell may be modified to include a term used to model shell buckling
during the thickening regime,
v r(d )

R2

dR
2

[S + z (R S)] dt

(5.3.15)


T
The complete diffusive flux term may be expressed D = Din , Dshell , where

(c) Deff B (c)


B (c) f (t )
(c)
(Sb )0 z
(1 ")
[b +z(Sb )]2
A

D "
(d )
2
v r " Sb z
[b + z (S b )]

D m2
(d )
Din (U , z; t ) =

v r m2 Sb
z
(S b )

D m1
(d )

v r m1 Sb z

D m0
v r(d ) m0 Sb
z

(5.3.16a)

176

CSH

5.3. Solving the Model Equations


and, noting that the solid particles in the shell region are no longer free to diffuse in space,
Dshell (U , z; t ) =

[S + z (R S)]2

(R S)



(c)
(c)
(c)

B shell f (t )
shell Deff B shell
(c)
1 "shell shell [S+z(RS)]2 (RS)0
z

(d
)

v r "shell

2, shell
v r(d ) m

(d
)

1, shell
vr m

0, shell
v r(d ) m

(5.3.16b)
The form of f (t ) and v r(d ) depends on the drying regime as discussed above.
The Convective Flux, F , term
The convective terms in the model equations arise from the virtual flux introduced following
the co-ordinate transformation, (5.1.17). Inspection of the model equations in Section 5.2
(e.g., (5.2.2),(5.2.6) and (5.2.7)), shows that the co-ordinate transformation adds virtual
flux terms of the form



1 drL
z
drR drL
.
(5.3.17)

=

t z t r rR rL dt z rR rL dt
dt z
|
{z
}
V F

The first of these virtual flux terms is already in a form suitable for incorporation into
the convective F term; the second of the two terms requires further thought. Reference
to (5.3.4) shows that the NAG D03PLF routine does not permit the coefficients of F to
have a z dependence. The second virtual flux term has such a dependence on the spatial
co-ordinate and is therefore rewritten using
z

f
z

(5.3.18)

[z f (z)] f (z) ,

i.e., the problematic coefficient is eliminated at the expense of introducing an additional


T

source term. Writing F = Fin , Fshell , the flux term is given by

(1 ") (c) B (c)


"

1
dS db
db

Fin (U , z; t ) =
+z

S b dt
dt
dt

1
m

0
m

177

, (5.3.19a)

CSH

5.3. Solving the Model Equations


and


(c)
(c)
1 "shell shell B shell


"shell

1
dS
dR dS

Fshell (U , z; t ) =
+z

2, shell

R S dt
dt
dt

1, shell
m

0, shell
m

(5.3.19b)
As explained above, this F term is evaluated by the numflx subroutine, discussed in detail
in Section 5.5.
The Source, S , term
T

The final terms requiring definition are the source terms, expressed S = Uin , Ushell . As
just discussed, S includes the terms introduced by applying (5.3.18) to the advective fluxes.
Inspection of (5.2.2) shows that

"

0D

t

=
crys

0D

(0)
2
Gm2 m

L3min N0

(5.3.20)

is a source in the continuous phase equation. Further, the equations describing the moment
system, (5.2.6) and (5.2.7), contain source terms relating to particle nucleation and growth.
The resulting source terms for implementation are therefore

(1 ") (c) B (c)

"

1
dS db

Sin (U , z; t ) =

S b dt
dt

m
1

0
m

(0)

3
0

Gm
m
+
L
N
2
2
6 min 0
D 2

(0)

G m2 m2 + 6 Lmin N0

(0)
1
2

N
2Gm
m
+
L
(0)
1
1
min 0
+
m2

(0)
1

0 + Lmin N0
Gm0 m

(0)
m1

N0

(0)

m0

(5.3.21a)

178

CSH

5.3. Solving the Model Equations


and


(c)
(c)
1 "shell shell B shell

"
shell

1
dR dS

Sshell (U , z; t ) =

2, shell

R S dt
dt

m
1,
shell

0, shell
m


(0)

Gm
m
2,
shell
2
D 2

(0)

Gm
m

1 2 2 (0) 2, shell

1, shell
+ m (0) 2Gm1 m
.
2

(0)
1

0, shell
m (0) Gm0 m

(5.3.21b)

5.3.3 Definition of the Coupled ODEs odedef


The NAG D03PLF allows a system of coupled ordinary differential equations to be solved in
parallel with the partial differential equation system. Indeed, in essence, this is just a matter
of appending these additional equations to the system of ordinary differential equations
obtained when the method of lines is applied to the partial differential equations.
The coupled ordinary differential equations must be such that they can be expressed in
the general form




u u
dv
Ri v, , , u| ,
,
; t = 0,
dt
z t

i = 1, 2, . . . , ncode

(5.3.22)

where ncode is the number of coupled ordinary differential equations and is a vector of
spatial coupling points. The solution values of these equations are contained in the vector
v , with the numerical approximations to these values denoted V . The ordinary differential
equations describing the evolution of these variables may only depend on the values and
derivatives of u the vector of partial differential equation solution values at the predefined coupling points, .
For the present model, the vector of ordinary differential equation solution values is
T
db dS dR
, Td (t ) , M (t )
.
v (t ) = b (t ) , S (t ) , R (t ) , , ,
dt dt dt

(5.3.23)

The first three terms are required to track the evolving solution domains as a result of
the co-ordinate transformation described in Section 5.1.3. At first sight, the second set of
three values representing the time derivatives of the first three terms might appear

179

CSH

5.3. Solving the Model Equations


superfluous. However, inspection of (5.1.18) shows that these are included in the virtual
flux term introduced as a result of the co-ordinate transformation. Further, recall from
(5.3.4) that whilst the flux term supplied to the D03PLF routine may show a functional
dependence on the ordinary differential equation solution values, v , it may not depend on
their time derivatives. It is therefore necessary to track these terms explicitly.
The relevant equations describing the time evolution of b , S and R in all the different
drying regimes are given in Section 5.2. For example, the ordinary differential equations
prior to shell formation are
db
dS
dt

dR
dt

dt
00
m
0A

=0 ,

(5.3.24a)

=0 .

(5.3.24b)

The equations required for the components v4 , v5 and v6 representing the time derivatives of the first three variables in v are
vi

dvi3
dt

= 0,

(5.3.25)

i = 4, 5, 6 .

The final two components in v represent the solution values of the ordinary differential
equations describing the droplet temperature and droplet mass. The rate of change of
droplet mass is given by
dM
dt

(5.3.26)

vap = 0 ,
+m

vap is the total rate of mass loss from the droplet, calculated by Algorithm 3.1.
where m
The solution value of this equation, i.e., the droplet mass, is used to determine the droplet
temperature using
dTd
dt

Q
b
MC
p, drop

(5.3.27)

=0 ,

where Q , the heat penetrating the droplet per unit time, is also calculated using Algorithm
3.1, (3.5.3).
The user-defined subroutine, odedef, evaluates R when supplied with the value of U
and its derivatives at the coupling points, . It only remains to specify the location of these
coupling points. It is clearly desirable to minimise the number of such points and inspection
of the ordinary differential equations requiring solution shows that they only depend on the
end values of u. Therefore two coupling points are required located at the end points of the

180

CSH

5.4. Boundary Conditions and Solution Grid


normalised z domain, i.e.,
= [0, 1]T .

(5.3.28)

This concludes the formulation of the model equations for solution using the NAG
D03PLF routine. The remaining sections in this chapter focus on issues relating to the
boundary conditions, numerical flux function and general stability and convergence tests.

5.4 Boundary Conditions and Solution Grid


The boundary conditions are a key part of the problem formulation and, further, it is
found that they have a major influence on the numerical stability of the solution method.
The grid points must be located in such a way that the boundary gradients are correctly
captured, as described in Section 5.4.2. However, before investigating the solution grid,
the implementation of the boundary conditions within the NAG D03PLF routine is briefly
discussed.

5.4.1 Definition of the Boundary Conditions bndary


Boundary conditions on the solution domain z = [0, 1] are implemented through the usersupplied subroutine, bndary. These boundary conditions must be expressed in the general
form

dv
Gi u, v, , z; t = 0 , i = 1, 2, . . . , npde .
(5.4.1)
dt
This form does not permit explicit dependence on the spatial derivatives of the solution
variables, u. Instead, where these are required, it is necessary to construct numerical approximations to these gradients using the numerically determined solution variables, Ui (t ),
at grid points in the neighbourhood of the boundary. Such approximations are obtained
using one-sided finite differences, the simplest of which is the two-point formula,

u
Ui Ui 1
.

z i
z

(5.4.2)

This formula has the advantage of simplicity, but is only first-order accurate in z . Higher
order methods generally make use of the solution values at more grid points; the formula
based on n grid points has the theoretical order one less than n , (Chung, 2002; Bieniasz,
2002). The coefficients for such one-sided formulae up to n = 6 are listed in Table 5.2

181

CSH

5.4. Boundary Conditions and Solution Grid


Table 5.2: Coefficients of the one-sided n -point gradient approximations written for a right hand boundary.

znpts5

2:

(z)1

3:

(2z)1

4:

(6z)

5:

(12z)1

6:

(60z)1

znpts4

-12

znpts3

znpts2

znpts1

znpts

-1

-4

-2

-18

11

-16

36

-48

25

75

-200

300

-300

137

for example, the three-point formula is



u

z

3Unpts 4Unpts1 + Unpts2


2z

z=1


+ O z 3 .

(5.4.3)

Once the boundary gradients have been approximated using these formulae, it is a simple matter to specify the boundary conditions listed in Section 5.2 in the form given by
(5.4.1). However, it is noted that such one-sided differencing has potential to cause numerical problems, particularly at the external boundary where it opposes the general direction
of information flow. Such numerical issues are investigated in the following sections.

5.4.2 Grid Resolution


The physical nature of the solution variables imposes a maximum grid spacing in the neighbourhood of the external boundary. Consider using the two-point formula, (5.4.2), to
provide a numerical approximation to an external Neumann boundary condition on u .
Further suppose that the values of u are required to lie in the range [0, 1], as is the case for
variables representing mass or volume fractions. Such a situation is illustrated schematically
in Figure 5.6. When
z 


u

z

!1

(5.4.4)

,
z=1

as in Figure 5.6a, the numerical approximations at the grid points reproduce the true solution well. However, when
z


u

z

!1

(5.4.5)

,
z=1

182

CSH

5.4. Boundary Conditions and Solution Grid


U

u

z z=1

u

z z=1

z 0

(a)

1
(b)

Figure 5.6: Numerical approximation of a Neumann boundary condition using the two point, one sided
 1
 1


difference formula, (5.4.2) with: (a) z  uz
; and (b) z uz
.
z=1

z=1

as illustrated in Figure 5.6b, the condition imposed by the boundary gradient forces the
numerical values at the boundary nodes to move away from the true solution profile. If
z >


u

z

!1

(5.4.6)

,
z=1

the numerical method is incapable of reproducing the required gradient whilst maintaining
physically realistic values at the boundary nodes. Here, the two point one-sided difference
formula has been used for illustration, but a similar conclusion holds when using expressions based on more grid points. This has severe consequences for the present model as
the routines used to calculate physical properties (Figure 5.5) can return numerical errors
when passed unphysical values. Whilst error capture routines prevent such incidents from
immediately crashing the program, repeated unphysical calls are an obstacle to accurate
simulation.
The external boundary conditions on the profiles in the new drying model frequently
impose steep gradients near the droplet surface. This results from the fact that internal
diffusion is often slow when compared with the evaporation or shrinkage rate. For example,
consider using the simple two point formula, (5.4.2), to approximate (5.2.10) the solids
volume fraction gradient at the external boundary prior to shell formation,

"

z

= (rR rL )
z=1

00
m

D0A

(5.4.7)

".

Taking representative values from the sodium sulphate test case discussed in Section 3.6.2

183

CSH

5.4. Boundary Conditions and Solution Grid


gives an idea of the magnitude of the gradients expected. At the point of shell formation,

"

z


3
5.8

10
kg m2 s1
4
0.65
7.5 10 m
 2 1
3
11
3
1.5

10
m
s

10
kg
m
z=1
190 ,

which, following the discussion above implies that the grid spacing near the boundary
should satisfy
z <

1
190

(5.4.8)

if the solids volume fraction is to lie in the physical range [0, 1] at the two external solution
nodes. Indeed, it is expected and is verified by numerical experiments below that this
is an absolute minimum requirement; a significantly finer grid is required to achieve stable,
converged profiles that are independent of the solution grid.
It is often impractical to employ a uniform solution grid across the entire domain given
the requirement on the minimum grid density at the external boundary. Instead, the grid
points may be placed according to an exponential scheme with the i th node located at
h

i
1
i
1 exp npts1
zi =
.
h i
1 exp 1

(5.4.9)

Figure 5.7 illustrates the construction of a grid containing 11 points using (5.4.9); employing this method to place points clusters the nodes closer together near the right hand
boundary. The parameter is used to adjust the degree of clustering. The grid in Figure 5.7
was constructed using = 0.5 but, as is shown in Figure 5.8, assigning a smaller value to
results in a greater density of points in the vacinity of the right hand boundary.
The use of a non-uniform grid must be reflected in the one sided finite difference formulae used to approximate the solution gradients. For example, the appropriate three point
formula based on the grid shown in Figure 5.9, is

u

z

=
npts

znpts + znpts1 Unpts Unpts1


znpts1

znpts

znpts

Unpts Unpts2

znpts1 znpts + znpts1

. (5.4.10)

This expression reduces to (5.4.3) when evaluated on a uniform grid, (Hirsch, 1992a).

184

CSH

5.4. Boundary Conditions and Solution Grid

1
0.9
0.8

i/(npts1)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Grid Point Location, z

Figure 5.7: Construction of a solution grid placing grid points using (5.4.9) and taking = 0.5.

1
0.9

=0.04
=0.06
=0.08
=0.10
=0.15

0.8

i/(npts1)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Grid Point Location, z

Figure 5.8: Illustrating the effect of varying the parameter, in (5.4.9). Lower values of result in
more grid points clustered in the neighbourhood of the right hand boundary.

