LAU-THESIS-2023 simulation numérique de poudre pharmaceutique

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

N UMERICAL SIMULATION OF FLUIDIZED BED DRYING OF

PHARMACEUTICAL POWDERS USING A TWO - PHASE MODEL

A Thesis Submitted to the


College of Graduate and Postdoctoral Studies
in Partial Fulfillment of the Requirements
for the degree of Master of Science
in the Department of Chemical and Biological Engineering
University of Saskatchewan
Saskatoon

By
Marcus Lau

©Marcus Lau, June/2023. All rights reserved


Unless otherwise noted, copyright of the material in this thesis belongs to the
author.
P ERMISSION TO U SE

In presenting this thesis in partial fulfilment of the requirements for a Postgraduate degree from the University of
Saskatchewan, I agree that the Libraries of this University may make it freely available for inspection. I further agree
that permission for copying of this thesis in any manner, in whole or in part, for scholarly purposes may be granted by
the professor or professors who supervised my thesis work or, in their absence, by the Head of the Department or the
Dean of the College in which my thesis work was done. It is understood that any copying or publication or use of this
thesis or parts thereof for financial gain shall not be allowed without my written permission. It is also understood that
due recognition shall be given to me and to the University of Saskatchewan in any scholarly use which may be made
of any material in my thesis.
Requests for permission to copy or to make other use of material in this thesis in whole or part should be addressed
to:

Head of the Department of Chemical and Biological Engineering


University of Saskatchewan
57 Campus Drive
Saskatoon, Saskatchewan
Canada
S7N 5C9

Or

Dean
College of Graduate and Postdoctoral Studies
University of Saskatchewan
116 Thorvaldson Building, 110 Science Place
Saskatoon, Saskatchewan S7N 5C9
Canada

i
A BSTRACT

Tablet production in the pharmaceutical industry is a common process for oral ingestion products. Before tablet
production, mixtures of active pharmaceutical ingredients and other excipients are granulated, generally through the
process of wet granulation. The wet granulation process agglomerates mixtures of powder components to create
homogeneous granules, typically using water as a liquid binder. Before the wet granules can be made into tablets, they
have to be dried to an acceptable level. Fluidized beds have been extensively used for the drying of these granules. To
better understand the fluidized bed process, mathematical models have been created to emulate the drying phenomena.
Simplified models, such as phenomenological models, aim to capture the drying characteristics without defing the
more complex kinetics and thermodynamics of the system. A recent approach based on a two-phase model is further
verified in this work using experimental data from a lab scale fluidized bed. The model was examined against the
pharmaceutical powder moisture content and temperature profiles from previous experimental work.
The model is comprised of the mass and energy balances of five distinct sections of the fluidized bed. Powder
moisture and heat transfer are governed by a stagnant gas film in equilibrium with the surrounding gas. The model
shows good correlation with the experimental data. It also displays the general characteristics of pharmaceutical powder
drying, with distinct constant drying rate and falling drying rate periods.
Upon model validation, optimization of the inlet gas parameters is explored. Optimization of the model is imple-
mented, controlling the inlet gas parameters and incorporating a stepwise change during the drying process. It was
found that a single stepwise change has a negligible effect on optimizing the process but is still useful around the end
point of the batch. In addition, optimization results and behaviours are discussed.

ii
ACKNOWLEDGEMENTS

First and foremost, I would like to express my gratitude to both my co-supervisors, Dr. Lifeng Zhang of the
Chemical and Biological Engineering department, and Dr. Raymond Spiteri of the Computer Science department. I’d
like to thank both of them for welcoming me to their respective research groups as well as their continuous support and
direction throughout this project.
I would also like to thank the members of Dr. Zhang’s research group, especially Chen Li, who all provided me with
valuable anecdotes about the experimental setup, as well as allowing me to get acquainted with the physical fluidized
bed dryer.
Lastly, I would like to acknowledge the financial assistance from the University of Saskatchewan and the Natural
Sciences and Engineering Research Council of Canada (NSERC).

iii
C ONTENTS

Permission to Use i

Abstract ii

Acknowledgements iii

Contents iv

List of Tables v

List of Figures vi

List of Abbreviations vii

1 Introduction 1
1.1 Project Motivation and Knowledge Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Literature Review 3
2.1 Tablet Production via Wet Granulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Fluidized Bed Dryers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2.1 Fluidization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.2 Drying Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Fluidized Bed Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3.1 Drying Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.2 Utilization of the Two-Phase Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Validation of Two-Phase Fluidized Bed Drying Model for Pharmaceutical Powders 16


3.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.1 Two-Phase Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.2 Single Particle Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.2.1 Mass Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.2.2 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3.3 Bubble Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.3.3.1 Mass Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3.3.2 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3.4 Dense Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3.4.1 Mass Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3.4.2 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3.5 Vessel Wall and Freeboard Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3.5.1 Mass Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.5.2 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.6 Heat and Mass Transfer Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.7 Model Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.8 Experimental Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.8.1 Applicable Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.4.1 Particle Moisture Content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

iv
3.4.2 Particle Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4.3 Relative Humidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4 Optimization of Fluidized Bed Drying using a Two-Phase Model 38


4.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.3.1 Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.3.2 Energy and Exergy Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3.2.1 Energetic Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3.2.2 Exergetic Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3.3 Inlet Gas Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3.3.1 Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3.3.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3.3.3 Step-Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3.4 Optimization using Matlab . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.4.1 Objective Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.4.1 Energy and Exergy Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.4.2 Five-Variable Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4.2.1 Profit Ratio Bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4.2.2 Profit Ratio Transition Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4.2.3 Step-Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

5 Conclusions and Recommendations 53


5.1 Summary of Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.1.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Appendix A Matlab Code 59


A.1 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
A.2 Model Controller and Post Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
A.3 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

v
L IST OF TABLES

2.1 Summary of modelling methods for fluidized beds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.1 Summary of Transfer Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27


3.2 Summary of Model Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3 Granule Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4 Summary of Constant Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.5 Pulsated Flow General Particle Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.1 Summary of modelling methods for fluidized beds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


4.2 Inlet Gas Parameter Bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Optimization results at Profit Ratio bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4 Optimization results in transition area for optimization function . . . . . . . . . . . . . . . . . . . . . 49

vi
L IST OF F IGURES

2.1 Mechanisms of granulation, adapted from [2] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4


2.2 General fluidized bed dryer configurations and residence type distributions for a) Batch, b) Back-mixed,
and c) Plug Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Bed pressure with increasing inlet gas velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4 Geldart’s classification for wet particles [6] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.5 Fluidization regimes with increasing inlet gas velocity, adapted from Mujumdar . . . . . . . . . . . . 8
2.6 General particle behaviour in fluidized bed drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.7 Rising bubble using three-phase configuration [25] . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.8 Two-phase schematic diagram for continuous (additions in blue) and batch drying . . . . . . . . . . . 14

3.1 Mass and energy transport within model, adapted from [29] . . . . . . . . . . . . . . . . . . . . . . . 19
3.2 Product bowl dimensions [4] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3 Moisture content profiles by dry basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4 Model particle temperature profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.5 Experimental relative humidity example for general trend . . . . . . . . . . . . . . . . . . . . . . . . 35
3.6 Model outlet relative humidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.1 General Two-phase fluidized bed model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40


4.2 Energy Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Exergy Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4 Moisture and temperature profiles of optimization solution at Profit Ratio bounds . . . . . . . . . . . 49
4.5 Moisture and temperature profiles of optimization solution within transition range . . . . . . . . . . . 50
4.6 Objective value vs. step change with high/low configuration . . . . . . . . . . . . . . . . . . . . . . 51

vii
L IST OF A BBREVIATIONS

cg Heat Capacity Coefficient - dry gas


cs Heat Capacity Coefficient - dry particle
cv Heat Capacity Coefficient - water vapour
cw Heat Capacity Coefficient - liquid water
db Bubble Diameter
De f f Effective Diffusion Rate
dp Particle Diameter
Gr Grashof Number
∆Hevap Enthalpy of Evaporation
Hb,c Heat Transfer Coefficient (Dense Phase to Cloud Phase)
Hc,d Heat Transfer Coefficient (Cloud Phase to Dense Phase)
Hd,b Heat Transfer Coefficient (Dense Phase to Bubble Phase)
hd,s f Heat Transfer Coefficient (Dense Phase to Particle)
hd,v Heat Transfer Coefficient (Dense Phase to Vessel)
hr,v Heat Transfer Coefficient (Freeboard Region to Vessel)
hv,a Heat Transfer Coefficient (Vessel to Ambient Air)
I˙ Enthalpy Flow Rate
Kb,c Mass Transfer Coefficient (Bubble Phase to Cloud Phase)
Kc,d Mass Transfer Coefficient (Cloud Phase to Dense Phase)
Kd,b Mass Transfer Coefficient (Dense Phase to Bubble Phase)
ks f ,d Mass Transfer Coefficient (Particle Stagnant Film to Dense Phase)
ṁ Water Flow Rate
M Moisture Content
Me Moisture Content Equilibrium
ms Particle Mass
MR Moisture Ratio
OF Objective Function
P Pressure
Pwv Water Vapour Partial Pressure
Q̇ Heat Flow Rate
Tin Inlet Temperature
Tout Outlet Temperature
Tb Bubble Temperature
Td Dense Phase Temperature
Tp Particle Temperature
Tv Vessel Temperature
vin Inlet Velocity
vm f Minimum Fluidization Velocity
Vb Bubble Control Volume
vb Bubble Velocity
Vf Bubble + Dense Control Volume
xcr Particle Critical Moisture Content Point
xout Outlet Moisture Content
xs f Stagnant Film Moisture Content
xb Bubble Phase Moisture Content
xd Dense Phase Moisture Content
xp Particle Moisture Content
z Height
Z Normalization Scale Factor
δ Bubble Phase Fraction

viii
εm f Voidage Space
ηen Energetic Efficiency
ηex Exergetic Efficiency
φ Particle Sphericity
ρg Gas density
ρs Dry Particle density
ρw Water density
τ Drying Duration

ix
C HAPTER 1
I NTRODUCTION

1.1 Project Motivation and Knowledge Gap


In the pharmaceutical industry, medicinal products are often produced in the form of tablets. Tablets are common
because they are easily adaptable to contain the proper mix of active ingredients, binders, and other constituents to
function in the desired way [1]. Wet granulation is one method used to create tablets. The wet granulation process uses
a liquid binder, usually water, to agglomerate the mix of powders together into granules [2]. Before these granules
can be compacted into tablets, the excess moisture must be removed. Fluidized bed dryers are frequently used for
this purpose. Fluidized beds are proficient at drying due the high mass and energy transfer that can occur within the
gas-solid mixture.
The nature of the pharmaceutical industry has not fully utilized advancements in drying technology in many wet
granulation operations. Many pharmaceutical processes still use batch fluidized bed dryers instead of a continuous
setup despite advancements in drying technology. Furthermore, the fluidized bed dryers have little to no inline mon-
itoring capabilities and often rely on operational knowledge to correctly dry the batch of wet granules. These issues
can be attributed to the regulatory practices in the pharmaceutical industry because strict manufacturing and product
licensing impede process changes or improvements [3]. Therefore, any enhancements to the batch drying process
through less intrusive means may be beneficial.
Many studies have been completed to model the drying process, in particular pharmaceutical granules. In the
literature, work has been done to verify models with experimental data from both lab and industrial scales. Once
verified, the models have been used for sensitivity analysis and scaling efforts to minimize the need for additional
real-world testing. Fluidized bed drying models have also been applied to optimization and optimal control techniques
to implement generic model control for the batch process. However, the focus of these models is usually to control
the drying process to perform in a predetermined profile instead of optimization of parameters. Moreover, research is
sparse for optimal control, with optimization of the inlet gas velocity and temperature profiles only being completed
for rice drying, which has a much higher critical moisture content point than pharmaceuticals, among other differences.

1
1.2 Research Objectives
The first objective of this study is to model an experimental scale fluidized bed dryer for drying pharmaceutical pow-
ders. The model will be validated by the experimental data captured by Milad Taghavivand during his work at the
University of Saskatchewan studying tribocharging phenomena [4]. The second objective is to optimize the drying
process using the experimentally validated model. Optimization will look for optimal operational parameters with
respect to the main operational goals (energy efficiency, drying time).
The thesis is structured as follows. Chapter 2 discusses previous literature and theory, examining fluidized bed
drying, fluidization, and modeling practices. Chapter 3 describes the two-phase model being employed and verifi-
cation with lab scale experimental data. After the two-phase is validated with experimental data, Chapter 4 focuses
on optimization for the drying process. A step-change to the inlet gas velocity and temperature is examined using
optimization methods. Together, Chapter 3 and 4 provide a draft for an upcoming submission for publication. Lastly,
Chapter 5 concludes the work and discusses further research opportunities.