185

CSH

5.4. Boundary Conditions and Solution Grid


Cnpts2
znpts2
znpts3

z(npts2) 1

znpts1
znpts2

z(npts1) 1

znpts
znpts1

znpts

Figure 5.9: An arbitrary distribution of grid points in the neighbourhood of the right hand boundary of
the z domain. The grid cells are defined by the primitive points, with cell interfaces located at zi 1 .
2

5.4.3 Grid Studies


The results obtained from solving the new droplet drying model were found to exhibit varying levels of dependence on the solution grid employed. Such effects are clearly undesirable
and, consequently, extensive grid studies were conducted for all of the simulations presented
in Chapters 3 and 4. In this section, two of these convergence studies are presented in detail.
Silica Grid Dependence
Sections 3.6.1 and 4.6.2 presented simulations of colloidal silica droplets drying in air. This
system consists of mono-disperse solid particles suspended in a pure liquid and, as such,
is the least computationally demanding of the applications investigated. Grid independent
results were obtained when solving the model on a sufficiently fine uniform mesh.
To investigate grid dependence in a quantitative manner, the predicted time of initial
shell formation, tshell , was taken as a comparative measure. Recall that shell formation is
predicted when the solids volume fraction at the surface rises above a pre-defined critical
level. This is, in turn, determined by the rates of moisture evaporation from and solids
transport within the droplet. The dependence of tshell on these variables makes it a good
candidate for assessing solution convergence.
Figure 5.10 illustrates how the predicted time to initial shell formation depends on grid
cell density for a droplet containing 500 nm silica particles drying in air at 178 C, (4.6.2).
Whilst the grid dependence is strong for very sparse meshes, the solution becomes essentially independent of the cell density above npts = 200. This is qualitatively confirmed in
Figure 5.11, which shows the solids volume fraction profiles at the point of shell formation.
Whilst satisfactory solutions are obtained using uniform meshes for the colloidal silica
system, it is instructive to compare results obtained when simulating this system on a nonuniform grid generated using (5.4.9). Figure 5.12 shows the solids volume fraction profiles
at the point of shell formation obtained from a grid described by (5.4.9) containing 20
cells and taking = 0.2. This profile is compared with those obtained from uniform grids
containing 20 and 500 cells. Clearly the profile produced using the non-uniform grid is
superior to that from a uniform grid containing the same number of points, whilst the

186

CSH

5.4. Boundary Conditions and Solution Grid

25
24.5
24

tshell / s

23.5
23
22.5
22
21.5
21
20.5
20
0

200

400

600

800

1000

npts

Figure 5.10: Plot showing the predicted time to shell formation, tshell , when drying a droplet of colloidal
silica, illustrating the dependence on the number of cells in the solution grid, npts.

0.65

Solids Volume Fraction / []

0.6
0.55

npts=20
npts=50
npts=100
npts=500

0.5
0.45
0.4
0.35
0.3
0.25
0.2
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Radial Position / mm

Figure 5.11: Plot illustrating the solids volume fraction within a droplet of colloidal silica at the point of
shell formation, showing the effect of increasing cell density in a uniform grid.

187

CSH

5.4. Boundary Conditions and Solution Grid

0.65

Solids Volume Fraction / []

0.6

npts=20
npts=500
npts=20, NonUniform Grid

0.55
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Radial Position / mm

Figure 5.12: Plot illustrating the solids volume fraction within a droplet of colloidal silica at the point
of shell formation. The profiles obtained using a uniform grid containing 20 and 500 cells are compared
with that returned when solving on a non-uniform grid containing 20 cells.

CPU time required is roughly similar. Calculating the l 2 -norm of the deviations from the
high accuracy solution obtained using a grid containing 500 cells confirms this observation;
the values l 2 -norm are shown along with the associated CPU times in Table 5.3.
Table 5.3: Comparison of accuracy and computational times for solutions obtained on a uniform and
non-uniform grid. The l 2 -norm reported is that of the deviations of associated solution and that obtained
using a uniform grid containing 500 cells.

l 2 -norm CPU time / s

Grid
npts = 500
uniform
npts = 20
uniform
npts = 20 non-uniform

1.16
0.170

51.8
2.17
2.73

The non-uniform grid concentrates grid points towards the right hand end of the domain and, consequently, the fidelity of the solution produced using this grid is worst towards
the centre of the droplet. However, most processes of interest occur at the droplet exterior
where the profile reproduction is good. The non-uniform grid therefore provides a method
of obtaining accurate results whilst incurring a considerably lower computational cost than
high resolution methods on a uniform mesh.

188

CSH

5.4. Boundary Conditions and Solution Grid


Sodium Sulphate Grid Dependence
Section 3.6.1 presented simulations of a droplet containing dissolved sodium sulphate drying in air. When solving the equations describing this system on a uniform grid, the solution fails to demonstrate grid independence before the maximum allowable number of
nodes is reached.31 Such behaviour is due to the steep gradients imposed on the solution
at the external droplet boundary, as discussed in Section 5.4.2. In particular, the relative
immobility of the growing crystals of sodium sulphate causes the solids volume fraction to
display extremely steep gradients. To overcome this difficulty, the model was solved using
a non-uniform grid described by (5.4.9). Comparisons of results obtained using uniform
and non-uniform grids for the colloidal silica example suggest that non-uniform grids may
be employed with confidence.
The convergence of the solution was monitored using the predicted time to shell formation which, for the reasons discussed in relation to the colloidal silica simulations, is
considered to be a good indicator of convergence in all variables. Simulating a droplet of
sodium sulphate with an initial particle number density of 3.39 108 m3 and maximum
nucleation rate, Nmax = 1020 m3 s1 , Figure 5.13(a) shows the predicted timing of shell formation, tshell , as the grid spacing parameter, , is decreased. Reference to Figure 5.8 shows
that lowering the values of increases the concentration of the grid points at the droplet
edge. The convergence with decreasing is clear, although it is perhaps more instructive
to plot tshell against the separation of the outermost grid points at znpts1 and znpts , as in
Figure 5.13(b). This second plot confirms that the spacing between these extreme grid
points is a key parameter; tshell is seen to increase linearly with decreasing znpts znpts1
until convergence is achieved.
The choice of finite difference formula used to approximate the boundary derivatives influences the solution prior to convergence. The results presented Figure 5.13 were obtained
using the three point formula given in (5.4.10); recall that the difference formulae used
must be those for a uneven grid. Figure 5.14 shows the predicted timing of shell formation
as the solution grid is refined, using two, three and four point finite difference formulae
to approximate the gradients at the boundaries. All three approximations converge to the
give the same value of tshell as the grid points become increasingly concentrated at the outer
boundary, but the formulae based on more points converge more quickly. This is expected,
since finite difference expressions based on more solution values are generally higher order.32
Figure 5.15 demonstrates the efficiency of the solutions obtained using different grid
31
32

Whilst the NAG D03PLF does not explicitly impose an upper limit on the number of grid points that
may be specified, it was found that memory limitations prohibited runs with npts > 103 .
For the finite difference approximation to a first derivative based on n solution values written on a
uniform grid, the formal order is n 1. When using a non-uniform grid, the formal order of the
approximation is, in general, less than n1, although the order is still expected to increase with increasing
n , (Hoffman, 1982).

189

CSH

5.4. Boundary Conditions and Solution Grid

50

tshell / s

45

40

35

30

25
0.04

0.06

0.08

0.1

0.12

0.14

0.16

(a)

50

tshell / s

45

40

35

30

25 12
10

10

10

10

10

10

znptsznpts1 / []

(b)
Figure 5.13: The timing of shell formation when drying a droplet of sodium sulphate converges with: (a)
decreasing ; and (b) decreasing znpts znpts1 . Reference to (5.4.9) shows that reducing concentrates
grid points at the right hand end of the domain which, in turn, reduces znpts znpts1 .

190

CSH

5.5. The Numerical Flux Function

50
45

tshell / s

40
35
30
25

2point formula
3point formula
4point formula

20 12
10

10

10

10

npts

npts1

10

10

/ []

Figure 5.14: Plot illustrating the influence of the finite difference formula used to approximate the
boundary gradients on the convergence behaviour of the solution.

spacings and finite difference approximations to the boundary gradients. The accuracy
of the various solutions is measured using the percentage error in the predicted time of
shell formation. The true value of tshell is taken as 47.60 s the value predicted using
the most refined mesh and the 4-point finite difference formula at the boundary. Plotted
against CPU time,33 the figure shows the conditions required for a given accuracy. All the
simulations presented in Section 3.6.1 were obtained using = 0.05 with the three point
finite difference approximation.

5.5 The Numerical Flux Function


5.5.1 Definition of the Numerical Flux Function numflx
Hyperbolic equations or hyperbolic terms in partial differential equations of mixed
type generally require special treatment, as discussed in Section 5.1.2. The effect of
such terms on the solution values in a given cell, Ci , may be approximated by the information fluxes across the cell boundaries. These boundary fluxes may be described by flux
vectors, Fi 1 which, in turn, may be approximated using the solution values in the cells at
2

33

All simulations were run on an AMD 2.2 GHz machine with 2 GB of RAM.

191

CSH

5.5. The Numerical Flux Function

50
2point formula
3point formula
4point formula

45

Error in tshell / %

40
35
30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

80

CPU Time / s

Figure 5.15: Plot demonstrating the efficiency of the solutions obtained using various grid spacings and
finite difference approximations to the boundary gradients.

either side of the interface,

Fi 1 = F Ui 1 , Ui ,

(5.5.1)

where F is the numerical flux function. The D03PLF routine uses such a numerical flux
function to handle F , the hyperbolic term in (5.3.4). For the present model, this term is

T
given by (5.3.19a) and (5.3.19b), where F = Fin , Fshell . The numerical flux function
itself is specified in the user-defined routine, numflx which, when passed Ui 1 and Ui is

required to evaluate F Ui 1 , Ui at the midpoint of the solution grid,


zi 1 =
2

zi + zi1
2

(5.5.2)

Reference to Figure 5.1 shows that this mid-point coincides, as expected, with the cell
boundary.
There are a vast number of ways in which the boundary fluxes might be approximated.
The simplest method is to use an arithmetic average of the hyperbolic flux term evaluated
using the values of Ui in the cells either side of the interface. That is,


1
Fin 1 = F Ui 1 , Ui =
F Ui1 + F Ui
.
2
2

192

(5.5.3)

CSH

5.5. The Numerical Flux Function


Sadly, this simple method is generally unstable for hyperbolic problems, (LeVeque, 2002),
and this was found to be the case in the present application. Instead, a more sophisticated
approach is required.

5.5.2 Roes Method for the Numerical Flux Function


For the constant coefficient linear problem,
u
t

+A

u
x

(5.5.4)

=0 ,

the flux function may be expressed in the form


Fi 1 =
2

1
F Ui 1 + F Ui |A| Ui Ui1 .
2
2

(5.5.5)

The matrix |A| is defined


|A| = R || R 1 ,

(5.5.6)

where R is the matrix of right-eigenvectors of A and || = diag (| p |), where p are the
eigenvalues of A. This way of writing the flux is similar to the arithmetic averaged expression, (5.5.3), with an additional correction term. For linear equations such as (5.5.4), the
flux expression (5.5.5) is identical to the upwind methods touched upon in Section 5.1.2.
The reason for introducing this form of the flux function is its extension for non-linear
problems in which it is known as Roes Method, (Roe, 1981); such an approach is recommended for use with the NAG D03PLF routine when solving non-linear problems, (NAG).
It is perhaps natural to approximate the non-linear equation
u
t

(5.5.7)

f (u) = 0 ,

with a linearised approximation at the cell interfaces,


u
t

1
+A
i
2

u
x

(5.5.8)

=0 ,

1 is some approximation to the Jacobian of f valid in the neighbourhood of each


where A
i 2
interface. Specifically,

1 =f0 U
,
A
i

(5.5.9)

is some average of U and U , the solution values in the cells either side of
where U
i
i 1

193

CSH

5.6. Numerical Parameter Studies


the interface; Roes linearisation method chooses this average in such a way that the results
satisfies certain useful conditions, (LeVeque, 2002). The Roe flux for the present problem
is derived in Appendix B. The appropriately linearised Jacobian is then used to replace A
in (5.5.5); for the inner core section of the drying droplet, this gives the flux function
Fi 1
2

1 
1

=
Fin Ui 1 + Fin Ui zi 1
2
2
2
(c)
0A2B (1"i )
0AB B i1




0
+ (c) + (c)
(c)
B +B i1
B B i1 B B i

0
1
0

0
0
1

0
0
0

0
0
0

0 0
0 0
0 0
1 0
0 1

Ui Ui1 ,

(5.5.10)
where




1

z

i 2 =


db
db
dS


+ zi 1


.
2

S b dt
dt
dt
1

(5.5.11)

The fluxes, Fin , in (5.5.10) are those derived in the previous section, i.e.,

(1 ") (c) B (c)


"

1
db
dS db

Fin (U , z; t ) =
+z

S b dt
dt
dt

1
m

0
m

(5.5.12)

The numerical flux function for the equations defined in the shell region follows from this
result by inspection, using Fshell given by (5.3.19b).
Considering the Roe flux function, (5.5.10), it is apparent that it reduces to aupwinding



scheme for all but the first component. The upwinding velocity is given by zi 1 , as
2
defined in (5.5.11). This is expected since the Roe flux function is known to be degenerate
with the switching upwind scheme for linear equations.

5.6 Numerical Parameter Studies


The NAG D03PLF routine used to solve the equation system describing the new droplet
drying model must be supplied with a number of key parameters which control the nature
of the solution method, the accuracy sought and time steps employed. The sensitivity of

194

CSH

5.6. Numerical Parameter Studies


the solution to the choice of these parameters is investigated in this section.

5.6.1 Solution Method


The general solution method employed has been discussed at length in Section 5.3.1. However, one further option must be specified to determine the algorithm used: whether sparse,
banded or full matrix methods should be used to solve the system. Banded methods were
not suitable due to the presence of coupled ordinary differential equations and, following
shell formation, non-locality of some boundary conditions. The system was solved using
both the sparse and full matrix methods and the solutions obtained were shown to be identical. However, the time taken using sparse matrix methods was, as expected, considerably
less. For this reason, all the results presented in this thesis were obtained employing these
sparse methods.

5.6.2 Error Tolerances


The accuracy of the numerical solution produced by the NAG D03PLF routine is controlled
by specifying relative and absolute error tolerances, denoted rtol and atol respectively.
Having specified values for rtol and atol, the NAG D03PLF forms the weights,
wi = rtoli |Ui | + atoli

1

i = 1, . . . , neqn ,

(5.6.1)

and then calculates either the l 1 - or l 2 -norm of a weighted local error estimate, ei , (NAG).
Having chosen the type of norm to be used, the local error test to be satisfied is
(5.6.2)

kei k1 or 2 < 1.0 .