2
C HAPTER 2
L ITERATURE R EVIEW

2.1 Tablet Production via Wet Granulation


In the pharmaceutical industry, much of the tablet manufacturing process is completed using the wet granulation
method. Granulation refers to small particles being brought together to form larger particles, with the original distinct
small particles still distinguishable [2]. Wet granulation is a popular method to granulate material and is characterized
by a liquid binder to promote agglomeration of the particles. The agglomerating liquid, usually water, wets the small
particles and binds the particles together through liquid bridge forces. This step is taken within the tablet production
process for the following benefits [2]:

1. Structural strength of produced granules


2. Improved flow properties
3. Reduction of caking tendencies and segregation into fines
4. Product uniformity for pharmaceutical standards

Granulation begins with the mixing of active pharmaceutical ingredients and other excipients necessary for the
product. The wet granulation process can be defined by the following mechanisms: wetting, nucleation, coalescence
and consolidation, and finally granule attrition or breakage [2]. The solids wetting from the liquid binder encourages
the initial coalescence, or nucleation, of the wet particles. These newly formed granules undergo further collisions with
each other and coalesce and consolidate into even larger granules. Lastly, granules that are weakly formed or deficient
in other manners undergo particle attrition and break back into smaller granules. These four mechanisms Figure 2.1
work concurrently in the wet granulation process, with the rates of each mechanism impacted by solids properties and
equipment selection.

3
Figure 2.1: Mechanisms of granulation, adapted from [2]

There are a variety of wet granulation techniques that can be utilized in pharmaceutical tablet production. The
most preferred method involves a high-shear mixer. Low-shear mixers, such as ribbon mixers, were used initially in
industry, with high-shear mixers taking over in more recent times due to superior control and production [2]. A typical
high-shear mixer is a mechanically generated fluidized bed that has an impeller to stir the material, a chopper to break
apart the material, and a liquid binder feed through a nozzle or injector. The vessel can also be fluidized by an upward
gas flow to enhance mixing. Impeller mixing speeds generally range between 100 and 500 rpm, with chopper speeds
from 1000 and 3000 rpm. This method is often chosen because it creates better quality granules for tablet production
[2]. Other wet granulation methods include batch fluidized beds, single-pot processing, and extrusion/spheronization.
Once the granules are produced, the liquid binder must be removed by drying. Fluidized beds are often the preferred
method, but microwave/vacuum drying has proved to be effective as well. A prominent example of this is the single-pot
process, which granulates and dries the material in one vessel to remove the transfer step for the wet granules to the
drying equipment [3].

2.2 Fluidized Bed Dryers


Fluidized bed dryers are a common piece of equipment for drying solid particles and granules. A vessel containing a
bed of moist material exerts an upwards flow of drying gas through the bed to remove the undesired moisture. When
operated correctly, the gas and solid mixture within the fluidized bed behaves similar to a boiling fluid. Fluidized bed
dryers are employed in many industries other than pharmaceuticals, such as food, biomass, and chemical manufacturing
[5]. Fluidized beds are highly effective for particle sizes from 50 µm to 2000 µm, exhibiting adequate fluidization
characteristics in this range [6]. Fluidized bed dryers are used for drying due to the thorough contact time between the

4
gas and solids, homogeneous mixing, and high thermal and mass transfer conditions [5].
Fluidized bed dryers can be found in both batch and continuous configurations, as well as many hybrid methods.
The batch process is arguably the simplest, with a bed of particles being dried sufficiently and gravity-fed out of the
vessel to be ready for the next batch. Continuous fluidized bed dryers can be categorized based on the residence times
of the particles. Back-mixed dryers recirculate some of the particles, creating a broad range of residence times. In
contrast, plug flow dryers try to constrict the residence time by inducing a continuous solid flow from the inlet to outlet
[7]. The differences can be seen in Figure 2.2, with the residence time distribution displaying the various residence
times for particles inside the bed.

Figure 2.2: General fluidized bed dryer configurations and residence type distributions for a) Batch, b) Back-
mixed, and c) Plug Flow

The fluidization of some solid particle beds can be difficult. Fluidization issues can arise due to certain particle
specifications, as well as cohesive forces caused by the moisture. Many strategies have been designed and investigated
to overcome defluidization. All the strategies involve additional agitation or energy into the fluidized bed to promote
better fluidization regimes for drying. Design concepts include, but are not limited to [5] [7] [8]:

1. Pulsated flow of the drying gas


2. Fluidized bed vessel vibration
3. Electrical or magnetic fields to lower cohesive forces
4. Mechanically assisted fluidization (vessel baffles, stirrers, etc.)
5. Spouted or jet stream flow
6. Fast fluidization or high recirculation rates of particles transported to the top of the vessel

5
2.2.1 Fluidization

Fluidization refers to the process of an upwards flow of fluid causing a bed of solid particles to have fluid-like behaviour.
For drying, the fluidizing fluid is generally air, but other applications can be found for various other gases and liquids.
For fluidization to occur, the fluid must exert enough upwards drag force to the solid particles to offset the gravitational
force [6]. In other words, fluidization of a bed of particles occurs when the drag forces are equal to or greater than the
weight of the bed. Fluidization can be distinguished by observing the fluidized bed pressure drop while increasing the
velocity of the fluidizing fluid. The pressure drop with increasing gas velocity can be seen in Figure 2.3

Figure 2.3: Bed pressure with increasing inlet gas velocity

The pressure drop of the fluidized bed increases as the fluidizing fluid velocity increases. This continues until
a constant pressure drop as seen in Figure 2.3. At this point, the pressure drop is equal to the weight of the solid
particles. The velocity of the fluidizing fluid that can sustain this pressure drop is often known as the minimum
fluidization velocity, vm f . The pressure drop remains the same at velocities higher than vm f because the bed weight
has already been sufficiently supported. At a certain point of increasing past the minimum fluidization velocity, the
pressure decreases as the particles start to be pneumatically transported out the top of the vessel [5]. Bed pressures also
display periodic fluctuations that occur due to the rise and fall of the particles [9].
The fluidization regime depends significantly on the solids being dried. Particles can be categorized by Geldart’s
classification for wet particles into four groups [5]. Geldart’s classification is based on particle size and the density
difference between the particle and fluidizing gas. Figure 2.4 displays the relationship between the two variables.

6
Figure 2.4: Geldart’s classification for wet particles [6]

The classifications can be seen by plotting the density differences against particle size, as seen in Figure 2.4. Each
particle group has their own characteristics and fluidization regime tendencies. Group A particles are in the best size
range for fluidization. These particles fluidize easily and tend keep in a smooth fluidization regime with minimal
bubbling or slugging of gas. Larger particles can be found in Group B, which are in the range of 150 µm and 1000
µm. Group B particles are not as efficient at fluidization as those in Group A, with bubbling or slugging regimes
occurring closer to the minimum fluidization velocity. Group C houses the smallest particles and powders. Group C
particles are considered the most difficult to fluidize and often have defluidization issues. Due to the small particle size,
cohesive forces between the particles are not overcome by the fluidizing gas. This often causes Group C particles to
have channeling of the gas flow because void pathways are created for the gas to bypass the particle bed. Lastly, Group
D contains the largest particles that can be fluidized. At this size, the particles are too large for ideal fluidization, and
large bubbles and slugging regimes are created [6].
The effect of water on fluidization is significant and justifies concern. It is known for pharmaceutical granules to
display Group C behaviour above 20% moisture and become like Group B particles once below 10% moisture [10].
The excess moisture causes water bridging forces between the solid particles and creates larger agglomerates, affecting
flow regimes. A higher superficial gas velocity is necessary to fluidize a wet particle bed and to overcome the additional
water weight. Auxiliary agitation such as pulsation can also be an effective strategy to promote fluidization.
Various fluidization regimes can occur for fluidizing gas through a bed of solid particles, depending on inlet gas
velocity and Geldart classification. Figure 2.5 displays the regimes as the gas velocity increases. Below the minimum
fluidization velocity, the gas flow mostly travels within the void spaces of the bed and does not have enough energy

7
to support fluidization. As the velocity increases past the minimum fluidization velocity, the excess gas travels with
a bubbling, slugging, turbulent, or channeling behaviour, depending on Geldart classification. At even higher gas
velocities, the bed weight cannot keep the particles within the vessel, and the particles are transported out the top of
the fluidized bed. This higher velocity can be used for ”fast fluidization”, characterized by high recirculation rates due
to the pneumatic transport of the particles [6].

Figure 2.5: Fluidization regimes with increasing inlet gas velocity, adapted from Mujumdar

2.2.2 Drying Process

The drying process within fluidized beds is a process of simultaneous heat and mass transfer. Generally, the process
can be split into two periods, the constant drying rate period and falling drying rate period [11]. During the constant
drying rate period, there is external moisture not bound to the particles present. This moisture on the particle surface is
freely available for the drying gas to evaporate into the gas stream. The rate of moisture removal is typically constant in
this period because the abundance of surface water creates an equilibrium of gas saturation within the vessel. Likewise,
the particle temperature during this period also finds an equilibrium value and settles at the wet bulb temperature of the
system. Once the unbounded moisture is removed, the drying process enters the falling drying rate period. This period
is defined by the internal moisture transport of the water found inside the particles. The drying rate is diminished by
the lack of readily available water, and the drying gas starts to warm up the particle bed. The internal moisture goes

8
through a desorption process and diffuses to the surface of the particle to evaporate. Therefore, particle attributes are
the dominating factor during the falling drying rate period.
The amount of unbound and bound water is a significant factor in the drying curve. Pharmaceutical powders
generally have a high amount of unbound surface moisture, and their drying is dominated by the constant drying rate
period. Figure 2.6 displays the typical temperature and moisture curves for porous particles in fluidized bed dryers.
Other material, such as food grains, have more internal moisture and begin the falling drying rate period sooner and
at a higher moisture content. The moisture content point where the falling drying rate period begins is known as the
critical moisture content point. Pharmaceutical materials have a relatively low critical moisture content point in the
range of 5%-20% moisture by dry basis [4], with other materials such as food and biomass at higher values ranging
between 40%-60% moisture by dry basis [12].

Figure 2.6: General particle behaviour in fluidized bed drying

The transport of internal moisture by diffusion is dependent on the material being dried. Pharmaceutical materi-
als generally have uniform diffusion. However, they are susceptible to thermal surface degradation that can impede
moisture transport [13]. Rice, corn, and other food material exhibit a variable diffusion rate depending on the moisture
content and temperature of the bed [14]. These materials can benefit from a tempering stage to increase moisture
transport and removal [15], employing a heating stage to enhance the drying process. Diffusivity coefficients are a
prominent aspect to fluidized bed drying, governing the falling drying rate period. Unfortunately, empirical correla-
tions for diffusivity are highly specific for individual materials and can also change when scaling up to larger fluidized
beds.
Other material factors affect the drying rate as well. Thermal properties [16] and particle shape of the solids material
impact tempering effects and hydrodynamics. Larger particle size distribution ranges require higher gas velocities and

9
smaller batch sizes to obtain adequate fluidization [17]. Dryer design also plays a significant role. The method of
inlet gas distribution (porous, punched, or perforated) affects fluidization [18]. Likewise, vessel shape (rectangular,
cylindrical, constant or varying diameter, etc) also influences the hydrodynamics.

2.3 Fluidized Bed Modelling


The modelling of fluidized beds has been extensively researched in published literature. Models are employed for a
better comprehension of the fluidization process. The difficulty of modelling fluidized beds can be attributed to the
various scales of phenomena occuring simultaneously. For example, the fluidization regimes can be seen at a macro
level, whereas particle-particle collisions occur at much smaller scales. Therefore, models must focus on a particular
scale depending on the research need to provide valuable insight. For modelling, the mathematical description of the
solids and gas within the fluidized bed can be from either a Lagrangian or Eulerian viewpoint. Lagrangian descriptions
discretize the solid or gas phases into individual particles, whereas an Eulerian approach uses continuum equations,
often employing Navier–Stokes theories to characterize the phenomena. With this in mind, there are five main cate-
gories for modelling fluidized beds and solids and gas interactions. These categories are divided based on the type of
mathematical description used for both the gas and solids [19].