It is possible to specify individual error tolerances for each ordinary differential equation
in the problem, as indicated by the indexes on rtol and atol in (5.6.1). These values
can be optimised for accuracy and speed, but for all the simulations in the present work, a
single value is used for all equations. This decision reflects the emphasis placed upon model
development and validation as opposed to computational speed. These tolerances might
be fine tuned if, at a later date, the new droplet drying model was being extended for use
within a computational fluid dynamics code where CPU time is clearly a more pressing
concern.
Within the present work, the accuracy is controlled through the relative error tolerance,
with atol set at all times to a small value, i.e., atol = 1012 . The advantage of controlling
accuracy using the relative tolerance is that errors in quantities with different orders of
magnitude for example, temperatures and mass fractions can be compared with ease.
Figure 5.16 illustrates the effect of varying the relative error tolerance on the timing of
195

CSH

5.6. Numerical Parameter Studies

47.135

47.13

ifail=10

shell

ifail=4

/s

47.125

47.12

47.115

47.11 10
10

10

10

10

10

10

10

Relative Error Tolerance / []

Figure 5.16: Plot illustrating the effect of varying the relative error tolerance, rtol, on the timing of shell
formation for a droplet of aqueous sodium sulphate solution. The solution algorithm is seen to fail when
the tolerance is set either very high and very low.

shell formation for a droplet containing aqueous sodium sulphate solution. As discussed in
Section 5.4, this parameter is thought to be a good measure of the accuracy of the solution
since it depends on a large number of the solution variables. Whilst the results in the figure
relate to a single trial on the sodium sulphate system, the behaviour shown is characteristic
of that observed across all the droplets simulated.
At high values of the relative error tolerance for rtol > 2 105 in the example
shown in Figure 5.16 the NAG D03PLF solution routine typically fails returning the
error flag ifail=10. This corresponds to the situation where a small change in rtol is
calculated to have no change on the solution values; in other words, the error tolerance
is too large to produce accurate solutions. If instead the relative error tolerance is set too
small rtol < 5 1010 in the example shown in Figure 5.16 then the solution
routine fails with the flag ifail=4, indicating that the solver has failed to initialise the
derivative of the ODE system. In essence, this means that the solver is unable to compute
the solution to the requested accuracy. Such an error is likely to be caused by just one of
the many ordinary differential equations comprising the system most likely that with
the smallest absolute solution value and setting individual values of rtol for each of the
neqn different equations would allow tighter tolerances to be specified for the majority of
the equations. However, as explained above, this was not investigated further within the
scope of the present thesis.

196

CSH

5.7. Conclusions of the Chapter


Over the range of rtol values for which the NAG D03PLF routine returns a solution,
the variation in predicted times to shell formation is seen from Figure 5.16 to be small.
Furthermore, the required CPU time varied by less than a factor of two between the slowest
and fastest runs, with no obvious correlation between rtol and CPU time. For these
reasons, it seems safe to assume that provided any solution is returned the choice of
numerical tolerance does not matter greatly. For most of the simulations presented in this
work the tolerances used were rtol = 106 and, as mentioned before, atol = 1012 .

5.6.3 Maximum Time Step


The NAG D03PLF routine uses an adaptive time step, t to integrate the ordinary differential equations describing the dry model through time. However, it is possible to specify an
upper limit on the allowable integration time steps, tmax ; setting such a limit ensures that
the CFL condition34 is satisfied at all times, (Schatzman, 2002; LeVeque, 2002).
It is found that the value assigned to tmax can, in practice, have a pronounced impact
upon the solutions obtained. The effect was most noticeable when simulating the drop
of detergent crutcher mix, discussed in Section 4.6.1. Figure 5.17 shows predicted solids
volume fraction profiles after drying a droplet of detergent crutcher mix for 50 minutes.
The sequence of four plots illustrates the qualitative effect of reducing the maximum time
step, tmax . The solution produced is unstable when using large values of tmax , with
oscillations plainly evident in the plots corresponding to tmax = 10 s and tmax = 1 s.
Reducing the maximum allowable time step stabilises the solutions produced and further
decreasing tmax demonstrates the solution has converged. The results presented in this
thesis were produced using tmax = 0.1 s which was found, in all cases, to yield converged
solutions.

5.7 Conclusions of the Chapter


The aim of this chapter was to describe and validate the numerical methods used to solve
the new droplet drying model introduced in Chapters 3 and 4. This model consists of a
system of advectiondiffusionreaction equations, coupled with additional ordinary differential equations. The partial differential equations are defined on spatial domains that are
themselves evolving in time. General numerical methods for solving such systems are discussed and the method of lines is identified as the most promising method. The moving
boundary problem was handled by applying a co-ordinate transformation to the equation

34

The CourantFriedrichsLewy (CFL) condition is a condition for convergence of hyperbolic partial


differential equations. Put simply, this states that the domain of dependence of the true solution must
lie within the domain of dependence of the numerical method, (Lewy et al., 1928).

197

CSH

5.7. Conclusions of the Chapter

0.65

0.65

tmax = 10 s

0.6

0.55

/ []

/ []

0.55
0.5

0.5

0.45

0.45

0.4

0.4

0.35

0.35

0.3

tmax = 1 s

0.6

0.2

0.4

0.6

0.3

0.8

0.2

Radial Position / mm

0.65

0.65

tmax = 0.5 s

0.6

0.8

0.6

0.8

0.55

/ []

/ []

0.6

tmax = 0.1 s

0.6

0.55
0.5

0.5

0.45

0.45

0.4

0.4

0.35

0.35

0.3

0.4

Radial Position / mm

0.2

0.4

0.6

0.8

0.3

Radial Position / mm

0.2

0.4

Radial Position / mm

Figure 5.17: Simulated profiles of the solids volume fraction is a droplet of crutcher mix after 50 minutes
of drying. The four plots illustrate the effect of varying the maximum integration time step, tmax . The
dotted line in each plot shows the stable solution produced using tmax = 0.1 s and the diamond indicates
the position of the dry shell interface.

system, thereby mapping the equations on to a fixed interval at the expense of introducing
a virtual flux term.
The NAG D03PLF routine was used to numerically solve the droplet drying model.
After stating the transformed model equations in full, it is shown how these may be formulated for solution by this library routine. Roes method is used to derive an expression
for the user-defined numerical flux function and the influence of the boundary conditions
is investigated. Numerical studies demonstrate the solutions obtained are grid independent
provided a sufficiently fine grid is used. The steep boundary gradients arising when simulating the droplet of sodium sulphate necessitates the use of an uneven solution mesh, with
grid points concentrated in the neighbourhood of the boundary. The validity of this grid
is established and grid independence of the solutions is again demonstrated. Investigations
demonstrate that using a higher order formula to approximate the boundary gradients re-

198

CSH

5.7. Conclusions of the Chapter


sults in faster convergence with grid refinement at the expense of CPU time. The chapter
ends by investigating the numerical error tolerances used, further supporting the validity of
the numerical results presented throughout this thesis.

199

CSH

Chapter 6
Conclusion
6.1 Conclusions of the Thesis
The goal of this thesis was to develop a new model for the drying of single droplets with
potential for use as a sub-model within spray dryer simulations. Whilst the literature as
reviewed in Chapter 2 contains several attempts to formulate such models, the attempts
all contain a number of deficiencies. A spatially resolved description of suspended solids
within the droplet is usually lacking and the models used to describe morphological development must be specified in advance. The lack of a clear physical basis for these models
makes their specification somewhat arbitrary and means that no existing model is capable
of simulating different dried-particle morphologies based upon droplet composition and
drying history.
This thesis presents the development of a new droplet drying model that seeks to overcome some of the weaknesses identified in previous descriptions. The continuous phase is
modelled as an ideal binary solution and volume averaging was applied to derive the appropriate transport equation for the solute. The suspended solids are described by means
of a population balance equation, allowing for particle nucleation, growth and transport
in physical space. Solving for the moments of this equation provides the spatially resolved
solids description that is lacking in most existing models and allows numerous properties
of interest such as particle number density, mean particle size and solids volume fraction to be simulated. Combined, the continuous and discrete phase descriptions provide
a comprehensive framework for describing drying droplets.
Formation of a surface shell is identified as the key event in the morphological evolution
of a drying droplet. The core droplet drying framework is augmented with various submodels to describe drying and structural evolution following shell formation. The choice
of sub-models implemented was motivated by a review of experimental observations from
the literature, along with a discussion of the physics of droplet drying in the presence of a
surface shell. New sub-models were developed to describe drying during a shell thickening
200

6.1. Conclusions of the Thesis


regime and during a period of slow boiling. In addition, the wet shell and dry shell models
from the literature were implemented.
The sub-models describing drying following shell formation are firmly grounded in
an understanding of the physical processes actually occuring during drying. Criteria may
therefore be developed for deciding which model to use at each point in the droplet drying
history. These criteria make use of information provided from the solid particle population
balance, such as mean particle size in the shell, as well as physical properties of the shell,
such as its Youngs modulus. In doing this the new droplet drying model provides, for the
first time, a framework within which the evolving particle composition and external drying
conditions can be used to inform developing droplet structure. Most importantly, the new
droplet drying model is capable of simulating structural evolution as influenced by droplet
composition and drying history, thus addressing the principle deficiency of existing models.
The new droplet drying model is successfully applied to simulate three different systems:
colloidal silica, aqueous sodium sulphate and detergent crutcher mix. The last of these systems is of considerable industrial interest and, whilst simplifications are made to facilitate
simulation using the new model, several interesting results are obtained. For all systems, experimental temperature and mass histories from the literature are reproduced by the model.
In addition, recently reported NMR observations of the detergent droplet allowed experimentally measured radial moisture profiles to be compared with those predicted by the
model. This represents the first time that a mechanistic drying model has been validated
against spatially resolved experimental data.
Simulations of the aqueous sodium sulphate system demonstrated the full capabilities
of the population balance approach for the solid particles. This study also emphasised
the importance of experimentally determined parameters when running simulations; the
lack of accurate information on particle nucleation kinetics hampered the production of a
greater range of quantitative results. Nevertheless, sensitivity studies enabled the influence of
particle nucleation rate and initial seeding on the timing of shell formation and mean solid
particle sizes to be investigated. Interesting trends were established: for example, increasing
the maximum nucleation rate was found to initially delay the onset of shell formation and,
when a shell did form, the average particle size was reduced.
The colloidal silica and detergent crutcher mix systems demonstrated the power of the
new model to simulate morphological evolution following shell formation. Using the criteria established to select a sub-model to use after the appearance of a surface shell, the
crutcher mix is predicted to dry via the dry shell route throughout. In contrast, the droplet
of colloidal silica demonstrates the use of all four sub-models and demonstrates the ability
of the new model to simulate different dried-particle morphologies depending on droplet
composition. Parameter studies illustrate how the structural evolution of the droplet is influenced by varying the structural properties of the shell, the drying air conditions and the

201

CSH

6.2. Suggestions For Future Work


initial droplet composition. Such studies are only possible as a result of the physically motivated criteria developed for choosing the appropriate sub-model at each point in the drying
history.
The droplet model is formulated as a system of advectiondiffusionreaction equations
defined on a moving spatial domain with coupled ordinary differential equations. A coordinate transformation maps this system to a fixed domain, on which it is solved using the
NAG D03PLF library routine. This routine implements the method of lines to reduce the
model to a system of ordinary differential equations which are then integrated through time
using a backward differentiation formula method. An uneven grid is required to capture the
steep boundary gradients which occur in some systems. Numerical studies demonstrate the
validity of this uneven mesh and show that the results obtained are grid independent. Together with studies investigating sensitivity to numerical solution parameters, this provides
confidence in the results presented in this work.
In summary, this thesis contained the following novel elements:
a framework to describe drying droplets based on volume averaged transport equations and a population balance describing suspended solids was developed;
the shell thickening sub-model was developed;
the wet and dry shell sub-models were extended, providing physical motivation and
demonstrating compatibility with the new model framework;
physically motivated criteria to determine the appropriate sub-model to use following shell formation were developed, thus allowing simulation of structural evolution
influenced by droplet composition and drying conditions;
the core and extended droplet drying models were used to simulate systems of practical interest and validated against experimental data.

6.2 Suggestions For Future Work


Suggestions for future work may be grouped in three broad categories: extensions to the
structure of the new model; acquisition of experimental data; and applications of the new
model.

6.2.1 Extending the New Model


There are a number of obvious ways in which the new droplet drying model presented in
this thesis might be extended. The core model is intended to provide a framework within

202

CSH

6.2. Suggestions For Future Work


which sub-models might be incorporated with relative ease. Four such sub-models are implemented within this work but that describing slow boiling is the least well developed. This
sub-model could be extended by modelling the vapour flow through the porous dry shell
and tracking the pressure rise at the evaporative interface. In combination with information concerning the structural properties of the shell, this would allow conditions leading
to droplet shattering to be simulated.
Continuous phase boiling prior to shell formation, or in the presence of a wet or thickening shell, is not presently incorporated within the drying model. Implementation of a
sub-model describing such inflation or puffing would be a significant advance, widening
the range of structural developments and dried-particle morphologies capable of being simulated. The infrastructure to handle a centrally located expanding bubble is in place as a
result of implementing the wet shell model this could be reused to implement a description of puffing. Previous attempts to simulate inflation have required specification of an
arbitrary maximum droplet size to prevent massively expanded shells being predicted. It
is possible that the solid phase description incorporated within the present model might
enable descriptions of inflation to be put on a more physical basis.
Beyond incorporation of new sub-models, the core droplet description might be extended. At present, the population balance permits evolution in the internal co-ordinate
through particle nucleation and growth. This might be extended to include agglomeration of suspended particles, thus enabling a more detailed description of shell formation.
Thought would need to be given to the coagulation kernel used but, in principle, it should
be possible for the present moment formulation to encompass such an extension.
Finally, the current model formulation allows crystallisation processes to be simulated
but there is no reason why solid production as a result of reaction should not be described in
a similar way. Indeed, homogenous chemical reactions in the continuous phase could also
be included if required. Such an extension would broaden the range of systems to which
the new droplet drying model could be applied.

6.2.2 Experimental Investigations


The lack of experimental data both regarding physical parameters and validation data
has proved the major problem in the course of producing this thesis. Prior to running a
simulation, a large number of physical parameters are required. These include diffusion coefficients, solution properties such as density, heat capacity and viscosity, material sorption
isotherms and crystal nucleation and growth kinetics. All of these are functions of temperature and composition and each has an appreciable impact upon the results obtained. To
accurately simulate a drying system it is necessary to have all of these data; sadly there are
very few systems which, to date, have been characterised in sufficient detail.
To overcome the poor characterisation of drying systems, modellers and experimentalists
203

CSH

6.2. Suggestions For Future Work


need to work more closely together. As demonstrated above, there are many parameters
requiring measurement and models can be used to indicate which of these most influences
the results. The efforts of experimentalists can thereby be focussed on the most important
factors and the results obtained can, in turn, be used to better judge the quality of different
simulation strategies. As shown through this work, experimental observations are also key
in guiding model development for example, the sub-models included here were chosen
on the basis of observed dried-particle morphologies. Direct feedback from experimentalist
to modellers would therefore guide the development of better models.
A second problem exists regarding validation data. Until recently, the only comparison
data against which models could be checked were droplet averaged temperature and mass
histories. This made judging the relative merits of more advanced spatially resolved and
mechanistic models difficult, if not impossible. In this work, for the first time, a mechanistic model is compared against spatially resolved experimental data obtained using NMR
measurements and this is a real advance. It is hoped that more such data might become
available in future. Here again closer collaboration between modellers and experimentalists
would be beneficial, with modellers able to suggest which measurements would be most
useful to distinguish between competing model descriptions.

6.2.3 Uses of the New Model


This thesis presents stand-alone simulations of three droplet drying systems. A natural extension would be to use the new model to simulate a wider range of systems skimmed
milk and concentrated saline solution are of particular interest. Such single droplet simulations fulfil the aims of the present project: providing useful physical insights into the
droplet drying process which may be used to elucidate the relationship between droplet
composition, drying rate and dried-particle structure. However, a wider hope was that the
droplet drying model might be incorporated within a full spray dryer simulation and this is
an obvious next step.
The new droplet drying model might be used to inform the development of cut-down
sub-models for inclusion within spray dryer simulations based on computational fluid dynamics. However, whilst no attempt has been made to optimise the current model with
respect to solution speed, reported CPU times suggest it might be feasible to implement the
complete model as it stands. Combining the ability of the new droplet model to simulate
multiple dried-particle structures with the detailed gas phase description provided by CFD
would produce a powerful simulation tool of considerable industrial interest.

204

CSH

Appendices

205

Appendix A. Abel Transformation

Appendix A
Abel Transformation
A.1 Theory of the Transformation
The Abel transformation is an integral transform most often although not always used
in the analysis of axially symmetric functions, (Bracewell, 1990). The spherical symmetry
assumed in the droplet drying model introduced in this thesis means that the discussion
here is limited to the special case of spherically symmetric functions.
Bracewell (1999) gives the Abel transform of a spherically symmetric function, f (r ),
35
as
Z
f (r ) r d r
F (s) = A [ f (r )] = 2
.
(A.1.2)
p
2
2
y
r s
The transform is seen to represent the projection of the function, f , along a set of parallel
lines on to, say, the y zplane. Clearly such a projection will have circular symmetry and,
as a result, the transformed function is expressible as F (s) where s 2 = y 2 + z 2 .
Many instrumental methods record cumulative intensities along a line-of-sight. The
inverse Abel transform is often used to determine an unknown function from such projected data. Such an approach is generally problematic however as many data points are
required for high-fidelity reconstruction and errors in the data are generally magnified by
the inversion procedure. Consequently, when comparing model predictions with experimental results it is preferable to Abel transform the simulation values and compare these
with the raw data, (Cremers and Birkebak, 1966). The problem of fidelity is still an issue
as seen in the next section, a large number of data points are required for an accurate
35

Some authors use the more general definition,


Z
F (s ) =
0

f (x)
(s x)

(A.1.1)

ds ,

where 0 < < 1, (Weisstein, retrieved 2008).

206

CSH

Appendix A. Abel Transformation


result from the numerical transform.

A.2 Testing the Transformation Algorithm


The NMR data of droplet moisture content used in this thesis goes further than the situation described above and collapses the three-dimensional physical distribution on to a
single line. This can be replicated in the simulation data by double application of the Abel
transform; the circular symmetry resulting from the first Abel transform renders the planar
projection suitable for further transformation.
A MatLab routine was written to perform the double Abel transform when passed a
one-dimensional data array. The routine was tested using the function
f1 (r ) = 2 1 r 2

for 0 < r < 1 ,

(A.2.1)

for which the analytical Abel transform is, (Bracewell, 1999)



F1 (s) = A [ f1 ] = 1 s 2 .

(A.2.2)

In turn, the Abel transform of F1 is


4
3
F2 (x) = A [F1 ] = 1 x 2 2 .
3

(A.2.3)

The radial domain, r , was sub divided into N equal intervals and the transform routine

was passed a vector containing the values f ri for i = 0, . . . , N . Figure A.1 shows the
results of the transformation, indicating the effect of increasing N . Clearly, low values of
N give poor reconstruction of the analytical solution, with N > 1000 needed for reasonable
fidelity; N = 105 gives near perfect results. However, the time taken for the transformation
increases with N 2 , as shown in Figure A.2. Further, the absolute times taken on an AMD
2.2 GHz machine with 2 GB of RAM, shown in Table A.1, are not insignificant.
Table A.1: Computational time to perform a double Abel Transform on (A.2.1). The CPU times shown
are the average of 10 runs on an AMD 2.2 GHz machine with 2 GB of RAM.

Average CPU Time / s

101

0.0400

102

0.0538

103

0.538

104

40.3

105

4122

207

CSH

Appendix A. Abel Transformation

4.5

Analytical Solution
1

N=10

N=10

Intensity / [a.u.]

3.5

N=10

N=104

N=105

2.5
2
1.5
1
0.5
0
0

0.2

0.4

0.6

0.8

x / [a.u.]

Figure A.1: Numerical transformation of (A.2.1) with profiles showing the effect of varying the number
of data points, N . The analytical solution, (A.2.3), is indicated by points and is almost coincident with
the transform that results when using N = 105 points.

Slope = 2

log

10

2
0

log

10

Figure A.2: The computation time, t , required to perform a double Abel transform on (A.2.1) is proportional to N 2 , where N is the number of data points used in the transform.

208

CSH

Appendix A. Abel Transformation


To balance fidelity of the reconstruction and computational speed, the drying simulation
results were up-sampled to N = 104 data points prior to transformation. The up-sampling
was necessary as the PDE solution grid was non-uniform and it would be prohibitive in
terms of computational time to run the entire simulation on such a fine mesh. Spline interpolation using intrinsic MatLab routines efficiently performed the required up-sampling.

209

CSH

Appendix B. Roe Flux Function

Appendix B
Roe Flux Function
B.1 Partial Differential Equation
It is sought to derive the Roe flux function, (Roe, 1981), for the non-linear equation,

u (z; t ) +

(B.1.1)

F (u, z; t ) = (u, z; t ) ,

where the equation is written in the form required by the NAG D03PLF routine. In this new
droplet drying model, it is seen from Section 5.3.2 that

(1 ") (c) B (c)


"

1
drL
drR drL

F (u, z; t ) =
+z

rR rL dt
dt
dt

1
m

0
m

, (B.1.2)

and

(1 ") (c) B (c)

"
1
drR drL

(u, z; t ) =

m
rR rL dt
dt

1
m

0
m

(B.1.3)

For convenience, these vectors have been written assuming ushell = 0, but the extension
to the more general case required following shell formation follows trivially. Further, the
z dependence of F is removed, allowing reincorporation of the source term .36 The
36

Recall that the source term in (B.1.1) was introduced in Section 5.3.2 when the model equations were
manipulated into the form required for the D03PLF routine.

210

CSH

Appendix B. Roe Flux Function


equation now reads

u (z; t ) + (z)

(B.1.4)

f (u; t ) = 0 ,

where
(z) =

drL

rR rL

dt

+z

drR
dt

drL

dt

(B.1.5)

B.2 Linearised Jacobian


To proceed with developing the Roe flux function for (B.1.1), a linearised approximation
of (B.1.4) is defined at each cell interface,

u
t

1
i 1 A
+
i
2

u
z

(B.2.1)

=0 .

1 is a linearised approximation to the Jacobian, f 0 (u), which is valid in the neighHere, A


i 2
bourhood of the interface.
1 . Roe Linearisation
The problem is now to determine the appropriate form for A
i 2
achieves this by imposing the condition

1 U U
A
=
f
U

f
U
,
i
i
i 1
i
i 1
2

(B.2.2)

and then integrating the Jacobian matrix, f 0 (u), along a path between Ui 1 and Ui 1 , (Roe,
1981; LeVeque, 2002). The straight-line path between Ui1 and Ui 1 may be parameterised

u ( ) = Ui 1 + Ui Ui 1 ,

(B.2.3)

where 0 1. Integrating the Jacobian along this line gives

f Ui

Z 1 d f (u ( ))
d
f Ui1 =
d
0
Z1
d f (u ( )) du
=
d
du
d
0
Z 1

d f (u ( ))
=
d
Ui Ui1
du
0

which, comparing with (B.2.2), gives


Z
1=
A
i
2

d f (u ( ))
du

(B.2.4)

d .

211

CSH

Appendix B. Roe Flux Function


Now, for the present problem

(1")0A2B
0AB B (c)
+ (c) 2 + (c)
B
B
( B B )

0
1
df

=
0
0
du

0
0

0
0

0 0 0
0 0 0
1 0 0
0 1 0
0 0 1

(B.2.5)

where the two terms of interest lie in the upper left corner of the matrix. Using the parameterisation, (B.2.3), gives

d f (u ( ))


du


 
1 "i 1 + "i "i 1 0A2B
=

2 ,
(c)
(c)
(c)
B + B i 1 + B i B i1
(1,1)

(B.2.6)


d f (u ( ))


du


(c)
(c)
(c)
0AB B i 1 + B i B i 1
=

.
(c)
(c)
(c)
B + B i 1 + B i B i 1
(1,2)

(B.2.7)

and

1 are obtained by integrating these expressions beThe corresponding components of A


i 2
tween = 0 and 1, according to (B.2.4). Using the standard integral,
Z
0

a + b
(c + d )2

d =

(ad b c)
c d (c + d )

b
d2

log 1 +

d
c

(B.2.8)

(B.2.6) integrates to give




(c)
(c)
(c)
1 "i 1 B i B i 1 + "i "i 1 B + B i 1

(c)
(c)
(c)
(c)

B
B
B
B
B
B
i
1
i
i1
i

1 = 0 2

A
.

i 2
A B
(c)
(c)

(1,1)

"

"
B i
B i1

i
i 1
log 1 +

2
(c)
(c)
(c)
B + B i
B i B i 1

(B.2.9)
This expression can be simplified considerably by making the entirely reasonable assump-

212

CSH

Appendix B. Roe Flux Function


tion37 that
(c)

(c)

B i B i1
(c)

B + B i

(B.2.10)

1 ,

in which case

1
A
i 2


0A2B 1 "i
=

.
(c)
(c)
(1,1)
B + B i 1 B + B i

(B.2.11)

Following a similar procedure for (B.2.7), using the standard integral


Z
0

a + b
c + b

d = 1 +

ac
b

log 1 +

(B.2.12)

gives

Ai 1
2

(1,2)

= 0AB 1

(c)
(c)

B i
B i1

log 1 +
. (B.2.13)
(c)
(c)
(c)
B i B i 1
B + B i 1
B

Again, it is reasonable to assume that the second term in the argument of the logarithm is
small, giving the approximation

Ai 1
2

(c)

(1,2)

0AB B i 1

(B.2.14)

(c)

B + B i 1

The linearised equation, (B.2.1), contains the position-dependent velocity, (z). In the
present implementation, this is evaluated at the mid-point of Ci , i.e., at z = zi 1 , and then
2

combined with the corresponding A 1 to give


i 2

1=
A
i
2

37

1
rR rL

drL

dt

+ zi 1

drR

drL

dt
dt
0A2B (1"i )



(c)
B +B i1

(c)

(c)

B +B i

0
0
0
0

0AB B i1
(c)

B +B i1

1
0
0
0

0 0 0

0 0 0

.
1 0 0

0 1 0

0 0 1

(B.2.15)

This assumption is justified since the change


values is likely to be small across almost
in the functional



all the grid cells, i.e., Ui Ui 1 implying Ui Ui1  1.

213

CSH

Appendix B. Roe Flux Function

B.3 Flux Function


Roes method obtains the numerical flux using the expression

=
F Ui 1 + F Ui Ai 1 Ui Ui 1 ,
2
2
2

Fi 1
2

(B.3.1)

where
1 | = R||R 1 .
|A
i

(B.3.2)

1 , and || = diag (| p |) is the matrix of


Here R is the matrix of right-eigenvectors of A
i 2
1 in (B.2.15) are seen by
absolute eigenvalues. For the present model, the eigenvalues of A
i 2

inspection to be




= zi 1
2

0A2B (1"i )



(c)
(c)
B +B i1 B +B i

1
1
1
1

(B.3.3)

For all the systems simulated in the present work, 0B 0A, and, consequently, B
(1, ). Since B (c) [0, 1], each bracketed term in the denominator of the first
eigenvalue is negative and so their product 
is greater
 than zero. The sign of the eigenvalues
therefore depends solely on the sign on zi 1 , and the numerical flux function from
2
Roes method takes the form
Fi 1 =
2

1 
1

F Ui1 + F Ui zi 1
2
2
2
(c)
0A2B (1"i )
0AB B i1
 + (c)  + (c)  + (c)
B B i1 B B i
B
B i1

0
1

0
0

0
0

0
0

0 0 0
0
1
0
0

0
0
1
0

0
0
0
1

Ui Ui1 , (B.3.4)

where




zi 12 =


dr
dr
dr
L
R
L
+ zi 1


.
2

rR rL dt
dt
dt
1

214

(B.3.5)

CSH

Nomenclature

Nomenclature
Greek Characters
B

x
r
M
T

"
"crit

0j

dry
mix

A
j

Dimensionless ratio of solvent and solute densities


Total stress tensor
State space of the internal coordinate
State space of the external coordinate
Mass transfer film thickness
Heat transfer film thickness
Surface tension
Solids volume fraction
Critical solids volume fraction triggering shell formation
Contact angle
Permeability
Thermal conductivity
Viscosity
Clustering parameter when constructing non-uniform grids
Poissons ratio
Fraction of sodium sulphate in the solute
Vector of spatial coupling points for coupled ODEs
Mass density
Material density of component j
Tortuosity
Characteristic drying time
Characteristic mixing time
Parameter relating BM and BT
Association factor
Mass fraction of component j , (liquid phase)

[-]
[Pa]
[m]
[m]
[Nm1 ]
[-]
[-]
[deg] or [rad]
[m2 ]
[Wm1 K1 ]
[Pas]
[-]
[-]
[-]
[-]
[kgm3 ]
[kgm3 ]
[-]
[s]
[s]
[-]
[-]
[-]

Operators
A

Abel transformation

(i )

B
Superficial volume average of B associated with phase i
(i )
B
Intrinsic volume average of B associated with phase i
B
Total volume average of B

Roman Characters
A

A
B

i 12

Area
Linearised Jacobian matrix
A scalar, vector or tensor quantity associated with a porous medium
215

[m2 ]
[-]