Table 2.1: Summary of modelling methods for fluidized beds

Model Description Scale

Discrete Bubble Lagrangian gas phase; Eulerian solid Industrial (10 m)


phase
Two-fluid Eulerian Phases Engineering (1 m)
Unresolved discrete Particle Lagrangian gas phase; Eulerian solid Laboratory (0.1 m)
phase (discretized into sections larger
than solid particles)
Resolved discrete particle Lagrangian gas phase; Eulerian solid Laboratory (0.01 m)
phase (discretized into sections smaller
than solid particles)
Molecular dynamics Lagrangian phases Mesoscopic (0.001 m)

The range of models listed in Table 2.1 displays the various combinations of Lagrangian/Eulerian descriptions for
both phases. The first three model categories are the main techniques found in practice to model fluidized beds. In
the Discrete Bubble method, an emulsion (or dense) phase represents a mixture of gas and solid particles in a Eulerian
grid. The rest of the gas, in the form of bubbles, is described individually using equations for motion [20]. This
allows the model to account for bubble-bubble, bubble-wall, and coalescence interactions, effectively describing the
bubbling phenomena for industrial-scale fluidized beds. In a two-fluid model, both phases are described using Eulerian

10
equations. The two-fluid model needs more parameter specifications because the model does not represent any detailed
interactions between particles or interactions between the particles and gas. However, despite the simplification of the
two-fluid model, many studies have shown good agreement with experimental data, successfully predicting bubble size
formation and voidage space [21], and have displayed adequate results with correlation variables. The two-fluid model
is also a practical choice for modelling the drying process in fluidized beds, accommodating experimental data easily
with various correlating variables. The last prevalent type of model is the unresolved discrete particle model. This
classification characterizes the gas within the fluidized bed with an Eulerian grid and models the interactions with the
individual solid particles. This method is considered unresolved due to an Eulerian grid being magnitudes larger than
the solid particles. The discretization of the gas is not small enough to fully model the particle/gas interactions, with
parameters still needed to describe the interactions.
The last two models in Table 2.1 cover the model techniques that are not normally used for fluidized beds but
still effectively can represent a fluidized system. The lack of usage is mainly due to the small scale that these models
represent and the large amount of computation power to model any suitable experimental or industrial systems. The
resolved discrete particle model is similar to its unresolved counterpart described above but uses an Eulerian grid size
for the gas at least a magnitude smaller than the solid particles within the system. This allows the model to account
for the flow of gas between the particles and characterizes the particle/gas interactions via boundary conditions at the
surface of the particles. In the last model category, fluidized beds can be modelled using a Lagrangian description for
both phases. Both phases are described as particles, using molecular dynamics to determine the individual positions
and velocities of the solid and gas particles.
As stated previously, fluidization regimes at a macro level are affected by the smaller scale interactions within the
fluidized bed. Consequently, scaling is problematic when evaluating fluidized bed models. The main issue is that
fluidization in larger scale vessels have vastly different fluidization patterns and flow regimes in comparison to lab
scale setups. Scale-up practices are comprised of empirical data and correlations, with industrial-sized designs unable
to fully rely on lab-scale results. An equation used to evaluate various scaling for fluidized bed dryers takes the form
of [22]:

∆τ2 (mB/A )2 G1 (Tin − TB )1


Z= = , (2.1)
∆τ1 (mB/A )1 G2 (Tin − TB )2

with Z representing the normalization scale factor and τ representing the time needed for a complete drying run. The
scaling in dependent on three ratios, with mB/A denoting the mass of the bed divided by the bed area, G denoting the
volumetric drying gas velocity, and Tin and TB representing the inlet and bed temperatures respectively. Equation (2.1)
has been further validated with pharmaceutical powders comparing two lab-scale experimental fluidized bed dryers.

2.3.1 Drying Models

Fluidized bed models incorporating drying are generally the more simplistic models due to the complexity caused by
the moisture. Fluidization regimes are greatly affected by the cohesive forces from the excess water in the bed [10].

11
Consequently, modelling particle-particle interactions during the drying process is currently impractical or at the very
least, not necessary. The motivation for modelling the drying process is usually to understand a particular drying
profile, so complex models are not always needed. With this in mind, there are still a variety of models with differing
complexities developed to describe the drying process.
Thin-layer drying modelling has been applied to many drying profiles with food material. Thin-layer models
describe the drying process as the drying of a thin layer of material exposed entirely to an airstream [15]. These models
can be theoretical, semi-theoretical, or empirical. Theoretical models are generally based on Fick’s second law of
diffusion, adapted to cartesian, cylindrical, or spherical structure. Fick’s second law of diffusion can be expressed as:
dM
= Deff O2 M , (2.2)
dt
with M representing the moisture content of the material and Deff representing the effective moisture diffusivity coeffi-
cient. Fick’s second law of diffusion is effective at describing biological drying, such as food or agricultural materials,
because they often possess a significant falling drying rate period. The equation uses an effective moisture diffusivity
coefficient so the mechanism of moisture transport is not needed to be known.
Semi-theoretical thin-layer models either apply Fick’s second law of diffusion or Newton’s law of cooling, with
simplifications or correlating variables to account for various drying profiles. An example of a simplistic diffusion
related thin-layer model is the Henderson and Pabis model [15]. This model is a single term general solution of Fick’s
second law and can be presented as:

MR = aexp(−kt) , (2.3)

with MR denoting a normalized moisture ratio, with a and k used as correlating variables. These correlating variables
do not have any physical meaning and serve as a method to fit experimental data. Newton’s law of cooling suggests that
the drying process is analogous to a solid body immersed in a cooler fluid. The Lewis model is the simplest cooling law
model [15] and determines the drying rate using the difference between the material’s actual and equilibrium moisture
content. The equation can be seen as:

dM
= −k(M − Me ) (2.4)
dt
with the addition of the drying constant, k, and Me representing the equilibrium moisture content at a given temperature.
By setting a boundary condition for the initial moisture content, the analytical solution for the Lewis model can be seen
as:

MR = exp(−kt) . (2.5)

In addition to these semi-theoretical thin-layer models, there are numerous adaptations and improvements made to
both Fick-based and Newton-based models. Generally, additional correlating variables are introduced, or changes to
the equation are made to account for specific drying cases.

12
Empirical thin-layer models make up the last category of thin-layer drying equations. Empirical models are based
on quadratic or exponential form, or a mixture of both, with correlating variables based on experimental data. These
models provide a trivial calculation that can be verified for a specific material.
Phenomenological models make up another family of models applied to fluidized beds. These models are built on
physical theory but have empirical correlations to supplement the model. Commonly, heat and mass transfer equations
are coupled with empirical equations for the rates of transfer along with other parameters such as minimum fluidization
specifications and bubble size and velocity. Models have been constructed using one, two, or three phases to describe
the interior of the fluidized bed. A one-phase model encapsulates the solid and gas mixture into one bulk phase for
mass and heat transfer [23]. Two-phase models add another phase to account for the voidage bubbles of gas. Two-
phase models split the fluidized bed into a bubble phase and emulsion (or dense) phase that accounts for both the solid
material and interacting interstitial gas [11]. Lastly, three-phase models incorporate a cloud-wake phase in addition to
the bubble and emulsion phases, as found in Figure 2.7. The cloud-wake phase essentially describes the interaction
between the solid material within the emulsion phase and the voidage bubbles [24]. The free-flowing bubbles infiltrate
the emulsion phase as they flow upwards, causing some of the solid material to temporary transfer to the cloud phase
above the bubble. Similarly, the bubbles leave a turbulent wake phase below each bubble, again containing solids
material from the emulsion phase. Each type of model has been successfully verified using various experimental data
for batch and continuous configurations.

Figure 2.7: Rising bubble using three-phase configuration [25]

There has been some work completed for fluidized bed drying modelling involving computational fluid dynamics
(CFD) and its coupling with the discrete element method (DEM) to describe the solid material and the individual

13
particle motion within the bed [26] [27]. These models use the discrete element approach to describe each solid
particle’s state, including the particle’s position, velocity, temperature, and moisture content. This information is then
used by CFD computations to determine the fluid flow and other interactions that ultimately influence the heat and
mass transfer from each respective particle. Liquid bridge forces have been modelled effectively using CFD-DEM,
which may lead to better understanding of moisture effects on hydrodynamics. CFD-DEM models may prove to be
increasingly useful in modelling fluidized bed drying in the future; the coupling of both methods lowers computational
complexity while still providing both fluidization information and heat and mass transfer phenomena.

2.3.2 Utilization of the Two-Phase Model

The two-phase model has been a successful method in portraying the drying process in fluidized beds for pharmaceu-
tical powders. Both batch and continuous configurations have been explored. The two-phase model is an extension of
Kunii and Levenspiel’s three phase model (bubble/cloud-wake/dense) [24], with simplifications to omit the cloud-wake
phase. As a result, the two-phase model partitions the fluidized bed volume into only bubble and dense phases, seen
in Figure 2.8. The bubble phase is comprised of the void spaces in the fluidized bed in the form of upward flowing
bubbles. Theoretically, minimal to no solid particles are associated with the bubble phase. On the other hand, the dense
phase is comprised of all of the solid particles, along with the interstitial gas that directly interacts with the particles
through mass and heat transfer.

Figure 2.8: Two-phase schematic diagram for continuous (additions in blue) and batch drying

14
The first advancement for the two-phase model was completed by Palancz [11], who used the Kunii and Levenspiel
equations and reduced the solids interaction to be represented by a single particle. Following this, a variety of con-
tinuous and batch modelling research has been completed, heavily relying on both Kunii and Levenspiel and Palancz.
Duncan and Li have verified the two-phase model for pharmaceutical powders, adding variation to the bubble temper-
ature and moisture in the vertical direction [28]. Gagnon et al. have further reviewed this model and included heat loss
to the vessel walls [29].
These two-phase models have been applied to optimization methods and process control. Duncan and Li have used
the model for generic model control with the addition of an extended Kalman filter to manage the moisture content
[30]. Gagnon et al. attempted to control the process using non-linear predictive control coupled with a non-linear
moving-horizon estimator [13]. Lastly, Villegas et al. determined optimal control functions for inlet gas temperature
and velocity for drying rice [31].

15
C HAPTER 3
VALIDATION OF T WO -P HASE F LUIDIZED B ED D RYING M ODEL
FOR P HARMACEUTICAL P OWDERS

This chapter is a portion of a draft to be submitted for publishing.

• M. L. Lau, R. J. Spiteri, and L. F. Zhang. Validation and Optimization of two-phase fluidized bed drying model
for pharamaceutical powders.

The first author conducted the simulations, post-processed and analyzed the results, and prepared the first draft of
the paper. He then worked under Dr. Spiteri and Dr. Zhang’s supervision to finalize the content.

16
3.1 Abstract
In the pharmaceutical industry, batch fluidized beds are commonly employed to dehydrate wet powder granules in the
tablet production process. The granules are generally produced by wet granulation, with the drying stage necessary
before tablet compaction. Fluidized beds are effective for the drying process due to their optimal solid/gas mixing and
high mass and energy transfer. However, fluidized bed drying is still a relatively low efficiency process and can ben-
efit from optimization, especially in the pharmaceutical industry, where extensive changes are difficult to implement.
Phenomenological models can simulate the drying process and constitutes one method of finding better solutions and
operation techniques. Before a model can be utilized, it must be validated with real world data to ensure the model
properly responds to various parameters.
A previously studied two-phase phenomenological model is verified using recent fluidized bed drying data from
a lab scale setup. The model is comprised of the mass and energy balances of five distinct sections of the fluidized
bed. Powder moisture and heat transfer are governed by a stagnant gas film in equilibrium with the surrounding gas.
Bed moisture content, temperature, and outlet humidity are examined in comparison to experimental data. The effect
of various inlet gas velocities and temperatures are compared and discussed using the model. The phenomenological
model displays good correlation with the three parameters examined.

3.2 Introduction
In the pharmaceutical industry, tablets are often produced by wet granulation of powders. This method typically
requires a drying process to remove the moisture, with fluidized bed dryers being commonly used. Fluidized beds are
effective at drying pharmaceutical powders because of the fluidization of the conglomerate granules providing optimal
contact time between the solids and fluidizing air. Despite the efficiency of the heat and mass transfer in fluidized beds,
drying is still an energy-intensive process that can have significant cost savings if optimized. As with other processes,
fluidized bed drying has been modeled using various levels of complexity and physical phenomena. Due to the addition
complexities from the drying aspect (wetting, particle diffusion, cohesion, etc.), models for fluidized bed drying are
typically material specific, with correlating parameters and equations to fit the experimental data. In some cases,
models employ both heat and mass transfer theory in tandem with empirical equations to describe the drying process.
These are known as phenomenological models, which provide a reasonable representation of the drying process while
having a low computational cost.
Phenomenological models are physical descriptions based on fundamental theory, simplified with empirical cor-
relations to conform to known results. In other words, the models are rooted in known theories but have additional
manipulation to be suitable for the data. These types of models have both benefits and downfalls. The addition of
empirical equations can bridge the gap between fundamental theory and physical phenomena when the complexity of
the system is too expensive to fully describe. Specifically for fluidized bed drying, the distinct differences in drying
characteristics of various materials (pharmaceutical powder, biomass material, food grains, etc.) can be accommo-

17
dated. The effects of the excess moisture on fluidization and agglomeration can also be accounted for. However, due to
the simplification of the problem, certain aspects of the actual process are lost. Although the empirical equations may
help to predict the drying accurately, the tangible reasoning behind the process outcomes may not be captured. This is
why phenomenological models often lack generality and are only be applicable to specific materials.
There have been many successful attempts at modelling pharmaceutical powder drying in fluidized beds. Gavi [23]
has applied a simplified mechanistic one-phase approach to model lab-scale drying. Wang et al. [32] established a
model using a three-phase theory along with computational fluid dynamics (CFD) to describe the process spatially.
Two-phase models, employing an emulsion phase (powder and interacting interstitial gas) and bubble phase, have also
been validated. The models are ultimately built on the work of Palancz [11], along with Kunii and Levenspiel [24],
who present the framework for many fluidized bed drying models. Li and Duncan [28] effectively described the drying
process by simplifying the powder diffusion attributes. The work was further validated and improved by Gagnon
[29], supplementing the model with vessel heat loss. The two-phase model can be further validated using the drying
data provided by Taghavivand et al. [4] to evaluate the model with another distinct dataset. This validation has been
explored in this work.

3.3 Materials and Methods

3.3.1 Two-Phase Model

The two-phase model examined in this work is directly from the efforts of Gagnon [29]. The model is a continuation of
Li and Duncan’s model [28], with the addition of freeboard and vessel interactions for heat and moisture transport. The
model partitions the fluidized gas and solids within the fluidized bed into two-phases, the dense phase and the bubble
phase. The dense phase is comprised of the particles and interstitial gas interacting with the particles. The bubble
phase makes up the rest of the gas flowing through the dense phase in the form of bubbles that do not directly interact
with the particles. Mass and energy transfer equations are based on these two-phases, with many of the correlating
variables for heat and mass transfer based on Kunii and Levenspiel’s theories [24]. The powder within the dense phase
is represented by one particle and its interactions with the drying gas.