CSH

Nomenclature

D
D
Deff
D(AB)
DA,air
E
EA
F
Fm
FT
F 1
i

Spalding mass transfer number


Spalding heat transfer number
Molar concentration
Coefficient vector for PDE diffusive terms
Specific heat capacity of component j
Grid cell
Solids diffusion coefficient
Vector of PDE diffusive terms
Effective diffusion coefficient
Binary diffusion coefficient of A and B
Diffusion coefficient of water vapour in the drying gas
Youngs modulus
Activation energy
Vector of PDE advective terms
Mass transfer film correction factor
Heat transfer film correction factor
Flux vectors at the interfaces of the cell Ci

[-]
[-]
[kmolm3 ]
[-]
[Jkg1 K1 ]
[-]
[m2 s1 ]
[-]
[m2 s1 ]
[m2 s1 ]
[m2 s1 ]
[Pa]
[J]
[-]
[-]
[-]
[-]

F
G
H
Habs
HR
H vap
I
J
K
L
L
L0
Lmin
M
Ma
Nj
N0
Nmax
N
Nu
Nu0
P
Pcap
Pcrit
Pe
P
Q
R
R0

Numerical flux function


Linear growth rate
Absolute humidity
Percentage absolute humidity
Relative humidity
Latent heat of vaporisation
Number of phases
Number of components in a phase
Permeability tensor
Internal coordinate particle diameter
Average particle diameter
Characteristic dimension of an averaging surface
Minimum stable crystal size
Droplet mass
a th integer moment integrated over the droplet
Mole flux of component j
Particle nucleation rate per unit volume
Constant representing the maximum nucleation rate per unit volume
Number density function
Modified Nusselt number
Nusselt number in the limit of a low driving force
Pressure
Capillary pressure
Critical pressure determining the onset of wet shell drying
External pseudo-pressure acting on a growing shell
Coefficient matrix for PDE time derivatives
Heat penetrating into the droplet
Radius of the drying droplet
Initial radius of the drying droplet

[-]
[ms1 ]
[-]
[-]
[-]
[Jkmol1 K1 ]
[-]
[-]
[m2 ]
[m]
[m]
[m]
[m]
[kg]
[ma ]
[kmolm2 s1 ]
[# m3 s1 ]
[# m3 s1 ]
[# m4 ]
[-]
[-]
[Pa]
[Pa]
[Pa]
[Pa]
[-]
[Js1 ]
[m]
[m]

BM
BT
C
C
b
C
pj
C

216

CSH

Nomenclature
R
Rg
Ri
S
S
S
SV
Se
Sw
S
Sh
Sh0
T
Tboil
Td
Uin
V
Vm
V (i )
V
V
W
X
a
aw
b
ei
g
h
h
k
k
k
ksink
ma
ma(0)
a
m
00
m
vap
m
n
n

nr
p
q
r
r 00
rc

State vector of the external coordinates of a population balance


Gas constant
Position of the i th grid point
Closed averaging surface
Saturation Ratio
Radius of the inner edge of the shell
Surface area per unit volume
Portions of the averaging surface representing exits
Portions of the averaging surface representing walls
Vector of PDE source terms
Modified Sherwood number
Sherwood number in the limit of a low driving force
Shell thickness
Boiling temperature of the continuous phase
Droplet temperature
Numerical approximation to ui over the cell Ci at time tn
Volume
Molar volume
Volume of phase i
Volume integration variable
Numerical approximation to v
Relative molecular mass
State vector of the internal coordinates of a population balance
Pore radius
Water activity
Radius of the central bubble
Weighted local error estimate
Acceleration due to gravity
Heat transfer coefficient
Population balance source term
Boltzmann constant
Mass transfer coefficient based on a mass fraction driving force
Mass transfer coefficient in the limit of a low driving force
Sink diffusion coefficient
a th integer moment of the internal coordinate
Maximum initial value of ma
Non-dimensionalised a th integer moment of the internal coordinate
Mass vapour flux from the droplet
Total rate of solvent evaporation
Mass flux with respect to a stationary reference frame
Unit normal vector
Radial mass flux
Partial pressure
Volumetric flow rate
Internal coordinate radial position
Volume average interfacial production rate
Characteristic pore radius
217

[m]
[Jkmol1 K1 ]
[m]
[-]
[-]
[m]
[m2 m3 ]
[-]
[-]
[-]
[-]
[-]
[m]
[ C] or [K]
[ C] or [K]
[-]
[m3 ]
[m3 kmol1 ]
[m3 ]
[m3 ]
[-]
[kgkmol1 ]
[m]
[m]
[-]
[m]
[-]
[ms2 ]
[Wm2 K1 ]
[# m4 s1 ]
[JK1 ]
[kgm2 s1 ]
[kgm2 s1 ]
[ms1 ]
[ma3 ]
[ma3 ]
[-]
[kgm2 s1 ]
[kgs1 ]
[kgm2 s1 ]
[kgm2 s1 ]
[Pa]
[m3 s1 ]
[m]
[kgm2 s1 ]
[m]
CSH

Nomenclature
rj
rM
t
t DS
tshell
t
t
u
u
v
v
vr
vrel
wj
wi
x
xi 1
2
xi
x
y
z
z
z

Homogeneous mass rate of production of species j


Radius of curvature
Time
Time until formation of a dry shell
Time until initial shell formation
Integration time step
Data output time step
Dry-mass basis moisture fraction
Vector of functions whose evolution is described by PDEs
Mass-averaged velocity
Vector of variables whose evolution is described by coupled ODEs
Radial velocity
Relative drying air velocity
Mass fraction of component j , (gas phase)
Weight for a local error estimate
Spatial coordinate
Locations of the interfaces of the cell Ci
Width of the cell Ci
Mole fraction of component j , (liquid phase)
Mole fraction of component j , (gas phase)
non-dimensionalised spatial coordinate
Position vector for the point z in a porous medium
Width of a non-dimensionalised grid cell

[kgm3 s1 ]
[m]
[s]
[s]
[s]
[s]
[s]
[-]
[-]
[ms1 ]
[-]
[ms1 ]
[ms1 ]
[-]
[-]
[m]
[m]
[m]
[-]
[-]
[-]
[-]
[-]

Superscripts and Subscripts


+

X
X
A
B
D
L
R
buck
c
d
eqm
gas
in
sat
shell

The inner edge of the outer coordinate system, S + r


The outer edge of the inner coordinate system, S r
Average value of X in the mass transfer film
Local deviation from the averaged value of X
Solvent
Solute
Solids
Left-hand end of the domain
Right-hand end of the domain
The buckling condition
Continuous phase
Discrete phase
Equilibrium conditions
Bulk drying gas conditions
The central (wet) core region, r [b (t ) , S (t )]
Saturated conditions
The shell region, r [S (t ) , R (t )]

218

CSH

Nomenclature

Abbreviations
atol
CFD
CPU
LAS
NAG
NMR
ncode
neqn
npde
npts
ODE
PDE
RTD
rtol
SEM
UK
US

Absolute error tolerance


Computational Fluid Dynamics
Central Processing Unit
Linear Alkylbenzene Sulfonate
Numerical Algorithms Group
Nuclear Magnetic Resonance
Number of coupled ordinary differential equations
Number of equations
Number of partial differential equations
Number of points in the computational grid
Ordinary Differential Equation
Partial Differential Equation
Residence Time Distribution
Relative error tolerance
Scanning Electron Microscope
United Kingdom
United States

219

CSH

Bibliography
B. Abramzon and W. A. Sirignano. Droplet vaporization model for spray combustion
calculations. International Journal of Heat and Mass Transfer, 32(9):16051618, 1989.
doi:10.1016/0017-9310(89)90043-4.
B. Adhikari, T. Howes, B. R. Bhandari, and V. Troung. Surface stickiness of drops of carbohydrate and organic acid solutions during convective drying: Experiments and modeling.
Drying Technology, 21(5):839873, January 2003. doi:10.1081/DRT-120021689.
B. Adhikari, T. Howes, B. R. Bhandari, and V. Troung. Effect of addition of maltodextrin
on drying kinetics and stickiness of sugar and acid-rich foods during convective drying:
experiments and modelling. Journal of Food Engineering, 62(1):5368, March 2004.
doi:10.1016/S0260-8774(03)00171-7.
B. Adhikari, T. Howes, and B. R. Bhandari. Use of solute fixed co-ordinate system and
method of lines for prediction of drying kinetics and surface stickiness of single droplet
during convective drying. Chemical Engineering and Processing, 46(5):405419, May
2007. doi:10.1016/j.cep.2006.06.018.
Aerosil. Product information: Aerosil 200F. Product data sheet, Evonik Industries, May
2008. URL www.aerosil.com.
L. Alamilla-Beltrn, J. J. Chanona-Prez, A. R. Jimnez-Aparicio, and G. F. GutirrezLpez. Description of morphological changes of particles along spray drying. Journal of
Food Engineering, 67(1-2):179184, March 2005. doi:10.1016/j.jfoodeng.2004.05.063.
W. F. Ames. Numerical Methods for Partial Differential Equations. Computer Science and
Applied Mathematics. Academic Press, New York, San Francisco, 2nd edition, 1977.
K. Annamalai, W. Ryan, and S. Chandra. Evaporation of multicomponent drop arrays.
Journal of Heat Transfer, 115(3):707716, August 1993. doi:10.1115/1.2910742.
P. W. Appel. Modern methods of detergent manufacture. Journal of Surfactants and Detergents, 3(3):395405, July 2000. doi:10.1007/s11743-000-0144-x.
R. Asmatulu. Enhancement of the dewetability characteristics of fine silica particles. Turkish
Journal of Engineering and Environmental Sciences, 26(6):513520, November 2002.
T. O. K. Audu and G. V. Jeffreys. The drying of drops of particulate slurries. Transactions
of the Institution of Chemical Engineers, 53:165172, 1975.
C. G. J. Baker and K. A. McKenzie. Energy consumption of industrial spray dryers. Drying
Technology, 23(1-2):365386, February 2005. doi:10.1081/DRT-200047665.
220

Bibliography
A. S. Barker. Infrared absorption of localized longitudinal-optical phonons. Physical Review
B, 7(6):25072520, March 1973. doi:10.1103/PhysRevB.7.2507.
A. E. Bayly. Crutcher mix sorption experiments. Personal Communication, 2007.
A. E. Bayly. Crutcher mix particle size distribution. Personal Communication, 2008.
A. E. Bayly, P. Jones, M. Groombridge, and C. McNally. Airflow patterns in a countercurrent spray drying tower - simulation and measurement. In Proceedings of the 14th
International Drying Symposium, volume B, pages 775781, So Paulo, Brazil, August
2004.
R. Benzi. Getting a grip on turbulence. Science, 301(5633):605606, August 2003.
doi:10.1126/science.1087141.
R. Benzi, S. Succi, and M. Vergassola. The lattice Boltzmann equation: theory and
applications. Physics Reports, 222(3):145197, December 1992. doi:10.1016/03701573(92)90090-M.
H. E. Bergna. Colloid Chemistry of Silica: An Overview, volume 131 of Surfactant Science
Series, chapter 3, pages 942. CRC Press, Taylor and Francis Group, 6000 Broken Sound
Parkway NW, Suite 300 Boca Raton, FL 33487-2742, 2006.
L. K. Bieniasz. High order accurate, one-sided finite-difference approximations to concentration gradients at the boundaries, for the simulation of electrochemical reactiondiffusion problems in one-dimensional space geometry. Computational Biology and Chemistry, 27(3):315325, July 2002. doi:10.1016/S1476-9271(02)00079-8.
V. S. Birchal and M. L. Passos. Modeling and simulation of drying milk emulsion in spray
dryers. Brazilian Journal of Chemical Engineering, 22(2):293302, April-June 2005.
V. S. Birchal, L. Huang, A. S. Mujumdar, and M. L. Passos.
Spray dryers: Modelling and simulation. Drying Technology, 24(3):359371, April 2006.
doi:10.1080/07373930600564431.
R. B. Bird, W. E. Stewart, and E. N. Lightfoot. Transport Phenomena. John Wiley and Sons,
1st edition, 1960.
S. Blei and M. Sommerfeld. Consideration of particle interactions in spray dryer modelling:
An extended Euler-Lagrange approach. In Proceedings of the 15th International Drying
Symposium (IDS 2006), pages 307314, August 2006.
F. Bortolotti, L. Pietrantoni, P. Swinkels, and M. Waas. Detergent compositions and process
for preparing them. US Patent: US5160657, 1992. URL http://v3.espacenet.com/
origdoc?DB=EPODOC\&IDX=US5160657\&F=0\&QPN=US5160657.
R. N. Bracewell. Numerical transforms.
doi:10.1126/science.248.4956.697.

Science, 248(4956):697704, May 1990.

R. N. Bracewell. The Fourier Transform and Its Applications. McGraw-Hill, 3rd edition,
1999.

221

CSH

Bibliography
Y. L. Bray and M. Prat. Three-dimensional pore network simulation of drying in capillary porous media. International Journal of Heat and Mass Transfer, 42(22):42074224,
November 1999. doi:10.1016/S0017-9310(99)00006-X.
P. R. Brazier-Smith, S. G. Jennings, and J. Latham. The interaction of falling water droplets:
Coalescence. Proceedings of the Royal Society of London. Series A, Mathematical and Physical
Sciences, 326(1566):393408, January 1972. URL http://www.jstor.org/stable/
78012.
L. J. Briggs. Limiting negative pressure of water. Journal of Applied Physics, 21(7):721722,
July 1950. doi:10.1063/1.1699741.
C. J. Brinker. SolGel Processing of Silica, volume 131 of Surfactant Science Series, chapter 47,
pages 615635. CRC Press, Taylor and Francis Group, 6000 Broken Sound Parkway
NW, Suite 300 Boca Raton, FL 33487-2742, 2006.
H. J. H. Brouwers.
Particle-size distribution and packing fraction of geometric random packings.
Physical Review E, 74(031309), September 2006.
doi:10.1103/PhysRevE.74.031309.
S. Bryant and A. Johnson. Wetting phase connectivity and irreducible saturation in simple
granular media. Journal of Colloid and Interface Science, 263(2):572579, July 2003.
doi:10.1016/S0021-9797(03)00371-0.
J. R. Cash. Efficient numerical methods for the solution of stiff initial-value problems and
differential algebraic equations. Proceedings of the Royal Society A: Mathematical, Physical
and Engineering Sciences, 459(2032):797815, April 2003. doi:10.1098/rspa.2003.1130.
J. Chaloud, J. Martin, and J. Baker. Fundamentals of spray-drying detergents. Chemical
Engineering Progress, 53(12):593596, 1957.
D. H. Charlesworth and W. R. Marshall. Evaporation from droplets containing dissolved
solids. AIChE Journal, 6(1):923, 1960. doi:10.1002/aic.690060104.
H. Chen, S. Kandasamy, S. Orszag, R. Shock, S. Succi, and V. Yakhot. Extended Boltzmann kinetic equation for turbulent flows. Science, 301(5633):633636, August 2003.
doi:10.1126/science.1085048.
X. Chen and G. Xie. Fingerprints of the drying behaviour of particulate or thin layer food
materials established using a reaction engineering model. Food and Bioproduct Processing,
75(4):213222, December 1997. doi:10.1205/096030897531612.
X. Chen, W. Pirini, and M. Ozilgen. The reaction engineering approach to modelling
drying of thin layer of pulped kiwifruit flesh under conditions of small Biot numbers.
Chemical Engineering and Processing, 40(4):311320, July 2001. doi:10.1016/S02552701(01)00108-8.
X. D. Chen and S. X. Q. Lin. The reaction engineering approach to modelling drying
of milk droplets. In Proceedings of the 14th International Drying Symposium, volume C,
pages 16441651, So Paulo, Brazil, August 2004.

222

CSH

Bibliography
X. D. Chen and X. Peng. Modified Biot number in the context of air drying of small moist
porous objects. Drying Technology, 23(1-2):83103, February 2005. doi:10.1081/DRT200047667.
H. Cheong, G. Jeffreys, and C. Mumford.
A receding interface model for
the drying of slurry droplets. AIChE Journal, 32(8):13341346, August 1986.
doi:10.1002/aic.690320811.
A. Cheyne, I. Wilson, and J. Bridgwater. Investigation of structural and processing properties of spray dried detergents. c1 - powder morphology and structure. Technical Report C1, University of Cambridge, Department of Chemical Engineering, Pembroke St.,
Cambridge. CB2 3RA, January 2002.
T. Chung. Computational Fluid Dynamics. Cambridge University Press, 1st edition, 2002.
K. Clement, A. Hallstrm, H. Dich, C. Le, J. Mortensen, and H. Thomsen. On the
dynamic behaviour of spray dryers. Chemical Engineering Research and Design, 69:245
252, 1991.
R. Clift, J. R. Grace, and M. E. Weber. Bubbles, Drops and Particles. Academic Press, 1978.
J. M. Coulson, J. F. Richardson, J. R. Backhurst, and J. H. Harker. Coulson and Richardsons
Chemical Engineering: Particle Technology and Separation Processes v. 2. Coulson and
Richardsons Chemical Engineering. Butterworth-Heinemann Ltd, 4th edition, 1996.
J. Crank. Two methods for the numerical solution of moving-boundary problems in diffusion and heat flow. The Quarterly Journal of Mechanics and Applied Mathematics, 10(2):
220231, 1957. doi:10.1093/qjmam/10.2.220.
J. Crank. Free and Moving Boundary Problems. Oxford University Press, 1st edition, 1996.
C. J. Cremers and R. C. Birkebak. Application of the Abel integral equation to spectrographic data. Applied Optics, 5(6):10571064, June 1966. URL http://ao.osa.org/
abstract.cfm?URI=ao-5-6-1057.
C. Crowe. Modelling spray-air contact in spray-drying systems, volume 1, pages 6399. Hemisphere Publishing Corporation, New York, 1980.
C. T. Crowe, M. P. Sharma, and D. E. Stock. The particle-source-in-cell (psi-cell) model for
gas-droplet flows. Transactions of the ASME Journal of Fluids Engineering, 99(2):325332,
June 1977.
E. L. Cussler. Diffusion, Mass Transfer in Fluid Systems. Cambridge University Press, 2nd
edition, 1997.
N. Dalmaz, H. O. Ozbelge, A. N. Eraslan, and Y. Uludag. Heat and mass transfer mechanisms in drying of a suspension droplet: A new computational model. Drying Technology,
25(2):391400, February 2007.
W. H. de Groot, L. Adami, and G. F. Moretti. Manufacture of Modern Detergent Powders.
Herman de Groot Academic Publisher: Wassenaar, 1995.

223

CSH

Bibliography
A. A. U. de Souza and S. Whitaker. The modelling of a textile dyeing process utilizing the
method of volume averaging. Brazilian Journal of Chemical Engineering, 20(4):445453,
October-December 2003a. doi:10.1590/S0104-66322003000400011.
S. M. A. G. U. de Souza and S. Whitaker. Mass transfer in porous media with heterogeneous
chemical reaction. Brazilian Journal of Chemical Engineering, 20(2), April-June 2003b.
doi:10.1590/S0104-66322003000200013.
J. A. Dirksen and T. A. Ring. Fundamentals of crystallization: Kinetic effects on particle size
distributions and morphology. Chemical Engineering Science, 46(10):23892427, 1991.
doi:10.1016/0009-2509(91)80035-W.
A. A. Dolinsky. High-temperature spray drying. Drying Technology, 19(5):785806, May
2001. doi:10.1081/DRT-100103770.
J. Douglas and T. M. Gallie. On the numerical integration of a parabolic differential equation subject to a moving boundary condition. Duke Mathematical Journal, 22(4):557
571, 1955. doi:10.1215/S0012-7094-55-02262-6.
T. M. El-Sayed, D. A. Wallack, and C. J. King. Changes in particle morphology during
drying of drops of carbohydrate solutions and food liquids. 1. Effect of composition and
drying conditions. Industrial Engineering and Chemical Research, 29(12):2346 2354,
December 1990. doi:10.1021/ie00108a007.
T. Elperin and B. Krasovitov. Evaporation of liquid droplets containing small solid particles. International Journal of Heat and Mass Transfer, 38(12):22592267, August 1995.
doi:10.1016/0017-9310(94)00337-U.
A. Erriguible, P. Bernada, F. Couture, and M. A. Roques. Modeling of heat and mass transfer
at the boundary between a porous medium and its surroundings. Drying Technology, 23
(3):455472, March 2005. doi:10.1081/DRT-200054119.
M. Eslamian and N. Ashgriz. Evaporation and evolution of suspended solution droplets
at atmospheric and reduced pressures. Drying Technology, 25(6):9991010, June 2007.
doi:10.1080/07373930701394977.
M. R. Etzel, S.-Y. Suen, S. L. Halverson, and S. Budijono. Enzyme inactivation in a
droplet forming a bubble during drying. Journal of Food Engineering, 27(1):1734, 1996.
doi:10.1016/0260-8774(94)00078-N.
L. Farber, G. I. Tardos, and J. N. Michaels. Evolution and structure of drying material
bridges of pharmaceutical excipients: studies on a microscope slide. Chemical Engineering
Science, 58(19):45154525, October 2003. doi:10.1016/j.ces.2003.07.003.
M. Farid. A new approach to modelling of single droplet drying. Chemical Engineering
Science, 58(13):29852993, July 2003. doi:10.1016/S0009-2509(03)00161-1.
G. Ferrari, G. Meerdink, and P. Walstra. Drying kinetics for a single droplet of skim-milk.
Journal of Food Engineering, 10(3):215230, 1989. doi:10.1016/0260-8774(89)900277.

224

CSH

Bibliography
D. Fletcher, B. Guo, D. Harvie, T. Langrish, J. Nijdam, and J. Williams. What is important
in the simulation of spray dryer performance and how do current CFD models perform?
In Third International Conference on CFD in the Minerals and Process Industries, pages
357364. CSIRO Australia, CSIRO Australia, December 2003.
D. A. Frank-Kamenetskii. Diffusion and Heat Transfer in Chemical Kinetics. Plenum Press,
2nd edition, 1969.
M. Frenklach. Method of moments with interpolative closure. Chemical Engineering Science,
57(12):22292239, June 2002. doi:10.1016/S0009-2509(02)00113-6.
D. D. Frey and C. J. King. Experimental and theoretical investigation of foam-spray
drying. 1. Mathematical model for the drying of foams in the constant-rate period.
Journal of Industrial and Engineering Chemisty Fundamentals, 25(4):723 730, 1986.
doi:10.1021/i100024a041.
N. Frssling. ber die Verdunstung fallender Tropfen. Gerlands Beitrage zur Geophysik, 52
(170), 1938.
W. H. Gauvin and S. Katta. Basic concepts of spray dryer design. AIChE Journal, 22(4):
713724, July 1976. doi:10.1002/aic.690220413.
A. Gerstlauer, S. Motz, A. Mitrovic, and E.-D. Gilles. Development, analysis and validation of population models for continuous and batch crystallizers. Chemical Engineering
Science, 57(20):43314327, October 2002. doi:10.1016/S0009-2509(02)00348-2.
C. J. Gibbins. A survey and comparison of relationships for the determination of the
saturation vapour pressure over plane surfaces of pure water and of pure ice. Annales
Geophysicae, 8:859886, 1990.
J. A. Goff. Saturation pressure of water on the new Kelvin temperature scale. In Transactions
of the American Society of Heating and Ventilating Engineers, pages 347354, 1957.
J. A. Goff and S. Gratch. Low-pressure properties of water from -160 to 212F. In 52nd
Annual Meeting of the American Society of Heating and Ventilating Engineers, pages 95
122, New York, 1946. The American Society of Heating and Ventilating Engineers.
R. A. Goffredi and E. J. Crosby. Limiting analytical relationships for prediction of spray
dryer performance. Industrial and Engineering Chemistry Process Design and Development,
22(4):665 672, 1983. doi:10.1021/i200023a021.
C. G. Greenwald and C. J. King. The effects of design and opetating conditions on particle
morphology for spray-dried foods. Journal of Food Process Engineering, 4(3):171187,
July 1981. doi:10.1111/j.1745-4530.1981.tb00254.x.
C. G. Greenwald and C. J. King. The mechanism of particle expansion in spray drying of
foods. AIChE Symposium Series, 78(218):101110, 1982.
J. D. Griffith. The Drying and Absorption Properties of Surfactant Granules. PhD thesis,
University of Cambridge, 2008.

225

CSH

Bibliography
J. D. Griffith, A. E. Bayly, and M. L. Johns. Evolving micro-structures in drying detergent
pastes quantified using NMR. Journal of Colloid and Interface Science, 315(1):223229,
November 2007. doi:10.1016/j.jcis.2007.06.050.
J. D. Griffith, A. E. Bayly, and M. L. Johns. Magnetic resonance studies of detergent drop drying. Chemical Engineering Science, 63(13):34493456, July 2008.
doi:10.1016/j.ces.2008.03.043.
V. M. Gunko, V. I. Zarko, V. V. Turov, E. F. Voronin, I. F. Mironyuk, and A. A. Chuiko.
Structural and Adsorptive Characteristics of Fumed Silicas in Different Media, volume 131
of Surfactant Science Series, chapter 38, pages 499530. CRC Press, Taylor and Francis
Group, 6000 Broken Sound Parkway NW, Suite 300 Boca Raton, FL 33487-2742, 2006.
J. J. Guo and J. A. Lewis. Aggregation effects on the compression flow properties and drying
behaviour of colloidal silica suspensions. Journal of the American Ceramic Society, 82(9):
23452358, September 1999. doi:10.1111/j.1151-2916.1999.tb02090.x.
E. Hairer and G. Wanner. Solving Ordinary Differential Equations II. Stiff and DifferentialAlgebraic Problems, volume 14 of Springer Series in Computational Mathematics. Springer
Verlag, Berlin, Heidelberg, New York, second revised edition, 1996.
T. C. Hales. The sphere packing problem. Journal of Computational and Applied Mathematics, 44(1):4176, December 1992. doi:10.1016/0377-0427(92)90052-Y.
C. S. Handscomb, M. Kraft, and A. E. Bayly. A new model for the drying of
droplets containing suspended solids. Chemical Engineering Science, In Press, 2008a.
doi:10.1016/j.ces.2008.04.051.
C. S. Handscomb, M. Kraft, and A. E. Bayly. A new model for the drying of droplets
containing suspended solids after shell formation. Chemical Engineering Science, In Press,
2008b. doi:10.1016/j.ces.2008.10.019.
P. Hansson. The Sink-Effect in Indoor Materials: Mathematical Modelling and Experimental
Studies. PhD thesis, University of Gvle, 2003.
D. J. E. Harvie, D. F. Fletcher, and T. A. G. Langrish. A computational fluid dynamics
study of a tall-form dryer. Food and Bioproducts Processing, 80(3):163174, September
2002. doi:10.1205/096030802760309188.
J. P. Hecht and C. J. King. Spray drying: Influence of developing drop morphology on drying rates and retention of volatile substances. 1. Single drop experiments. Industrial and
Engineering Chemical Research, 39(6):17561765, June 2000a. doi:10.1021/ie9904652.
J. P. Hecht and C. J. King. Spray drying: Influence of developing drop morphology on
drying rates and retention of volatile substances. 2. Modeling. Industrial and Engineering
Chemical Research, 39(6):17661774, June 2000b. doi:10.1021/ie990464+.
C. Hirsch. Numerical Computation of Internal and External Flows, volume 1: Fundamentals
of Numerical Discretization of Wiley Series in Numerical Methods in Engineering. John
Wiley & Sons, 1st edition, 1992a.

226

CSH

Bibliography
C. Hirsch. Numerical Computation of Internal and External Flows, volume 2: Computational Methods for Inviscid and Viscous Flows of Wiley Series in Numerical Methods in
Engineering. John Wiley & Sons, 1st edition, 1992b.
C. W. Hirt and B. D. Nichols. Volume of fluid (VOF) method for the dynamics
of free boundaries. Journal of Computational Physics, 39(1):201225, January 1981.
doi:10.1016/0021-9991(81)90145-5.
J. D. Hoffman. Relationship between the truncation errors of centered finite-difference
approximations on uniform and nonuniform meshes. Journal of Computational Physics,
46(3):469474, June 1982. doi:10.1016/0021-9991(82)90028-6.
A. L. Horvath. Handbook of Aqueous Electrolyte Solutions Physical Properties, Estimation
and Correlation Methods. Ellis Horwood Series in Physical Chemistry. Ellis Horwood
Limited, Chichester, 1st edition, 1985.
M. J. Hounslow, R. L. Ryall, and V. R. Marshall. A discretized population balance for nucleation, growth and aggregation. AIChE Journal, 34(11):18211832, November 1988.
doi:10.1002/aic.690341108.
L. Huang and A. S. Mujumdar. Development of a new innovative conceptual design for
horizontal spray dryer via mathematical modeling. Drying Technology, 23(6):11691187,
June 2005. doi:10.1081/DRT-200059328.
L. Huang, K. Kumar, and A. Mujumdar. Use of Computational Fluid Dynamics to evaluate alternative spray dryer chamber configurations. Drying Technology, 21(3):385412,
January 2003. doi:10.1081/DRT-120018454.
L. Huang, K. Kumar, and A. S. Mujumdar. Computational fluid dynamic simulation of
droplet drying in a spray dryer. In Proceedings of the 14th International Drying Symposium,
volume A, pages 326332, So Paulo, Brazil, August 2004.
T. N. Hunter, G. J. Jameson, and E. J. Wanless. Determinatino of contact angles of nanosized silica particles by multi-angle single-wavelength ellipsometry. Australian Journal of
Chemistry, 60(9):651655, September 2007. doi:10.1071/CH07133.
D. H. Huntington. The influence of the spray drying process on product properties. Drying
Technology, 22(6):12611287, December 2004. doi:10.1081/DRT-120038730.
R. K. Iler. The Chemistry of Silica: Solubility, Polymerization, Colloid and Surface Properties
and Biochemistry. Wiley: New York, September 1979.
A. R. Imre. Binary liquids under absolute negative pressure. Periodica Polytechnica Chemical
Engineering, 44(1):3948, 2000.
F. P. Incropera and D. P. DeWitt. Fundamentals of Heat and Mass Transfer. John Wiley and
Sons, 5th edition, 2002.
H. M. Jaeger and S. R. Nagel. Physics of granular states. Science, 255(5051):15231531,
March 1992. doi:10.1126/science.255.5051.1523.