18
Figure 3.1: Mass and energy transport within model, adapted from [29]

Model equations can be organized into the five groups found in Figure 3.1; the dense and bubble phases, the particle,
the freeboard region and vessel walls. These groups make up all the mass and energy balances needed for computation.
In Figure 3.1, water transport is represented by blue arrows, and energy transfer is represented by red arrows. Bulk
properties and transport phenomena are used for the dense phase, vessel wall, and freeboard regions. Powder mass
and heat transfer occurs solely with the dense phase and is governed by empirical concepts for the evaporation of the
surface moisture and internal moisture within the particle, leaving the particle from a stagnant gas film. The bubble
phase is discretized into layers in the vertical direction to represent the freely flowing bubbles. This layered description
attempts to describe the bubble interaction with the moist air in the dense phase as the bubbles move upwards through
the fluidized bed.

19
3.3.2 Single Particle Representation

The interaction between the pharmaceutical granules and the drying gas is reduced to one single particle representing
the whole mass. The particle only interacts with the interstitial gas in the dense phase for heat and mass transport.
Interactions with the gas occur through a thin stagnant gas film, which is in equilibrium between the gas and particle.
This configuration is based on the following assumptions:

1. Particles are equal in size, sphericity, and other physical properties. Particle temperature and moisture content is
constant throughout the particle. Moreover, no particle size distribution is considered.

2. Particles have constant volume, with no shrinking during the drying process.

3. Particles have uniform temperature and moisture content throughout the solid bed.

4. No particle-particle interactions are considered in the model.

5. The particles are considered perfectly fluidized, with no variation of temperature and moisture content between
particles.

The drying process can be separated into two periods, the constant drying rate and falling drying rate period. The
phases can be interpreted using the particle moisture characteristics. The constant drying rate period is associated with
excess external moisture at the particle surface whereas the falling drying rate period is related to the internal moisture
and its resistance to evaporation. The falling rate period begins when the external or excess moisture has been fully
evaporated. This change occurs at what is known as the critical moisture content, xcr , which remains constant for the a
given material.

3.3.2.1 Mass Balance

The mass balance for the particle involves the loss of water to the stagnant gas film. Because the film is in equilibrium
with both the gas and solids in the dense phase, the mass flow rate from the particle to the film is equal to the mass
flow rate from the film to the dense phase gas. By describing the moisture by dry basis, the rate of change in particle
moisture content, x p , can be attributed to the mass flow rate from the stagnant film to the dense phase, ṁs f ,d . The mass
of the dry particle ms is included assumed to be constant, with the equation written as:

dx p
ms = −ṁs f ,d . (3.1)
dt
The mass flow rate, ṁs f ,d , is driven by the difference in moisture content between the film, xs f , and the dense phase,
xd . By using a mass transfer coefficient, ks f ,d , and volume and surface area equations for the particle, the equation can
take the form of [32]

dx p 6
ρs =− ks f ,d ρg (xs f − xd ) , (3.2)
dt φ dp

20
where ρs and ρg denotes the density of a dry particle and gas respectively, φ is particle sphericity, and d p describes the
particle diameter.
During the constant drying rate period, there is excess water at the surface of the particle to interact with the
stagnant gas film. Therefore, the film moisture content, xs∗f , during the constant drying rate period can be described as
saturated and can be determined using partial pressures and Dalton’s law by

Pwv
xs∗f = 0.622 (3.3)
P − Pwv
using the partial pressure for water vapour, Pwv and the vessel pressure, P. The partial pressure of the water vapour is
a function of the temperature of the gas. Because the dense phase temperature is equal to the particle temperature, Tp ,
the partial pressure equation takes the form of

 
5173
Pwv = 1 × 105 exp 13.869 − . (3.4)
Tp + 273.15
For the falling rate period, the moisture available for the stagnant gas film starts to the decrease due to diffusion
of the internal moisture to the particle surface. An empirical function is used to describe the surface with fitting
coefficients α and κ. The equation uses the particle moisture content, x p , and the critical moisture content to produce
the suitable values. A piecewise function split by the critical moisture content is used to correctly weight the saturation
moisture content.

1 x p ≥ xcr

ψ= α α
(3.5)
 xαp (xcrα +κ)

x p < xcr
xcr (x p +κ)

Equation (3.5) can be used to calculate the film moisture throughout the entire process by combining it with Equation
(3.3) to make an equation for the stagnant film moisture content, xs f , in the form of

xs f = xs∗f ψ (3.6)

to calculate the film moisture throughout the entire process.

3.3.2.2 Energy Balance

In terms of the particle, there are two sinks for the drying gas energy. The energy of the dry gas is used to evaporate
the particle moisture and increase the particle enthalpy, I p . The simple balance takes the form of

dI p
= Q̇d,p − I˙p,d , (3.7)
dt
where Q̇d,p represents the energy stream supplied by the gas and I˙p,d represents the energy needed for desorption and
evaporation of the particle moisture. Through substitutions, Equation (3.7) is transformed into

dTp 6 6  
ρ pd (cs + x p cw ) = hd,s f (Td − Tp ) − ks f ,d ρg (xs f − xd ) ∆Hevap + (cwv − cw )(Tp − Td ) . (3.8)
dt φ dp φ dp

21
On the left-hand side of the equation, the change in particle enthalpy is described by the change in particle temperature,
Tp , along with the specific heat capacity of the solid particle, cs , and the particle moisture, cw . On the right-hand side
of the equation, Qs f ,p is described using a heat transfer coefficient, hd,s f and is driven by the difference in temperatures
between the dense phase, Td , and the particle. The second term contains the amount of water transfer between the film
and dense phase to calculate the amount of energy used by water evaporation. The enthalpy transfer associated with
moisture removal is mostly due to the enthalpy of evaporation, ∆Hevap , but it also includes the change in enthalpy for
the evaporated water. The addition of the water vapour heat capacity, cwv is needed to fully account for the transition.
The amount of energy needed for evaporation can be determined by [32]:

∆Hevap = 3168 − 2.4364Tp . (3.9)

3.3.3 Bubble Phase

The bubble phase represents the void spaces within the fluidized bed occupied by free-flowing bubbles through the
vessel. There is a solid framework of research exploring the characteristics of the bubbling phenomena in fluidized
beds. Bubble velocity, size, and bubble interactions have be successfully modeled and verified. In this model, the
bubble phase is instead treated as a control volume alongside the dense phase.

1. Bubbles are spherical and have uniform diameter and velocity throughout vessel

2. Mass and energy transfers occur with the interstitial gas of the dense phase and not directly with the pharmaceu-
tical powder

3. Bubbles move through the vessel in plug flow with constant velocity throughout

4. Bubbles do not coalesce or interact with each other

The bubble phase control volume is arranged to describe the heat and mass balances as a function of height. In
other words, the model represents how the bubbles heat the dense phase and pick up moisture along the way as they
flow through. Bubble phase volume is determined by:

Vb = δV f , (3.10)

where Vb represents the bubble control volume and V f represents the full volume of the two-phase system (dense and
bubble phase). The fraction of fluidized volume that is part of the bubble phase, δ , is calculated with

vin − vm f
δ= . (3.11)
vb
In Equation (3.11), vm f is the minimum velocity for fluidization determined for the system. The velocity difference is
divided by the velocity of the bubbles, vb . In this model, vb is considered constant as plug flow and calculated using

22
the equation [24]

vb = (vin − vm f ) + 0.711(gdb )1/2 . (3.12)

3.3.3.1 Mass Balance

The mass balance of water transported to the bubble phase is based on the vertical height of the bubble control volume.
The vertical direction is characterized by the z-coordinate. By viewing the bubble control volume with respect to the
vertical direction, and assuming mass transport is always at steady-state in this direction, the mass balance can be
written as

ṁb (z + dz) − ṁb (z) = d ṁd,b , (3.13)

where the difference in bubble water transport rate as a function of height is determined by the infinitesimal mass
transfer rate value from the dense phase to the bubble phase, d ṁe,b . By evaluating each term and converting the mass
flow rates to dry basis moisture content, the equation takes the form of

dxb  
vb = Kd,b xd − xb (z) . (3.14)
dz
Equation (3.14) describes the bubble phase moisture content, xb , as a function of height. The mass transport is driven
by the moisture difference of the dense phase moisture content, xd , and the bubble moisture content at its respective
height in the vessel. As usual, a mass transfer coefficient governs the transport, as is represented by Kd,b . By adding
the boundary condition that the bubble moisture content is equal to the inlet gas at z = 0, the analytical solution to the
differential equation is

 
Kd,b
xb (z) = xd − (xd − xin )exp − z (3.15)
vb
to determine the bubble moisture content in a simpler fashion. However, the bubble temperature as a function of height
is not as trivial and pairs with the bubble moisture content. Therefore, the Equation (3.13) is used within the model.

3.3.3.2 Energy Balance

The energy balance of the bubble phase is similar to the mass balance, having the enthalpy fluxes characterized by
the same control volume with respect to the vertical z-direction. The bubble phase transports heat to the dense phase,
represented by Q̇b,d , and also receives an enthalpy flow from the dense phase due to the evaporated water, I˙d,b . By
assuming steady-state across the z-direction, the bubble phase energy balance using the enthalpy fluxes takes the form
of

I˙b (z + dz) − I˙b (z) = −d Q̇b,d + d I˙d,b . (3.16)

23
The addition of bubble phase enthalpy can be described by the change in bubble temperature. By using the temperature
and moisture differences between the dense and bubble phase, along with heat and mass transfer coefficients, the
balance can be transformed into:

 
  dTb   Hb,d  
vb cg + xb (z)cv = − Tb (z) − Td + Kd,b cv xe − xb (z) , (3.17)
dz ρg
with the introduction of Hb,d for the bubble to dense phase heat transfer coefficient and Kd,b representing the mass
transfer coefficient for the moisture moving to the bubbles. Again, a boundary condition for the differential is set at
z = 0, by fixing the bubble temperature equal to the inlet gas temperature at the initial height.

3.3.4 Dense Phase

The dense phase represents the second component of the fluidized volume, V f , and is comprised of all the powder
particles and interacting interstitial gas. Unlike the bubble phase, the dense phase is considered perfectly mixed with
no variation in moisture and temperature in the vertical direction of the bed. Similar to the bubble phase, the dense
phase is considered to be in steady-state. For mass and energy balances with the bubble phase, averages are used to
represent the parameters, which are denoted by x̄b and T̄b for bubble moisture content and temperature. Although the
powder is a portion of the dense phase, the gas and powder interactions are accounted for in the mass and energy
balances.

3.3.4.1 Mass Balance

The dense phase mass balance is comprised of moisture content gain from both the inlet gas and powder moisture. The
moisture exits the dense phase to the bubble phase as well as the freeboard region before exiting the fluidized bed. The
mass equation can be written as

ṁin + ṁ pt,d = ṁd,r + ṁd,b , (3.18)

with ṁ pt,d representing the water flow rate from the mass of powder to the dense phase and ṁd,r representing the
water exiting the dense phase to the freeboard region. Because the dense phase is assumed to be perfectly mixed, bulk
moisture content is equal to the moisture content exiting the freeboard region. This serves to reduce the mass balance
to three terms driven by the moisture content differences. After applying area and volume conversions for each mass
flux, the equation becomes

vm f 6(1 − ε f )
(xd − xin ) = ks f ,d (xs f − xd ) − δ Kd,b (xe − x̄b ) , (3.19)
zf φ dp
with z f and ε f representing the height and voidage space of the fluidized powder. The first term represents the change
in moisture content between the beginning of the dense phase and the final moisture at the outlet to the freeboard
region. On the right hand side, the terms are organized to describe the collection of moisture from the particle and the
loss to the bubble phase.

24
3.3.4.2 Energy Balance

The energy balance of the dense phase can be captured with three energy sources and four energy sinks. With sources
on the left hand side of the equation and sinks on the right, the balance takes the form of

I˙in + I˙pt,d + I˙b,d = I˙d,r + I˙d,b + Q̇d,pt + Q̇d,v . (3.20)

Equation (3.20) introduces the enthalpy and heat flows to and from the dense phase and powder with I˙pt,d and Q̇d,pt ,
respectively. Inlet enthalpy rate flow is denoted as I˙in , with bubble to dense phase heat flow as Q̇b,d . Enthalpy from the
evaporated moisture transporting into the bubble phase is accounted for with I˙d,b , and finally, Q̇d,v represents the loss
of heat to the vessel wall. Similar to the mass balance, an average temperature for the bubble phase, T̄b is used. The
resulting energy balance equation can be seen as

vm f δ Hb,d 6(1 − ε f )  hd,s f  4hd,v


(cg + xin cv )(Td − Tin ) = (T̄b − Td ) + ks f ,d (xs f − xd )cv + (Tp − Td ) − (Td − Tv ) . (3.21)
zf ρg φ dp ρg dv ρg

Equation (3.21) looks similar to the mass balance, with simplification by organizing the equation into four terms. The
first term represents the enthalpy change of the dense phase gas between the outlet and inlet in the form of heat. On
the right hand side, the first term contains the energy added by the warmer bubbles. The second, larger term contains
three separate flows, I˙pt,d , Q̇d,pt , and I˙d,b . The enthalpy flow from dense to bubble phase in the term is less intuitive.
It is determined by manipulating and substituting the dense phase mass balance in Equation (3.18) to simplify the
equation. Lastly, the final term on the right-hand side denotes the vessel heat loss and is governed by the difference in
temperature from the vessel wall, Tv , and another heat coefficient, hd,v .