227

CSH

Bibliography
M. Kadja and G. Bergeles. Modelling of slurry droplet drying. Applied Thermal Engineering,
23(7):829844, May 2003. doi:10.1016/S1359-4311(03)00014-0.
R. Keey. Drying of Loose and Particulate Matter. Hemisphere Publishing Corporation, 1st
edition, 1992.
R. B. Keey. Drying Principles and Practice. Pergamon Press, 2nd edition, 1975.
R. B. Keey. Introduction to Industrial Drying Operations. Pergamon Press, 1st edition, 1978.
S. Kentish, M. Davidson, H. Hassan, and C. Bloore.
Milk skin formation
during drying.
Chemical Engineering Science, 60(3):635646, February 2005.
doi:10.1016/j.ces.2004.08.033.
M. Kohout, A. P. Collier, and F. tepnek. Effective thermal conductivity of wet particle assemblies. International Journal of Heat and Mass Transfer, 47(25):55655574, December
2004. doi:10.1016/j.ijheatmasstransfer.2004.07.031.
M. Kohout, A. P. Collier, and F. tepnek. Mathematical modelling of solvent drying
from a static particle bed. Chemical Engineering Science, 61(11):36743685, June 2006a.
doi:10.1016/j.ces.2005.12.036.
M. Kohout, Z. Grof, and F. tepnek. Pore-scale modelling and tomographic visualisation
of drying in granular media. Journal of Colloid and Interface Science, 299(1):342351,
July 2006b. doi:10.1016/j.jcis.2006.01.074.
K. Kota and T. A. G. Langrish. Fluxes and patterns of wall deposits for skim milk
in a pilot-scale spray dryer. Drying Technology, 24(8):9931001, August 2006.
doi:10.1080/07373930600776167.
T. Kudra. Sticky region in drying - definition and identification. Drying Technology, 21(8):
14571469, January 2003. doi:10.1081/DRT-120024678.
T. Langrish and D. Fletcher. Prospects for the modelling and design of spray dryers in
the 21st century. Drying Technology, 21(2):197215, January 2003. doi:10.1081/DRT120017743.
T. Langrish and T. Kockel. The assessment of a characteristic drying curve for milk powder
for use in computational fluid dynamics modelling. Chemical Engineering Journal, 84(1):
6974, September 2001. doi:10.1016/S1385-8947(00)00384-3.
T. A. G. Langrish and I. Zbicinski. The effect of air inlet geometry and spray cone angle on
the wall deposition rate in spray dryers. Chemical Engineering Research and Design, 72a:
420430, 1994.
J. B. Laurindo and M. Prat. Numerical and experimental network study of evaporation
in capillary porous media. phase distributions. Chemical Engineering Science, 51(23):
51715185, December 1996. doi:10.1016/S0009-2509(96)00341-7.
A. Lee and C. Law. Gasification and shell characteristics in slurry droplet burning. Combustion and Flame, 85(1):7793, May 1991. doi:10.1016/0010-2180(91)90178-E.

228

CSH

Bibliography
R. J. LeVeque and H. C. Yee. A study of numerical methods for hyperbolic conservation
laws with stiff source terms. Journal of Computational Physics, 86(1):187210, January
1990. doi:10.1016/0021-9991(90)90097-K.
R. J. LeVeque. Finite Volume Methods for Hyperbolic Problems. Cambridge Texts in Applied
Mathematics. Cambridge University Press, 1st edition, 2002.
W. K. Lewis. The rate of drying of solid materials. Industrial and Engineering Chemistry, 13
(5):427432, May 1921. doi:10.1021/ie50137a021.
H. Lewy, K. Friedrichs, and R. Courant. ber die partiellen differenzengleichungen der
mathematischen physik. Mathematische Annalen, 100(1):3274, 1928. URL http://
resolver.sub.uni-goettingen.de/purl?GDZPPN002272636.
X. Li and I. Zbicinski. A sensitivity study on CFD modeling of cocurrent spray-drying proces. Drying Technology, 23(8):16811691, August 2005. doi:10.1081/DRT-200065093.
H. Liang, K. Shinohara, H. Minoshimab, and K. Matsushimab. Analysis of constant rate
period of spray drying of slurry. Chemical Engineering Science, 56(6):22052213, March
2001. doi:10.1016/S0009-2509(00)00505-4.
D. R. Lide. Handbook of Chemistry and Physics. CRC Press, 73rd edition, 1992.
S. X. Q. Lin and X. D. Chen. Prediction of air-drying of milk droplet under relatively high
humidity using the reaction engineering approach. Drying Technology, 23(7):13951406,
July 2005. doi:10.1081/DRT-200063486.
S. X. Q. Lin and X. D. Chen. A model for drying of an aqueous lactose droplet using
the reaction engineering approach. Drying Technology, 24(11):1329 1334, November
2006. doi:10.1080/07373930600951091.
S. X. Q. Lin and X. D. Chen. The reaction engineering approach to modelling the cream
and whey protein concentrate droplet drying. Chemical Engineering and Processing, 46
(5):437443, May 2007. doi:10.1016/j.cep.2006.05.021.
H. Maris and S. Balibar. Negative pressures and cavitation in liquid helium. Physics Today,
53(2):2934, February 2000. doi:10.1063/1.882962.
W. R. Marshall, Jr. and E. Seltzer. Principles of spray drying part I - fundamentals of
spray-dryer operation. Chemical Engineering Progress, 46(10):501508, October 1950a.
W. R. Marshall, Jr. and E. Seltzer. Principles of spray drying part II - elements of spray-dryer
design. Chemical Engineering Progress, 46(11):575584, November 1950b.
K. Masters. Spray Drying Handbook. Longman Scientific and Technical, UK, 5th edition,
1992.
H. Masuda, K. Higashitani, and H. Yoshida. Powder Technology Handbook. CRC Press, 3rd
edition, 2006.

229

CSH

Bibliography
R. P. Mayer and R. A. Stowe. Packed uniform sphere model for solids: Interstitial access
opening sizes and pressure deficiencies for wetting liquids with comparison to reported
experimental results. Journal of Colloid and Interface Science, 294(1):139150, February
2006. doi:10.1016/j.jcis.2005.07.005.
T. Metzger, M. Kwapinska, M. Peglow, G. Saage, and E. Tsotsas. Modern modelling methods in drying. Transport in Porous Media, 66(1-2):103120, January 2007.
doi:10.1007/s11242-006-9025-z.
M. Mezhericher, A. Levy, and I. Borde. Theoretical drying model of single droplets containing insoluble or dissolved solids. Drying Technology, 25(6):10251032, June 2007.
doi:10.1080/07373930701394902.
E. E. Michaelides. Particles, Bubbles and Drops Their Motion, Heat and Mass Transfer.
World Scientific, 2006.
H. Minoshima, K. Matsushima, H. Liang, and K. Shinohara. Basic model of spray
drying granulation. Journal of Chemical Engineering of Japan, 34(4):472478, 2001.
doi:10.1252/jcej.34.472.
H. Minoshima, K. Matsushima, H. Liang, and K. Shinohara. Estimation of diameter of
granule prepared by spray drying of slurry with fast and easy evaporation. Journal of
Chemical Engineering of Japan, 35(9):880885, 2002. doi:10.1252/jcej.35.880.
M. M. Montazer-Rahmati and S. H. Ghafele-Bashi. Improved differential modeling and
performance simulation of slurry spray dryers as verified by industrial data. Drying Technology, 25(9):14511462, September 2007. doi:10.1080/07373930701536817.
W. D. Murray and F. Landis. Numerical and machine solution of transient heat conduction
problems involving phase change. Transactions of the ASME, Series C. Journal of Heat
Transfer, 81:106112, 1959.
NAG Fortran Library Routine Document: D03PFF. NAG.
NAG Fortran Library Routine Document: E02BEF. NAG.
NAG. Numerical Algorithms Group. Website, retrieved 2008. URL http://www.nag.
co.uk.
S. Nagata. Mixing: Principles and Applications. Kodansha Scientific Books. Tokyo: Kodansha, 1st edition, 1975. a Halsted Press book.
S. Neic and J. Vodnik. Kinetics of droplet evaporation. Chemical Engineering Science, 46
(2):527537, 1991. doi:10.1016/0009-2509(91)80013-O.
D. Oakley and R. Bahu. Spray/gas mixing behaviour within spray dryers. In A. Mujumdar
and I. Filkova, editors, Drying 91, pages 303313, Amsterdam, 1991. Elsevier.
D. E. Oakley. Scale-up of spray dryers with the aid of computational fluid dynamics. Drying
Technology, 12(1-2):217235, 1994. doi:10.1080/07373939408959954.
D. E. Oakley. Produce uniform particles by spray drying. Chemical Engineering Progress, 10
(1):4854, 1997.
230

CSH

Bibliography
D. E. Oakley. Spray dryer modeling in theory and practice. Drying Technology, 22(6):
13711402, December 2004. doi:10.1081/DRT-120038734.
G. Oberman, T. Farrell, and E. Sizgek. Drying of a liquid droplet suspended in its own
vapour. School of Mathematical Sciences, QUT, Brisbane, Australia, September 2004.
F. M. Orr, Jr., R. A. Brown, and L. E. Scriven. Three-dimensional menisci: Numerical
simulation by finite elements. Journal of Colloid and Interface Science, 60(1):137147,
January 1977. doi:10.1016/0021-9797(77)90264-8.
L. Ozmen and T. Langrish. A study of the limitations to spray dryer outlet performance.
Drying Technology, 21(5):895917, January 2003a. doi:10.1081/DRT-120021691.
L. Ozmen and T. A. G. Langrish. An experimental investigation of the wall deposition of
milk powder in a pilot-scale spray dryer. Drying Technology, 21(7):12531272, January
2003b. doi:10.1081/DRT-120023179.
C. Palencia, J. Nava, E. Herman, G. Rodrguez-Jimenes, and M. Garca-Alvarado. Spray
dryer dynamic modeling with a mechanistic model. Drying Technology, 20(3):569586,
March 2002. doi:10.1081/DRT-120002818.
S. E. Papadakis and C. J. King. Air temperature and humidity profiles in spray drying. 1.
features predicted by the particle source in cell model. Industrial and Engineering Chemisty
Research, 27(11):21112116, November 1988. doi:10.1021/ie00083a026.
S. V. Pennington and M. Berzins. New NAG library software for first-order partial differential equations. ACM Transactions on Mathematical Software, 20(1):6369, March 1994.
doi:10.1145/174603.155272.
P. Perr and I. W. Turner. A 3-d version of transpore: a comprehensive heat and mass transfer
computational model for simulating the drying of porous media. International Journal
of Heat and Mass Transfer, 42(24):45014521, December 1999. doi:10.1016/S00179310(99)00098-8.
R. H. Perry and D. W. Green. Perrys Chemical Engineers Handbook. McGraw Hill, 7th
edition edition, 1997.
J. Phillips. The NAG Library: A Beginners Guide. Clarendon Press, Oxford, 1986.
E. Platzer and M. Sommerfeld. Modelling of turbulent atomisation with and Euler/Euler
approach including the drop size prediction. In 9th International Conference on Liquid
Atomization and Spray Systems, number Paper 2-3, Sorrento, Italy, 2003. ICLASS.
G. O. Porras, F. Couture, and M. Roques. An convection-diffusion model for convection
drying of a shrinking media composed of a binary liquid. Drying Technology, 25(7-8):
12151227, July 2007. doi:10.1080/07373930701438600.
M. Prat. Recent advances in pore-scale models for drying of porous media. Chemical Engineering Journal, 86(1-2):153164, February 2002. doi:10.1016/S1385-8947(01)002832.