3.3.5 Vessel Wall and Freeboard Region

The vessel walls and freeboard region are the last volumes to account for in the fluidized bed vessel. The freeboard
region of the fluidized bed is defined by the void gas area above the fluidized product in the vessel. The vessel walls
are configured to account for a dynamic change in temperature during the process. It is inevitable that the vessel walls
gather some energy from the gas in both the fluidized area of the fluidized bed as well as the freeboard region. Assuming
uniform temperature distribution across the dense phase and freeboard region respectively, the energy balance of the
vessel walls is described as:

dIv
= Q̇d,v + Q̇r,v − Q̇v,a , (3.22)
dt
with vessel wall enthalpy denoted by Iv . The heat flows to the vessel wall from the dense phase, Q̇d,v , and the freeboard
region, Q̇r,v . The heat flow from the outer side of the vessel to the ambient air is represented by Q̇v,a . The surface
area and volume of the vessel walls, along with heat transfer coefficients for each heat flow term, transform the energy

25
balance into:

mv cv dTv dvo
= z f hd,v (Td − Tv ) + (zv − z f )hr,v (Tout − Tv ) − zv hv,a (Tv − Ta ) , (3.23)
πdv dt dv
with mv , cv , and zv representing the mass, specific heat capacity, and height of the vessel. The inner diameter of
the vessel is denoted by dv , and the outer diameter by dvo . Each term has an associated heat transfer coefficient,
represented by hd,v , hr,v , and hv,a respectively. Because the freeboard region is considered perfectly mixed, the outlet
gas temperature, Tout is equal to the freeboard temperature, Tr . Therefore, Tout is used instead for simplification.
The freeboard region requires a mass and energy balance to describe the outlet temperature and moisture content. It
can also be thought of as the recombining of the gas from the dense phase and bubble phase before exiting the system.

3.3.5.1 Mass Balance

The mass balance of the freeboard region can be written as:

ṁr,out = ṁd,r + ṁb,r , (3.24)

with ṁr,out as the outlet moisture flow and ṁd,r and ṁb,r representing the mass flow of water from both the dense and
bubble phases, respectively. The dense phase bulk moisture content, xd , can be used for the mass balance. However,
due to the nature of the bubble phase and its increasing moisture content as the bubbles rise, only the moisture content
at the top of the bubble phase is used. The equation then takes the form of:

vin xout = vm f xd + δ vb xb (z f ) . (3.25)

3.3.5.2 Energy Balance

The energy balance is similar to the mass balance, with the addition of the heat loss to the vessel wall. The balance is
written as:

I˙r,out = I˙d,r + I˙b,r − Q̇r,v . (3.26)

By converting each term to thermal energy and organizing the terms, the equation changes to:

 
4(zv − z f )hr,v  4(zv − z f )hr,v
vin (cg + xout cv ) + Tout = δ vb cg + xb (z f )cv ]Tb (z f ) + vm f [cg + xd cv ]Td + Tv . (3.27)
dv ρg dv ρg

3.3.6 Heat and Mass Transfer Coefficients

The model employs heat and mass transfer coefficients between each phase. These coefficients have been extracted
from previous literature by Gagnon et al. [29].

26
Table 3.1: Summary of Transfer Coefficients

Coefficient Equation Description


hd,s f
Particle Mass ks f ,d = cg Lewis factor [5]
Transfer
λg  d p vin ρg 0.5 cg µg 1/3 
Particle Heat hd,s f = d p 2 + 0.6 µg λg
Ranz-Marshall correla-
Transfer tion [24]
1 1 −1

Bubble/Dense Kd,b = Kb,c + Kc,d Normalized bubble-
Mass Transfer cloud-dense coefficient
[24]
1
−1
Bubble/Dense Hd,b = Hb,c + H1c,d Normalized bubble-
Heat Transfer cloud-dense coefficient
[24]
λg   vin dv µg 0.325 ρ p cs dv1.5 g0.5 0.23 cg µg 0.3
Dense/Vessel hd,v = 47 dv 1−εf ρ p d 3p g λg λg
Empirical equation
Heat Transfer from horizontal tube
and gas heat transfer
[33]
λg  vin dv ρg 0.55 cg µg 1/3
Freeboard/Vessel hr,v = 3.95 dv µg λg
Empirical equation us-
Heat Transfer ing silica sand [34]
 cg µg 1/6 2
λg 0.387(Gr λg )
Vessel/Ambient hv,a = zv 0.825 + 0.492λg 9/16 8/27 Empirical equation for
[1+( cg µg ) ]
Heat Transfer convection from verti-
cal plate [35]

Table 3.1 is a summary of the heat and mass transfer coefficients. Many of the derived equations need the additional
drying gas parameters of thermal conductivity, λg , and dynamic viscosity, µg . The bubble phase and dense phase
coefficients are calculated using a third (cloud) phase as described by Kunii and Levenspiel [24]. The mass and heat
transfer coefficient equations from the bubble phase to cloud phase and the cloud phase to the dense phase are as
follows, with the new subscript c denoting the cloud phase.

vm f D0.5 g0.25
Kb,c = 4.5 + 5.85 wv1.25 (3.28)
db db

εm f Dwv g0.5 0.5


 
Kc,d = 5.71 (3.29)
db2.5

vm f ρg cg (λg ρg cg )0.5 g0.25


Hb,c = 4.5 + 5.85 (3.30)
db db1.25

0.5
εm f λg ρg cg g0.5

Hc,d = 5.71 (3.31)
db2.5

27
The bubble-cloud-dense phase transfer coefficients are normalized to create the bubble to dense phase coefficients,
Kd,b and Hd,b . The coefficients introduce another variable for the diffusivity of water vapor in air, Dwv . The final heat
transfer coefficient in Table 3.1, hv,a also introduces a new variable, Gr. This is known as the Grashof number and can
be determined by

gz3v |Tv − Ta |ρg2


Gr = (3.32)
0.5(Tv + Ta )µg2

Minimum Fluidization Velocity The minimum fluidization velocity of the inlet gas, vm f , and voidage space at the
minimum fluidization velocity, εm f are important parameters for the model. Both variables dictate the partition of gas
flow in the vessel and contact time with the pharmaceutical granules. The voidage space at minimum fluidization can
be determined using an empirical equation [36] written as

0.029  0.021
µg2

−0.72 ρg
εm f = 0.586φ , (3.33)
ρg (ρ p − ρg )gd 3p ρp
where φ is the particle sphericity, µg is the dynamic viscosity of the gas, and ρg and ρ p represent the density for the gas
and particle. The particle diameter is represented by d p , and g is the gravitational constant. The minimum fluidization
voidage space can be used with the Ergun equation [24] to calculate the minimum fluidization velocity by

d 2p (ρ p − ρg )g εm3 f φ 2
vm f = , (3.34)
150µg 1 − εm f

3.3.7 Model Algorithm

The system of equations for the model form a differential-algebraic equation (DAE) problem. The particle moisture
content, particle temperature, and vessel temperature are determined using differential equations, with the rest of the
system determined by algebraic constraints. The Matlab solver ode15s is used to solve the DAE due to its problem
stiffness. The variables are summarized in Table 3.2, along with the associated equations.

28
Table 3.2: Summary of Model Formulation

Parameter Variable Equation

Particle Moisture Content xp Equation (3.2)


Particle Temperature Tp Equation (3.8)
Vessel Temperature Tv Equation (3.23)
Bubble Phase Moisture Content xb Equation (3.14)
Bubble Temperature Tb Equation (3.17)
Dense Phase Moisture Content xd Equation (3.19)
Dense Phase Temperature Td Equation (3.21)
Outlet Gas Moisture Content xout Equation (3.25)
Outlet Gas Temperature Tout Equation (3.27)

The bubble phase moisture and temperature equations have a partial derivative with respect to the z-direction. In
order to deal with the spatial derivative, the bubble phase volume is discretized into n layers to create a group of
algebraic equations. The change in moisture and temperature with respect to the z-direction is determined by the finite
difference method, using the forward difference to explicitly solve each layer. By discretizing the spatial derivative and
using the forward Euler method, the moisture and temperature equations for the bubble phase, Equations (3.14) and
bube ncanbewrittenas :

Ke,b
xb,n+1 = xb,n + ∆z (xd − xb,n ) , (3.35)
vb

Hb,d  
ρg + Kd,b xd − xb,n cv  
Tb,n+1 = Tb,n + ∆z   Td − Tb,n , (3.36)
vb cg + xb,n cv
with ∆z as the layer height, and n and n + 1 representing the additional layers as they accumulate until the fluidized
powder height, z f . Finally, the average values for moisture and temperature, x̄b and T̄b , can be calculated by taking the
average of each layer. In this work, 40 layers were used to discretize the bubble phase, although no significant issues
were found with minimizing layers.

3.3.8 Experimental Data

The model was analyzed in comparison to the data collected by Taghavivand et al. [4] with various inlet gas velocities
and temperatures. The provided data display the powder moisture content as a function of time, but they do not include
any particle temperature data. However, particle temperature characteristics can still be discussed based on the general
trends in particle temperature in literature and recent pulsated flow data from the same experimental set up.
The details of the experiment can be found in [4]. The fluidized bed is cylindrical with a diameter increasing from
11 cm to 25 cm, as can be seen in Figure 3.2. The variable diameter section is where the bed of fluidized particles reside.

29
To account for this, the vessel diameter is set for the average of these values at 18 cm for the fluidized powder volume,
v f , and set for the full 25 cm for the freeboard region. It is important to note that the vessel diameter only serves as
a surface area and volume parameter as no hydrodynamics are described in the model. Consequently, the change in
gas velocity due to increasing diameter is neglected; however, the superficial velocity serves more as a volumetric flow
rate for each control volume. Therefore, the exclusion of capturing the increasing diameter geometry is considered not
significant for the results.

Figure 3.2: Product bowl dimensions [4]

The pharmaceutical material in the experiment is a mixture of typical pharmaceutical tablet composition. The
powders were mixed together with water to reproduce the granulated mixture. The wet granules were then screened
using a 3.36 mm size mesh screen for a more homogeneous granule size. The compositions of the granules are as
follows.

30
Table 3.3: Granule Formulation

Component Mass % by Dry Basis

Lactose Monohydrate 50%


Microcrystalline Cellulose 44%
Hydroxypropyl Methylcellulose 4%
Croscarmellose Sodium 2%
Water 42%

3.3.8.1 Applicable Parameters

The model needs various coefficients and parameters to simulate the experimental data correctly. Table 3.4 describes
the specific parameters pertaining to this experimental setup. Most variables are physical parameters of the materials
of the system.

Table 3.4: Summary of Constant Variables

Parameter Description Units Value Rationale

cg Heat Capacity Coeffi- J/kgK 1005 Physical parameter


cient - dry gas
cv Heat Capacity Coeffi- J/kgK 1996 Physical parameter
cient - water vapour
cs Heat Capacity Coeffi- J/kgK 1200 Based on powder mixture prop-
cient - dry particle erties
cw Heat Capacity Coeffi- J/kgK 4200 Physical parameter
cient - liquid water
ρg Gas density kg/m3 1.27 Physical parameter
ρs Dry Particle density kg/m3 1560 Physical parameter
ρw Water density kg/m3 992.3 Physical parameter
P Pressure millibar 1013.89 Pressure to overcome bed
weight of particles
xcr Critical Moisture Con- % dry basis 8% Experimental observation
tent Point
db Diameter of bubbles in m 0.035 Estimation
bubble phase
dp Diameter of particles m 0.001 Estimation
φ Particle Sphericity 0.9 Estimation

31
The last three variables in Table 3.4 are based on estimation and correlation to the experimental data due to their
values being uncertain. Bubble diameter is a complex feature of fluidized bed, so the bubble diameter used in the
model would fundamentally be an ”effective” bubble diameter. Likewise, particle diameter and sphericity are difficult
to determine because the granules are complex structures of agglomerates. Accordingly, the model is used in tandem
with the experimental data to estimate these values.

3.4 Results and Discussion

3.4.1 Particle Moisture Content

The main parameter in the drying process that is observed in the decrease in particle moisture content as it dries. As
stated before, moisture content is written in dry basis form, meaning the ratio of water mass and dry particle mass. The
model was examined using three inlet gas velocities of 1.0 m/s, 1.4 m/s, and 1.8 m/s, and four inlet gas temperatures
of 38◦ C, 50◦ C, 65◦ C, and 75◦ C. Therefore, a total of twelve moisture content datasets were compared to the model.
Figure 3.3 displays the moisture content data in comparison to the model, organized by the inlet gas temperature.
The model exhibits fairly good correlation with the data, albeit with some inconsistencies. At the lowest inlet gas
temperature of 38◦ C, the model predicts a faster moisture removal rate at all velocities. In contrast, the model also
underestimates drying rate for the lowest inlet velocity of 1.0 m/s at the two higher temperatures. The average R-
squared value for twelve datasets is 0.993, with a lowest value of 0.983 occurring at 1.0 m/s and 65◦ C inlet gas velocity
and temperature. These discrepancies do not have a clear interpretation but still predict the general relationship between
inlet gas parameters and moisture removal.

32
Figure 3.3: Moisture content profiles by dry basis

3.4.2 Particle Temperature

Although no direct particle temperature data is available for these experimental runs, the temperature profiles from the
model can still be examined. Recent work with the experimental fluidized bed and specific powder in question show
analogous results in particle temperature during the constant and falling rate regions. These results have the additional
complexity of inlet gas pulsated flow but nonetheless still provide valuable insight on the temperature ranges. From
the experimental work, it appears that the particle temperature during both the constant and falling drying rate periods
is a function of only inlet gas temperature and is not affected by the velocity. The general temperatures are as follows.

33
Table 3.5: Pulsated Flow General Particle Temperatures

Inlet Gas Temperature (◦ C) Constant Drying Rate Period (◦ C) Falling Drying Rate Period (◦ C)

40 21 30
50 23 37
60 25 45

The values of the inlet gas unfortunately do not coincide with the inlet gas temperatures of the provided dataset.
However, the general trend and relationship between the inlet gas temperature and particle temperature is still clear
and should be predicted by the model. Figure 3.4 displays the model’s particle temperature across the twelve inlet gas
configurations.

Figure 3.4: Model particle temperature profiles

From the temperature profiles in Figure 3.4, the model demonstrates multiple characteristics that are suitable for
pharmaceutical powders. Each drying run starts at the initial 20◦ C and quickly finds an energy equilibrium between the

34
moisture removal and inlet gas for the constant drying rate period. In this period, the model prediction shows slightly
higher particle temperatures with increasing velocity but still follows the same trend as Table 3.5 regarding the varying
inlet gas temperature. Following the constant drying rate period, the falling drying rate period begins with the second
increase in temperature. Profiles with no secondary increase are caused by the moisture content not reaching the critical
moisture point. At this transition to the falling rate, the particle rises to a certain temperature before it appears to stay
constant. Again, the model predicts the rise in temperature to the final value consistent with the values in Table 3.5.
The model appears to underestimate the increase at higher temperatures. This suggests a possible overestimation of
energy lost to the vessel at higher temperatures.

3.4.3 Relative Humidity

The relative humidity of the outlet gas from the top of the fluidized bed was also captured during experimental work
for some runs [4]. In addition, outlet relative humidities from the pulsated inlet gas flow experimental runs were also
recorded. The experimental data display no observable trends with respect to inlet gas parameters and follow the same
characteristics on each run. Figure 3.5 displays the general relative humidity trend with a pulsated drying run for an
example.

Figure 3.5: Experimental relative humidity example for general trend

The humidity ramps up to a constant of approximately 80% relative humidity for the constant drying rate period.
Once the particle reaches the critical moisture point, the outlet air gradually becomes dry as the moisture removal rate
decreases.

35
Figure 3.6: Model outlet relative humidity

The model exhibits slightly lower relative humidities during the constant drying rate period, as seen in Figure
3.6. Also, the initial increase is almost immediate in comparison to the experimental relative humidity. This quick
transient response is due to mass balance constraints, with the model predicting the equilibrium values quickly with no
restrictions. The model also predicts slightly different relative humidities with the various inlet gas temperatures and
velocities. The low relative humidities may be due to the ratio of gas in the bubble and dense phases. In the constant
drying rate period, the dense phase in the model is considered fully saturated (100% relative humidity). The dry air in
the bubble phase mixes with the saturated gas from the dense phase to lower the relative humidity at the outlet. The
ratio of gas into the bubble phase along with the minimal mass transport of water to the bubble phase may be the cause
of the lower relative humidities. Another possibility is an unreliable value calculated for the minimum fluidization
velocity, vm f . The minimum fluidization velocity is calculated using the Ergun equation [24]. This equation does
not account for the moisture content of the particles, so cohesion and agglomeration effects caused by the water are
neglected. The effective minimum fluidization velocity may be significantly higher, especially at the beginning of the
process.

36
Regardless, the model predicts a few trends worth discussing that are not shown in experiment. First, at a constant
inlet gas temperature, the relative humidity decreased with increasing inlet gas velocities. As the inlet gas velocity
increases, more of the inlet gas goes into the bubble phase. The additional dry bubble phase gas lowers the outlet relative
humidity. Second, at a constant inlet gas velocity, the relative humidity increases with increasing inlet temperatures.
This trend was not expected because the increase in particle temperature was thought to coincide with an increase in
the amount of water the gas can carry (dictated by the partial pressure of the water vapour. However, it appears that
the additional thermal energy causes even more moisture removal by the drying gas than previously anticipated. These
trends are insightful aspects that can be further verified by more experimental data.

3.5 Conclusions
In this work, a phenomenological model is described and verified with lab-scale experimental data. The mass and
energy balances are organized by characterizing the fluidized bed into five distinct phases for moisture and heat trans-
port. Moisture and heat transports are governed by mass and heat coefficients established by empirical correlations
from previous work. The model displayed the same characteristics as experimental data and responds appropriately to
various inlet gas velocities and temperatures.
It was also found that the model predicted lower relative humidities in comparison to experiment, among other dif-
ferences. Experimental data also display a constant relative humidity with changing inlet gas velocity and temperature.
Conversely, the model’s relative humidity appears to be dependent on the two parameters. These discrepancy are signif-
icant and could benefit from more scrutiny. The differences could be caused by either incorrect model configurations,
experimental issues such as defluidization, or a combination of both.
The model in general displays that it captures the experimental fluidized bed drying process adequately enough for
optimization use. Although the simplification of phenomenological model lacks the ability to describe hydrodynamic
and kinetic phenomena, it can still provide useful bulk characteristics of the fluidized bed dryer with changing inlet gas
parameters.

3.6 Acknowledgements
The authors gratefully acknowledge financial support from NSERC and the University of Saskatchewan.

37
C HAPTER 4
O PTIMIZATION OF F LUIDIZED B ED D RYING USING A T WO -P HASE
M ODEL

4.1 Abstract
Batch fluidized beds are often employed in drying pharmaceutical powders and granules for tablet production. Batch
fluidized beds are defined by a continuous upwards gas flow through a bed of particles. In this study, a two-phase
model is utilized to examine optimization potential for the drying of pharmaceutical granules in a fluidized bed. The
model applies the common bubble and emulsion/dense phases of two-phase fluidized bed models, including heat loss
to the vessel and freeboard. The bed of pharmaceutical powder is represented by a single particle and its interactions
with the phases within the model. Energy and exergy analysis across a range of inlet gas configurations is explored.
To optimize the process, the model is explored using Matlab’s Global Optimization Toolbox, controlling the inlet gas
parameters and incorporating a stepwise change during the drying process. It was found that a single stepwise change
has a negligible effect on optimizing the process but is still useful around the end point of the batch. In addition,
optimization results and behaviours are discussed.

4.2 Introduction
The drying of wet granules is an important part of the producing tablets in the pharmaceutical industry. Before granu-
lated material can be compacted into tablets, it must be dried of the excess moisture that was added for the granulation
process. This drying process is commonly carried out in batch fluidized beds that employ an upwards gas flow through
the bed of wet granules to fluidize the mixture. Fluidized beds are chosen due to their optimal mixing qualities and
high heat and mass transfer capabilities. However, the process is not without waste; much of the drying gas energy is
not consumed and leaves the vessel before any useful interaction. Moreover, process control within the industry can be
limited, relying on operational expertise to select constant gas parameters with minimal realtime monitoring.
Mathematical modelling can be employed to better understand the drying process and explore optimization ideas
and theories. Models of various complexity for fluidized bed drying have been successfully examined and verified
by experimental data. The two-phase semi-empirical model verified in Chapter 3 is used to study the drying process.
The two-phase model applies a dense and bubble phase to explain the bed mixture of gas and powder, but is also

38
incorporates expressions to describe the individual particles, vessel walls, and freeboard region (empty section above
fluidized bed of powder). Each section contains its own heat and mass transfer and empirical equations. In this way,
the batch operation can be analyzed as a whole for energy and exergy utility from both the prospective of the entire
vessel as well as from the various areas of phases within the fluidized bed.
Optimization of the fluidized bed drying process generally involves minimizing energy loss, maximizing energy and
exergy efficiency, or maximizing production throughput. Aside from changing the fluidized bed vessel configuration
(dimensions, shape, air distribution method, mixing methods, etc.) or altering the batch size of powder, operations
mainly focus on inlet gas parameters, temperature and velocity. Both these parameters affect heat and mass transfer, and
furthermore, mixing efficiency, fluidization regimes, and material quality. When modelling, one must stay cognizant
of the bounds of the optimization space, to stay in efficient fluidization and product stability, both thermally and
kinetically.
Oftentimes, the inlet gas parameters are set to a constant throughout the drying process. This can lead to inefficien-
cies due to changing drying stages and the various behaviours that would benefit from different inlet gas parameters.
To counteract this, the optimization in this chapter introduces a simple one-time step-change that occurs during the
drying process. In this change, both velocity and temperature of the inlet gas are allowed to be modified. A step-wise
change can counteract issues such as thermal degradation when the product is dry and fluidization issues during the
beginning high moisture stages. Of course, full optimal control of these parameters would be ideal, but given practical
constraints, a step-change configuration may be the only feasible option.
In this study, energy and exergy efficiencies under constant inlet gas conditions are examined to analyze the perfor-
mance of a simpler setup within the two-phase model. Following this, the study delves into optimization of a one-step
change inlet gas parameter fluidized bed drying process. The time of the step-change is left as an optimization param-
eter, as well as the inlet gas velocity and temperature before and after the step-change. This configuration creates a
five-variable design space to explore using Matlab’s global optimization toolbox. By setting up an objective function
that contains energy and production considerations, optimization analysis results in an optimum step-change near the
endpoint of the process to lower the energy utility of the inlet gas. The step-change induces a 1.1% small savings in
energy and would continue to increase in significnace the longer the dryer operates past the end-point. Regardless of
the small improvement, the optimization analysis demonstrated that lowering the inlet gas parameters right before or
near the endpoint of drying will have a positive effect, especially if the dryer is run past the desired powder mois-
ture content. Other considerations are discussed, with varying magnitude of energy and production significance in the
objective function (energy cost vs revenue of dry product).

4.3 Materials and Methods

4.3.1 Model Description

The two-phase model has been verified by pharmaceutical-related experimental results produced by Taghavivand et
al. [4] but also earlier by Gagnon [29] and Li and Duncan [28]. Numerous additions have been implemented over

39
the years, allowing the model to describe more components of the fluidized bed drying process. Other research has
utilized this model with various optimization setups. As an overview, the two-phase model is a summation of heat and
mass transfer balances of bulk phases, incorporating various empirical correlations to bridge the sections together. The
model reacts quickly to changes in inlet gas parameters and does not capture the transient behaviour one may observe
with more gradual changes in real-time fluidization regimes. However, the bulk energy and moisture flows from the
model provide sufficient data to study inlet gas variations. Figure 4.1 displays the various energy and moisture flows
across the five sections.

Figure 4.1: General Two-phase fluidized bed model

With the various energy and moisture flows as seen in Figure 4.1, efficiencies can be analyzed at a section by
section level. In this study, the energy transfer between gas and powder is of most concern. Any energy utilized to heat
and pressurize the inlet air is considered wasteful if not used in some manner (moisture removal or powder tempering).
Therefore, optimization is most concerned with the enthalpy flow rate associated with the inlet gas, I˙in , and outlet gas,
I˙r,out . The moisture content of the bed of powder must reach a certain value to signify the endpoint of the process. This
is dependent on the mass flow rate of water escaping the particles, ṁs f ,d , throughout the drying time. The state of the
particle is observed for both temperature, Tp , and moisture content, x p , representing these parameters for the entire bed

40
of powder.

4.3.2 Energy and Exergy Efficiency

In addition to optimization, the model can be used to assess energy usage and efficiency over various inlet gas con-
figurations. This analysis is completed for both energy and exergy perspectives. Both energy and exergy analyses are
valuable for various reasons. Energy analysis is more direct, examining the proportion of energy used for drying the
powder, and in contrast, the amount of unused energy that leaves the system. This insight can show the direct rela-
tionship of drying efficiency and inlet gas parameters. In contrast, exergy analysis examines the theoretical amount of
useful work potential within the system. Exergetic efficiency is a useful tool because it accounts for the irreversibilities
and energy loss to entropy generation, phenomena that energy analysis does not consider. Both energetic and exergetic
analysis of fluidized bed drying processes have provided meaningful results in previous literature. In this study, a range
of inlet gas configurations is be examined for both energetic and exergetic efficiencies.
Energy and exergy analysis is not a novel approach for fluidized bed drying. Energetic and exergetic efficiencies
have been analyzed for drying carrot cubes [37] and harvested paddy [38]. Syahrul et al. studied drying efficiencies
and the effect of gas temperature and velocity, initial particle moisture, and bed hold up on wheat drying [39]. Assari et
al. successfully utilized a three-dimensional two-fluid model to study the energy and exergy efficiencies as a function
of time through the drying process [40]. Özahi and Demir validated their model and thermal analysis using both corn
and unshelled pistachios [12].