231

CSH

Bibliography
M. S. Rahman. Food Properties Handbook (Contemporary Food Science). CRC Press Inc, 1st
edition, 1995.
D. Ramkrishna. Population Balances: Theory and Application to Particulate Systems in Engineering. Academic Press, 1st edition, 2000.
W. E. Ranz and W. R. Marshall. Evaporation from drops part I. Chemical Engineering
Progress, 48(3):1416, 1952.
J. A. Rard and D. G. Miller. The mutual diffusion coefficients of Na2 SO4 H2 O and
MgSO4 H2 O at 25 C from Rayleigh interferometry. Journal of Solution Chemistry, 8
(10):755766, October 1979. doi:10.1007/BF00648779.
D. Reay. Fluid flow, residence time simulation and energy efficiency in industrial dryers.
In M. Roques, editor, Proceedings of the 6th International Drying Symposium, volume 1,
Versailles, France, 1988. keynote address.
W. O. Roberts. Manufacturing and Applications of Water-Borne Colloidal Silica, volume 131
of Surfactant Science Series, chapter 12, pages 131175. CRC Press, Taylor and Francis
Group, 6000 Broken Sound Parkway NW, Suite 300 Boca Raton, FL 33487-2742, 2006.
P. L. Roe. Approximate Riemann solvers, parameter vectors and difference schemes.
Journal of Computational Physics, 43(2):357372, October 1981. doi:10.1016/00219991(81)90128-5.
G. F. C. Rogers and Y. R. Mayhew. Thermodynamic and Transport Properties of Fluids: S. I.
Units. Wiley Blackwell, 5th edition, 1994.
D. Rosenblatt, S. B. Marks, and R. L. Pigford. Kinetics of phase transitions in the system
sodium sulfate-water. Industrial and Engineering Chemistry Fundamentals, 23(2):143
147, May 1984. doi:10.1021/i100014a002.
R. Rokar and V. Kmetec. Evaluation of the moisture sorption behaviour of several excipients by BET, GAB and microcalorimetric approaches. Chemical and Pharmaceutical
Bulletin, 53(6):662665, 2005. doi:10.1248/cpb.53.662.
H. Salman and M. Soteriou. Lagrangian simulation of evaporating droplet sprays. Physics
of Fluids, 16(12):46014622, December 2004. doi:10.1063/1.1809132.
Y. Sano and R. B. Keey. The drying of a spherical partical containing colloidal material into
a hollow sphere. Chemical Engineering Science, 37(6):881889, 1982. doi:10.1016/00092509(82)80176-0.
M. Schatzman. Numerical Analysis A Mathematical Introduction. Clarendon Press, Oxford, 1st edition, 2002. Translated from an earlier French edition.
G. W. Scherer. Theory of drying. Journal of the American Ceramic Society, 73(1):314,
January 1990. doi:10.1111/j.1151-2916.1990.tb05082.x.
G. W. Scherer. Bending of gel beams: method for characterizing elastic properties and
permeability. Journal of Non-Crystalline Solids, 142:1835, 1992. doi:10.1016/S00223093(05)80003-1.
232

CSH

Bibliography
G. W. Scherer. Hydraulic radius and mesh size of gels. Journal of Sol-Gel Science and
Technology, 1(3):285291, January 1994. doi:10.1007/BF00486171.
H. Schlichting. Boundary Layer Theory. McGraw-Hill, New York, 7th edition, 1979.
E. U. Schlnder. Drying of porous material during the constant and the falling rate period:
A critical review of exisiting hypotheses. Drying Technology, 22(6):15171532, December
2004. doi:10.1081/DRT-120038738.
P. Seydel, J. Blmer, and J. Bertling. Modeling particle formation at spray drying using population balances. Drying Technology, 24(2):137146, March 2006.
doi:10.1080/07373930600558912.
V. Shabde, S. Emets, U. Mann, K. Hoo, N. Carlson, and G. Gladysz. Modeling a hollow micro-particle production process. Computers and Chemical Engineering, 29(11-12):
24202428, 2005.
T. K. Sherwood. The drying of solids I. Industrial and Engineering Chemistry, 21(1):
1216, January 1929a. doi:10.1021/ie50229a004.
T. K. Sherwood. The drying of solids II. Industrial and Engineering Chemistry, 21(10):
976980, October 1929b. doi:10.1021/ie50238a022.
T. K. Sherwood. The drying of solids III. mechanism of the drying of pulp
and paper. Industrial and Engineering Chemistry, 22(2):132136, February 1930.
doi:10.1021/ie50242a009.
B. Shi and R. W. Rousseau. Crystal properties and nucleation kinetics from aqueous solutions of Na2 CO3 and Na2 SO4 . Industrial and Engineering Chemistry Research, 40(6):
15411547, March 2001. doi:10.1021/ie0006559.
M. A. Silva, P. J. A. M. Kerkhof, and W. J. Coumans. Estimation of effective diffusivity in
drying of heterogeneous porous media. Industrial and Engineering Chemistry Research, 39
(5):14431452, May 2000. doi:10.1021/ie990563n.
R. K. Sinnott, J. M. Coulson, and J. F. Richardson. Coulson and Richardsons Chemical
Engineering: Chemical Engineering Design v. 6. Coulson and Richardsons Chemical Engineering. Butterworth-Heinemann Ltd, 3rd edition, 1999.
J. C. Slattery. Flow of viscoelastic fluids through porous media. AIChE Journal, 13(6):
10661071, November 1967. doi:10.1002/aic.690130606.
J. C. Slattery. Single-phase flow through porous media. AIChE Journal, 15(6):866872,
November 1969. doi:10.1002/aic.690150613.
J. C. Slattery. Two-phase flow through porous media. AIChE Journal, 16(3):345352, May
1970. doi:10.1002/aic.690160306.
J. C. Slattery. Advanced Transport Phenomena. Cambridge Series in Chemical Engineering.
Cambridge University Press, 1st edition, 1999.

233

CSH

Bibliography
D. M. Smith, G. W. Scherer, and J. M. Anderson. Shrinkage during drying of silica gel.
Journal of Non-Crystalline Solids, 188(3):191206, October 1995. doi:10.1016/00223093(95)00187-5.
Snapdata. UK laundry detergents 2005 household market industry research report. Market
industry research report, Snapdata International Limited, 2005a.
Snapdata. US laundry detergents 2005 household market industry research report. Market
industry research report, Snapdata International Limited, 2005b.
M. Sommerfeld. Validation of a stochastic lagragian modelling approach for inter-particle
collisions in homogeneous isotropic turblence. International Journal of Multiphase Flow,
27(10):18291858, October 2001. doi:10.1016/S0301-9322(01)00035-0.
D. Southwell and T. Langrish. Observations of flow patterns in a spray dryer. Drying
Technology, 18(3):661685, March 2000. doi:10.1080/07373930008917731.
D. B. Spalding. Convective Mass Transfer. Edward Arnold, 1st edition, 1963.
F. tepnek, M. Marek, and P. M. Adler. The effect of pore-space morphology on the
performance of anaerobic granular sludge particles containing entrapped gas. Chemical
Engineering Science, 56(2):467474, January 2001. doi:10.1016/S0009-2509(00)002505.
J. Straatsma, G. V. Houwelingen, A. E. Steenbergen, and P. D. Jong. Spray drying of food
products: 1. Simulation model. Journal of Food Engineering, 42(2):6772, November
1999. doi:10.1016/S0260-8774(99)00107-7.
S. Succi. The Lattice Boltzmann Equation for Fluid Dynamics and Beyond. Oxford University
Press, June 2001.
S. P. Sullivan, L. F. Gladden, and M. L. Johns. Simulation of power-law fluid flow through
porous media using lattice Boltzmann techniques. Journal of Non-Newtonian Fluid Mechanics, 133(2-3):9198, February 2006. doi:10.1016/j.jnnfm.2005.11.003.
N. R. Tas, P. Mela, T. Kramer, J. W. Berenschot, and A. Van den Berg. Water plugs in
nanochannels under negative pressure. In Proceedings of the 7th International Conference
on Miniaturized Chemical and Biochemical Analysis Systems, pages 1316, October 2003.
S. P. Timoshenko. Theory of Elastic Stability. McGraw-Hill, 1st edition, 1936.
S. Torquato, T. M. Truskett, and P. G. Debenedetti. Is random close packing of
spheres well defined?
Physical Review Letters, 84(10):20642067, March 2000.
doi:10.1103/PhysRevLett.84.2064.
N. Tsapis, E. Dufresne, S. Sinha, C. Riera, J. Hutchinson, L. Mahadevan, and D. Weitz.
Onset of buckling in drying droplets of colloidal suspensions. Physical Review Letters, 94
(1):018302, January 2005. doi:10.1103/PhysRevLett.94.018302.
J. Van der Lijn. Simulation of Heat and Mass Transfer in Spray Drying. PhD thesis, University
of Wageningen, 1976.

234

CSH

Bibliography
J. Van der Lijn, P. J. A. M. Kerkhof, and W. H. Rulkens. Droplet heat and mass transfer
under spray drying conditions. In International Symposium on Heat and Mass Transfer
Problems in Food Engineering, Wageningen, Netherlands, 1972. University of Wageningen.
M. A. van Drunen, H. G. Merkus, B. Scarlett, and G. M. van Rosmalen. Barium sulfate precipitation: Crystallization kinetics and the role of the additive PMAPVS. Particle and Particle Systems Characterization, 13(5):313321, October 1996.
doi:10.1002/ppsc.19960130511.
R. Verdurmen, P. Menn, J. Ritzert, S. Blei, G. Nhumaio, T. S. Srensen, M. Gunsing,
J. Straatsma, M. Verschueren, M. Sibeijn, G. Schulte, U. Fritsching, K. Bauckhage,
C. Tropea, M. Sommerfeld, A. P. Watkins, A. J. Yule, and H. Schnfeldt. Simulation
of agglomeration in spray drying installations: The EDECAD project. Drying Technology, 22(6):14031461, December 2004. doi:10.1081/DRT-120038735.
J. G. P. Verhey. Vacuole formation in spray powder particles 1. air incorporation and bubble
expansion. Netherlands Milk and Dairy Journal, 26:186202, 1972.
H. K. Versteeg and W. Malalasekera. An Introduction to Computational Fluid Mechanics.
Pearson, Prentice Hall, 2nd edition, 2007.
H.-J. Vogel, J. Tlke, V. P. Schulz, M. Krafczyk, and K. Roth. Comparison of a latticeBoltzmann model, a full-morphology model, and a pore network model for determining
capillary pressure-saturation relationships. Vadose Zone Journal, 4(2):380388, 2005.
doi:10.2136/vzj2004.0114.
D. A. Wallack, T. M. El-Sayed, and C. J. King. Changes in particle morphology during drying of drops of carbohydrate solutions and food liquids. 2. effects of drying
rate. Industrial and Engineering Chemisty Research, 29(12):2354 2357, December 1990.
doi:10.1021/ie00108a008.
D. Walton and C. Mumford. The morphology of spray-dried particles - the effect of process
variables upon the morphology of spray-dried particles. Chemical Engineering Research
and Design, 77:442460, July 1999a. doi:10.1205/026387699526296.
D. Walton and C. Mumford. Spray dried products - characterization of particle
morphology. Chemical Engineering Research and Design, 77:2137, January 1999b.
doi:10.1205/026387699525846.
C. K. Wei, H. T. Davis, E. A. Davis, and J. Gordon. Heat and mass transfer in waterladen sandstone: Convective heating. AIChE Journal, 31(8):13381348, August 1985a.
doi:10.1002/aic.690310813.
C. K. Wei, H. T. Davis, E. A. Davis, and J. Gordon. Heat and mass transfer in waterladen sandstone: Microwave heating. AIChE Journal, 31(5):842848, May 1985b.
doi:10.1002/aic.690310521.
E. W. Weisstein. Abel transform. From MathWorld A Wolfram Web Resource., retrieved
2008. URL http://mathworld.wolfram.com/AbelTransform.html.

235

CSH

Bibliography
S. R. L. Werner, R. L. Edmonds, J. R. Jones, J. E. Bronlund, and A. H. J. Paterson. Single droplet drying: Transition from the effective diffusion model to a modified receding interface model. Powder Technology, 179(3):184189, January 2008a.
doi:10.1016/j.powtec.2007.06.009.
S. R. L. Werner, R. L. Edmonds, J. R. Jones, and A. H. J. Paterson. Single droplet drying.
2. maltodextrin DE5 drying kinetics and mathematical model validation. Submitted for
publication, 2008b.
S. R. L. Werner, R. L. Edmonds, A. H. J. Paterson, J. E. Bronllund, and J. R. Jones. Single
droplet drying. 1. review of major model types, controlling mechanisms and the constant
internal temperature assumption. Submitted for publication, 2008c.
F. E. W. Wetmore and D. J. LeRoy. Principles of Phase Equilibria. International Chemical
Series. McGraw-Hill, 1st edition, 1951.
S. Whitaker. Fluid motion in porous media. Industrial and Engineering Chemistry, 61(12):
1428, December 1969. doi:10.1021/ie50720a004.
S. Whitaker. Simultaneous heat, mass, and momentum transfer in porous media: a theory
of drying. Advances in Heat Transfer, 13:119203, 1977.
S. Whitaker. Heat and mass transfer in granular porous media. Advances in Drying, 1:23,
1980.
F. M. White. Viscous Fluid Flow. McGraw-Hill Higher Education, 3rd edition, 2005.
A. E. Wijlhuizen, P. J. A. M. Kerkhof, and S. Bruin. Theoretical study of the inactivation
of phosphatase during spray drying of skim-milk. Chemical Engineering Science, 34(5):
651660, 1979. doi:10.1016/0009-2509(79)85110-6.
C. R. Wilke and P. Chang. Correlation of diffusion coefficients in dilute solutions. AIChE
Journal, 1(2):264270, 1955. doi:10.1002/aic.690010222.
WMO. World Meteorological Organization. General meteorological standards and recommended practices, appendix a. WMO Technical Regulations, WMO-No. 49, 1988a.
WMO. World Meteorological Organization. General meteorological standards and recommended practices, appendix a corrigendum. WMO Technical Regulations, WMO-No.
49, August 1988b.
D. Won, D. M. Sander, C. Y. Shaw, and R. L. Corsi. Validation of the surface sink model
for sorptive interactions between VOCs and indoor materials. Atmospheric Environment,
35(26):44794488, September 2001. doi:10.1016/S1352-2310(01)00223-0.
M. W. Woo, W. R. W. Daud, A. S. Mujumdar, M. Z. M. Talib, W. Z. Hua,
and S. M. Tasirin. Comparative study of droplet drying models for CFD modelling. Chemical Engineering Research and Design, 86(9):10381048, September 2008.
doi:10.1016/j.cherd.2008.04.003.
A. G. Yiotis, A. K. Stubos, A. G. Boudouvis, I. N. Tsimpanogiannis, and Y. C. Yortsos. Porenetwork modeling of isothermal drying in porous media. Transport in Porous Media, 58
(1-2):6386, January 2005. doi:10.1007/s11242-004-5470-8.
236

CSH

Bibliography
A. G. Yiotis, I. N. Tsimpanogiannis, A. K. Stubos, and Y. C. Yortsos. Pore-network study of
the characteristic periods in the drying of porous media. Journal of Colloid and Interface
Science, 297:738748, May 2006. doi:10.1016/j.jcis.2005.11.043.
M. C. Yuen and L. W. Chen. On drag of evaporating liquid droplets. Combustion Science
and Technology, 14(4-6):147154, 1976. doi:10.1080/00102207608547524.
I. Zbicinski and X. Li. Conditions for accurate CFD modeling of spray-drying process. Drying Technology, 24:11091114, September 2006. doi:10.1080/07373930600778221.
I. Zbicinski and R. Zietara. CFD model of counter-current spray drying processes. In
Proceedings of the 14th International Drying Symposium, volume A, pages 169176, So
Paulo, Brazil, August 2004.
I. Zbicinski, M. Piatkowski, and W. Prajs. Determination of spray-drying kinetics in a
small scale. Drying Technology, 23(8):17511759, August 2005. doi:10.1081/DRT200065197.
Q. Zheng, D. J. Durben, G. H. Wolf, and C. A. Angell. Liquids at large negative pressures:
Water at the homogeneous nucleation limit. Science, 254(5033):829 832, November
1991. doi:10.1126/science.254.5033.829.

237

CSH

Anda mungkin juga menyukai