4.3.2.1 Energetic Efficiency

To determine the energy efficiency, the model’s energy balances can be used. The increase in particle temperature is a
secondary and undesired effect, leaving the energy efficiency equation as
Energy used to evaporate moisture
ηen = . (4.1)
Energy brought in by inlet gas
Equation (4.1) can be further described by calculating the energy for moisture evaporation, ṁs f ,d , by multiplying the
mass flux of evaporation, ṁevap , and the enthalpy of desorption and evaporation, ∆Hevap . The inlet gas energy is
denoted by Q̇d,p
∆Hevap ṁevap
ηen = (4.2)
Q̇d,p

4.3.2.2 Exergetic Efficiency

To obtain an exergetic efficiency, an entropy balance must first be determined. An exergy balance can be described by
defining the energy balance and subtracting the terms by the product of a reference ambient temperature, T0 , and the
entropy balance of the system. Once the exergy balance is configured the efficiency can be determined similar to its
energy counterpart
Exergy used to evaporate moisture
ηex = (4.3)
Exergy brought in by inlet gas

41
Before the terms can be described, an entropy balance of the system can be defined by

ṁevap ∆Hevap
ρg Gd (Sout − Sin ) = + Ṡgen . (4.4)
Tp

Equation (4.4) describes the balance by the difference in entropy in the gas flow, Sin and Sout respectively. The gas
has a volumetric flow rate, Gd , that interacts with the particles as it passes through the vessel, along with a density
parameter, ρg . The energy of desorption and evaporation is divided by the particle temperature, Tp . Lastly, a term is
added to account for the entropy generation within the process, Ṡgen . Equation (4.4) does not contain any terms for loss
of entropy to environment or change in entropy to the particles, due to the model assumes no variation of temperature
within the bed of particles. With the entropy balance, an exergy balance can be generated.

T0
Q̇gas − T0 ρg Gd (Sout − Sin ) = (1 − )ṁevap ∆Hevap + T0 Ṡgen (4.5)
Tp

With interest in only the inlet gas exergy and moisture evaporation exergy consumption, the values of Sout and Ṡgen are
not needed. However, a definition for Sin is necessary for exergy efficiency.

Tin
Sin = ln (4.6)
Tp

With equation (4.6), the exergetic efficiency equation can be constructed and determined by
 
T0
1 − Tp ṁevap ∆Hevap
ηex = . (4.7)
Q̇gas − T0 ρg Gd ln TTinp
.

4.3.3 Inlet Gas Parameters

The inlet gas parameters are the only elements of the drying process that are modified and examined. Within the model,
the inlet gas velocity and temperature parameters dictate the magnitude of energy flow into the system. These variables
also affect heat and mass transfer coefficients, controlling the rate of flow of energy and moisture between the five
phases. However, in real world applications, the effects are even more confounding. Many effects are not captured by
the model, but they still should be addressed due to their significance to the drying process. With this in mind, it seems
valuable to discuss process behaviour in a general sense in terms of the two main gas parameters.

4.3.3.1 Velocity

The velocity of the inlet gas is one of the main variables affecting fluidization. Sufficient inlet gas flow is necessary
to support fluidization, regardless of how favourable the material is (homogeniety, sphericity, Group A Geldert’s clas-
sification, etc.). Because of this, a minimum fluidization, vm f , must always be maintained. Excess moisture can also
complicate fluidization; the additional cohesive forces causes the particle bed to behave as inferior Geldert groups and
require a higher vm f . Proper fluidization is of most concern because it ensures optimal contact area between the gas
and bed of powder, supporting the mass transfer relied on for moisture removal.

42
Fluidization inherently becomes easier as the powder dries [10]. The increased interparticle forces from the excess
moisture has been demonstrated through experimentation, necessitating an increased minimum fluidizidation velocity,
higher bed voidage (more space between particles to counteract interparticle forces), and increased Hausner ratio. For
pharmaceutical powders, the moisture effect on fluidization seems to subside when the moisture contents reaches 20%
and continues to decrease until approximately 10%, where the powder starts to behave as a fully dry bed [10].
How the moisture obstructs fluidization has a significant effect on the drying process. Unfortunately, the two-
phase model implemented does not capture these implications. In terms of the model, inlet gas velocity impacts mass
balance equations and various heat and mass transfer coefficients. With this in mind, the optimization results will
need to be considerate of real-life limits. For example, low initial inlet gas velocity will not be considered a reasonable
optimization result. Even though the energy savings may be significant, application in real-life would prove impractical.

4.3.3.2 Temperature

From a bulk perspective, the influence of inlet gas temperature on the drying process is relatively straightforward.
Many studies have observed the favourable increase in drying rate with increasing inlet gas temperature. The issue
arises with high temperatures causing thermal degradation of the powder. Therefore, operations must ensure inlet gas
temperatures do not cause the powder to reach excessive temperatures. Degradation usually occurs at the end of the
drying process when there is a lack of moisture to consume the thermal energy from the inlet gas. In the beginning
stages, excess moisture found on the outer surface particles is the predominant energy sink for the inlet gas. This effect
tempers the powder temperature until no more surface moisture is available.
The model captures the powder temperature appropriately from a bulk perspective. Within the model, an increase
in inlet gas temperature introduces a higher energy flux to this system. The increase thermal energy raises the saturation
capacity of the gas, allowing it to carry more moisture from the bed. The assumption of a perfect mixture of powder and
gas, with each particle experiencing the same interaction with the gas, is effective for modelling the temperature profile
of the powder bed. However, similarly to inlet gas velocity, the model cannot represent the effects of defluidization.
Less-than-efficient fluidization can cause hot spots and disparities for temperature in various areas of the fluidized bed.
This issue is another justification to ensure the initial velocity is sufficient for mixing, both in modelling and real-world
application.

4.3.3.3 Step-Change

Changing the inlet gas parameters during the drying process is a simple method of maximizing energy usage for the
various conditions the fluidized bed experiences. In some circumstances, the batch process has a user-driven parameter
tuning throughout the various set points of the process. Adaptability and monitoring capabilities are usually limited
in these cases. Therefore, understanding the broad effect of a trivial step-change time, as well as its corresponding
new inlet gas parameters, may be applicable. Various studies have investigated changing inlet gas parameters in a
more dynamic manner. Optimal control using the two-phase model has been explored with beneficial results [29] [31].
Despite the advancements, a rigid one step-change configuration has not been explored. With the varying operability

43
and monitoring methods available, a step change may be all that is viable.
From an operation standpoint, an apparent beneficial configuration is to run the inlet gas at a high temperature
and velocity initially, and then drop to lower parameters once most of the moisture is removed. This setup incurs
the quickest drying time, whilst attempting to preserve energy at the latter portion of the process. Furthermore, the
decrease in inlet gas temperature moderates the increase in powder temperature when the particles are drier and more
susceptible to thermal degradation. Accordingly, this setup is anticipated to be one a possible optimization ”solution”.
From a model standpoint, a step-change creates an interesting region within the process. As previously determined,
the drying process for pharmaceutical powders has two distinct sections, the constant and falling drying rate periods.
These periods are named appropriately, associated to their respective drying rates. Additionally, it also represents the
time when readily available water is depleted, lowering energy efficiency. Due to the distinct sections, the relationship
between the time of inlet gas step-change and drying behaviour change will be significant and observable.

4.3.4 Optimization using Matlab

Using the above parameters, optimization can be contained within a five-variable design. Optimization is completed
using Matlab’s differential solvers for the model equations, controlled by Matlab’s Global Optimization Toolbox. Mat-
lab’s ”GlobalSearch” manages the design space, controlling the model within the parameter bounds. The optimization
is performed across a range of objective functions. The bounds are held constant throughout the range of optimization
jobs completed.

Table 4.1: Summary of modelling methods for fluidized beds

Variable Bounds

Initial Velocity 0.9 - 2.0 m/s


Initial Temperature 303.15 - 353.15 K
Final Velocity 0.9 - 2.0 m/s
Final Temperature 303.15 - 343.15 K
Time of Step Change 25 - 133 minutes

The design space is bound based on process considerations previously discussed. Inlet gas temperature is bounded
by particle temperature limits, mainly the upper thermal degradation limit. The minimum velocity bound is determined
by both the calculating minimum fluidization velocity and experimental observations [22]. In many cases, the real
minimum fluidization velocity (increased by moisture effects) is much higher than the minimum velocity set in the
design space. This will be considered if any low velocity solutions occur. The upper velocity bound is dictated mostly
by experimental observation; pneumatic transport of the particles out the top of the vessel is unacceptable.
Although there are many parameters of interest within the model, this setup focuses on the main purpose: powder
moisture content. In this case, the drying process begins with the powder moisture content at 45% by dry basis and is
considered complete when the powder reaches 3%.

44
4.3.4.1 Objective Function

The objection function used in this setup is relatively simple. The expression pits daily energy consumption, Qday ,
against daily production rates, mday , by introducing monetary values for each variable.

OF = Qday Energy($/J) − mday Revenue($/kg) (4.8)

The daily energy consumption, Qday is calculated using the model and describes the total amount of energy used to dry
the powder to the 3% endpoint. Consumption is based on heating and pressurizing the inlet gas with respect to standard
conditions. The daily amount of dried product, mday , is established by using the endpoint time and determining the
daily number of 3 kg batches that could be completed.
The above setup requires two arbitrary monetary terms attached to each process component. In this way, the opti-
mization serves to find the relationship between energy costs and dry powder production. In a real world application,
the fluidized bed drying process is only a segment of the entire tablet production operation. Therefore, revenue values
would vary and would not be directly tied to production rates from fluidized beds. On the energy side, costs can be
influenced by compressor and heating efficiency or assisted with thermal energy from a concurrent process. Because
of this, optimization is completed across a range of energy costs and product revenues in order to provide observations
for various conditions.
To discuss the range of objective functions and their respective optimization solutions, a ratio variable is introduced.
Instead of discussing the two monetary terms separately, a ratio of the monetary terms can be introduced, written as:

Revenue($/kg)
Profit Ratio = , (4.9)
Energy($/J)

and utilized in the next section. The ”Profit Ratio” value can be applied over a range of values to observe how it affects
the optimization solution. The ratio can be set to a relatively low value to imply high energy costs, favouring saving
energy consumption. Consequently, a higher value implies high profit rates, favouring faster drying times to maximize
throughput. Lastly, a ”Profit Ratio” value within the transitional range between these two extremes will capture the
solution when trying to balance the two opposing terms.

4.4 Results and Discussion

4.4.1 Energy and Exergy Analysis

The efficiency equations were applied to a similar range of inlet gas velocities and temperatures as the optimization
space. The bounds included parameters that would not be suitable for fluidization; however, the range displays a
broader view of the model’s analysis.

45
Table 4.2: Inlet Gas Parameter Bounds

Variable Bounds

Initial Velocity 0.6 - 2.0 m/s


Initial Temperature 293.15 - 393.15 K

Energetic and exergetic efficiencies were recorded over a grid search of the inlet gas variables. A 30 × 30 grid size
is implemented to cover the range. Figure 4.2 and Figure 4.3 display each respective efficiency over these ranges.

Figure 4.2: Energy Efficiency

Figure 4.2 displays the various energetic efficiencies calculated from the model. The grid search did not expose
any optimum constant inlet gas parameters within a reasonable range. A trench is created at approximately 306 K,
showing a minimum efficiency that is fairly independent of the inlet gas velocity. As the temperature increases, the
efficiency climbs. Moreover, inlet gas velocity gradually becomes more of a factor, improving efficiency as the velocity
increases. In terms of energy efficiency, the model implies that the optimum configuration is at the highest constant
inlet gas velocity and temperature practical. However, the fluidized bed drying is a relatively inefficient process, with
model predictions of 45-55% energy efficiency across the grid search. Consequently, the small increases to efficiency
may not offset the associated additional energy cost.

46
Figure 4.3: Exergy Efficiency

The exergetic efficiencies found in Figure 4.3 of the fluidized bed drying process is even lower in comparison to its
energetic counterpart. Exergetic efficiencies are found within 2.5%-7% for the reasonable operating range. Exergetic
efficiency does not depend on inlet gas velocity and is based predominantly on inlet gas temperature. This is intuitive
for exergy; as the inlet gas temperature increases, the quality of the energy being introduced to the system increases.

4.4.2 Five-Variable Optimization

The optimization results can be organized into three sections based on the objective function; the two extremes (pri-
oritizing saving energy or maximizing throughput) and the transition period where optimization solutions changeover.
These sections are dictated by the Profit Ratio value. A low value promotes saving energy, whereas a high value
encourages production.

4.4.2.1 Profit Ratio Bounds

At the two extremes, the optimization solution reaches the bounds of the inlet gas parameters. Based on the design
space bounds, the drying process duration can vary greatly, with the fast-production batch taking only 23% of the time
in comparison to the energy-saving batch. At both extremes, no step change is implemented.

47
Table 4.3: Optimization results at Profit Ratio bounds

Optimization Type Profit Ratio Inlet Gas Velocity Inlet Gas Tempera- Drying Time Energy
ture

Energy Saving 1510000 0.9 m/s 303.15 K 2 hrs 49 min 3.63 MJ


Production Throughput 1521800 2.0 m/s 343.15 K 39 min 4.34 MJ

Table 4.3 displays the main drying specifications at the Profit Ratio extremes. The lower bound solution from the
model standpoint is perfectly reasonable; however, does not correlate with real-world application. With this particular
experimental setup, operating the fluidized bed 0.9 m/s will not be able to overcome the minimum fluidization velocity
necessary when the powder is wet. This is a good example of the limitations of the two-phase model. Regardless of the
design space variable bounds, at low Profit Ratios, the optimization solution will always converge to the lowest inlet
gas parameters. In other words, raising the inlet gas velocity bound to an acceptable value for fluidization will give us
the solution we are looking for, but it will not convey any meaningful conclusions.
Increasing the significance of the revenue within the objective function places all the importance on production
rates. A high Profit Ratio causes the optimization solution to dry the powder as fast as possible. Initial inlet gas
parameters are set to the upper velocity and temperature ranges. The process is quick, with the model predicting the
batch completed within 39 minutes, using approximately 4.34 MJ. In real world application, this would be ideal to
ensure fluidization, but issues may arise with high gas temperatures at the latter portions of the process. The particle
moisture and temperature profiles of the two Profit Ratio extremes can be seen in Figure 4.4.

48
Figure 4.4: Moisture and temperature profiles of optimization solution at Profit Ratio bounds

4.4.2.2 Profit Ratio Transition Point

At a certain Profit Ratio range, there is a transition section between the energy-saving solution and high-energy solu-
tions.

Table 4.4: Optimization results in transition area for optimization function

Initial Velocity Initial Temperature Final Velocity Final Temperature tchange

1.97 m/s 343.15 K 1.96 m/s 305.68 K 37 min 17 sec

Table 4.4 displays a solution found at the Profit Ratio of 1508000, found within the transitional range. The optimizer
consistently settled into a similar configuration throughout the various Profit Ratios in the transitional range. The initial
inlet gas temperature and velocity are set to the high values and fall back to a lower temperature in the final 4 minutes of
the drying process, changing at 3.7% powder moisture content. These results were susceptible to the user-defined initial
variable values changing the solution; it was necessary to run the Global Optimizer at various initial configurations to
ensure the optimum solution was found. Figure 4.5 displays the optimization solution at the Profit Ratio of 1508000.

49
Figure 4.5: Moisture and temperature profiles of optimization solution within transition range

The inlet gas is set at the highest possible temperature and velocity and is not lowered until near the endpoint of
drying. In Figure 4.5, the time of the step change is found at approximately 37 minutes, with the endpoint occurring
around 41 minutes. The process consumes about 4.30 MJ. This is consistent for other Profit Ratios found within the
transition range. From these solutions, it seems that once the energy cost reaches a certain point (relative to profit), the
most efficient solution is to operate the fluidized bed at the highest temperature and velocity possible, and a decrease
in inlet gas temperature is advantageous at the end of the drying process.
The optimization solution is interesting because it appears that a step change to lower the inlet gas velocity and
temperature does not help the process in a significant manner. This is intuitive during the constant drying rate phase.
Lowering the inlet gas parameters would not be useful; the surface moisture is readily available, and the fluidized bed
should operate at the highest parameters possible. However, as the drying rate decreases in the falling drying rate
phase, the drying efficiency continues to lessen, giving an opportunity to turn down the inlet gas parameters. However,
the optimization solution shows that this section still benefits from the faster and hotter gas flow, only turning the
parameters down at the very end of the drying process.
The result still addresses an issue with the batching nature of the process. Although the energy savings may
be negligible (1% savings in comparison to no step-change), the lowering of the inlet gas parameters may be more
significant if the fluidized bed is still operating after the moisture endpoint of 3%. A step-change to ensure lower
energy utility at the last portion of the process will lower energy waste while still ensuring the powder is dried to an
acceptable level. For example, in the solution found at Profit Ratio = 1508000, lowering the inlet gas parameters from
the high configuration to the low configuration equates to 59.2 kJ/min savings (the lower configuration using roughly
46% of the energy compared to the initial inlet gas flow). This energy loss will become more substantial the longer

50
the fluidized bed unnecessarily stays on. Moreover, powder temperature can be maintained at a lower level, mitigating
thermal degradation.

4.4.2.3 Step-Change

The high low configuration is further examined by moving the step change time. Because of the high initial parameters
and low final parameters, examining the model through a range of step changes indirectly displays the relationship
between energy-saving and production priorities. Figure 4.6 displays the objective function value across the step
changes using the high/low configuration at Profit Ratio =1508000.

Figure 4.6: Objective value vs. step change with high/low configuration

Figure 4.6 shows a minimum value at the previously optimized step change of approximately 37 minutes. Lowering
the inlet gas parameters at any time before this has adverse effects on the objective function; greatly increasing across
the falling drying rate period and levelling out if the step change is during the constant drying rate period. Finally, the
plot displays a relative local minimum at the very left, showing an energy saving configuration. This ”always low”
configuration shows a better efficiency in regard to the objective function in comparison to any step change within the
constant drying rate period, further indicating that a step-change within the middle of the drying process is ineffective.

4.5 Conclusions
In this work, a phenomenological model was employed to further investigate an experimental scale fluidized bed
drying process. The model examined energy and exergy efficiency across a range of operating conditions, followed
by optimization analysis of the inlet gas parameters utilizing a single step-change in the optimization design. In
general, the analysis demonstrated that higher inlet gas velocity and temperature were ideal, regardless of the inherent
inefficiency of the drying process.

51
Energy and exergy analysis of the drying process revealed highest efficiencies at the highest inlet gas velocity
and temperature. Energy efficiency is dependent on both inlet gas velocity and temperature, ranging between 45-55%
across the inlet gas operating range. In contrast, exergy efficiency is only dependent on temperature and spans between
2.5-7%.
The optimization setup used Matlab’s Global Optimization Toolbox and optimized the inlet gas parameters based
on energy usage and dry powder throughput. Across a span of objective functions, the optimizer demonstrated that a
step-change did not provide any significant savings, and in fact, could lead to a more inefficient configuration if placed
within the drying process. Based on the optimization solutions, the ideal operation would be to run the fluidized bed
at the highest possible inlet gas velocity and temperature until near the endpoint of the drying process. Lowering inlet
gas parameters near the endpoint is beneficial but may be hard to time. Regardless, ensuring the initial high conditions
are turned down near or at the endpoint saves a significant enough energy to justify.

4.6 Acknowledgements
The authors gratefully acknowledge financial support from NSERC and the University of Saskatchewan.

52
C HAPTER 5
C ONCLUSIONS AND R ECOMMENDATIONS

5.1 Summary of Results


Through this work, a two-phase model has verified and further substantiated by previous experimental results [4].
Various models were considered in the initial stages. The complexity of fluidization invites extensive models to be
studied however, many of these more ’complicated’ models either could not be feasibly verified in time, or they could
not fully capture the process at the scale of the experimental data. Ultimately, a two-phase model involving five distinct
regions of mass and heat transfer and incorporating emulsion/dense and bubble phases to characterize the powder bed
was chosen.
The experimental results were comprised of moisture profiles for the pharmaceutical powder and temperature data
for the outlet gas of the fluidized bed vessel. The experiment covered three inlet gas velocities over four inlet gas
temperatures, making up a total of 12 moisture profiles. The model agreed with the moisture profiles. Modelling of
the powder temperature coincides with the outlet gas temperatures recorded although no direct details are available for
comparison.
Once verified, the model provided many insights on the drying process. Energy and exergy efficiency varies with
varying inlet gas velocity and temperature. Inlet gas velocity has an effect on energy efficiency and not on exergy.
Increasing inlet gas temperature increases both energy and exergy efficiency. From an energy standpoint, the optimum
inlet gas configuration is found at the highest possible velocity and temperature, showing an increase energy efficiency
of 4.0% in comparison to the minimum possible configuration.
Optimization methods offered some interesting solutions under various objective function configurations. The
optimum inlet gas configuration settled to the outer bounds in two cases.

1. When conserving energy consumption, the optimum configuration is at the lowest inlet gas velocity and tem-
perature. This is in contrast to energy efficiency results; the inefficiency of the process overrides any efficiency
increases that can occur when increasing temperature.

2. When production throughput of dry powder is the focus, the optimum configuration predictably moves to the
highest inlet gas velocity and temperature to dry the powder as fast as possible.

The balance of the above terms provides a more interesting solution. In this transition section, the optimization
solution utilizes the inlet gas step change near the end of the drying process. The initial inlet gas velocity and temper-

53
ature are set to high values, dropping to a lower configuration with minutes left before the powder is sufficiently dried.
The optimization only incurs a 1% energy savings, but it also displays the importance of lowering inlet gas parameters
nearing the end of the drying process.
The optimization study also uncovered many insufficiencies of the model. The bulk perspective of the model
allows for simpler computations, but it suffers from too broad of view. The lack of any fluidization description gives
no warning of defluidization issues that would occur experimentally. Moreover, no fluidization effects on drying are
being considered.

5.1.1 Contributions

The contributions of this work can be summarized as follows.

1. The two-phase fluidized bed drying model was verified with an entirely different experimental setup. The flu-
idized bed dimensions, material, and inlet gas distributor were different. Different pharmaceutical powder was
also used, altering many physical parameters of the particles used within the model.

2. The optimization analysis displays the utility of the model for investigating the drying process, as well as the
energetic and exergetic efficiencies.

5.2 Conclusions
The following conclusions have been developed through this work.

1. The two-phase model previously studied by Gagnon [29] was successfully used to examine and analyze the
experimental setup found in this study.

2. The fluidized bed drying process of pharmaceutical powders is a relatively wasteful process. Energy and exergy
efficiency analysis confirms this, ranging between 47-53% energetic efficiency and 2-7% exergetic efficiency
across the design space. In both perspectives, results advocate for the highest possible inlet gas temperature and
velocity to maximize efficiency.

3. A one-time step change of the inlet gas parameters to optimize the drying process is relatively insignificant for
pharmaceutical fluidized bed drying. With this in mind, optimization still considers lowering of gas parameters
useful near the end of drying process, especially if the dryer is left on past the drying endpoint.

5.3 Recommendations
This study provided some utilization of a successful two-phase model for fluidized bed drying. However, the limits
of the model are fairly constrictive. Much of the phenomena seem to need more complex representation and higher
computational resources than the two-phase model operates with. However, improvements can be appended to the
two-phase model to account for fluidization and its effects on the drying profile. Some possible avenues are as follows:

54
1. Separating the dense phase into sections

2. Finding common behaviour and determine coefficients for various fluidized bed configurations

Moreover, usage of a simplified approach such as the two-phase model could benefit from industry guidance. The
model may be effective at providing useful suggestions to industry-scale fluidized bed processes. Various correlating
variables can be modified easily and adjusted to fit the drying data. Once verified, specific tests relevant for the
operating issues can be completed. For example, the single step-change configuration found in this study could be a
possibility. Other parameters, such as batch size or multiple step-changes could be examined. In this way, the utility of
the model will be as applicable as possible.

55
B IBLIOGRAPHY

[1] Jonas Berggren. Engineering of pharmaceutical particles: modulation of particle structural properties, solid-state
stability and tabletting behaviour by the drying process. 5:58, 2003.

[2] Dilip M. Parikh. Handbook of pharmaceutical granulation technology. Handbook of Pharmaceutical Granulation
Technology, Second Edition, pages 1–625, 2005.

[3] Chris Vervaet and Jean Paul Remon. Continuous granulation in the pharmaceutical industry. Chemical Engineer-
ing Science, 60(14):3949–3957, 2005.

[4] Milad Taghavivand, Kwangseok Choi, and Lifeng Zhang. Investigation on drying kinetics and tribocharging
behaviour of pharmaceutical granules in a fluidized bed dryer. Powder Technology, 316:171–180, 2017.

[5] A.S. Mujumdar. Handbook of Industrial Drying. M. Dekker, 1987.

[6] Ray Cocco, S. B.Reddy Karri, and Ted Knowlton. Introduction to fluidization. Chemical Engineering Progress,
110(11):21–29, 2014.

[7] Wan Ramli Wan Daud. Fluidized Bed Dryers - Recent Advances. Advanced Powder Technology, 19(5):403–418,
2008.

[8] Maysam Saidi, Hassan Basirat Tabrizi, and John R. Grace. A review on pulsed flow in gas-solid fluidized beds
and spouted beds: Recent work and future outlook, 2019.

[9] Baoguo Hao and Hsiaotao T. Bi. Forced bed mass oscillations in gas-solid fluidized beds. Powder Technology,
149(2-3):51–60, 2005.

[10] Michael Wormsbecker and Todd Pugsley. The influence of moisture on the fluidization behaviour of porous
pharmaceutical granule. Chemical Engineering Science, 63(16):4063–4069, 2008.

[11] B. Paláncz. A mathematical model for continuous fluidized bed drying. Chemical Engineering Science,
38(7):1045–1059, 1983.

[12] Emrah Özahi and Hacimurat Demir. Drying performance analysis of a batch type fluidized bed drying process
for corn and unshelled pistachio nut regarding to energetic and exergetic efficiencies. Measurement: Journal of
the International Measurement Confederation, 60:85–96, 2015.

56
[13] Francis Gagnon, André Desbiens, Éric Poulin, Pierre Philippe Lapointe-Garant, and Jean Sébastien Simard. Non-
linear model predictive control of a batch fluidized bed dryer for pharmaceutical particles. Control Engineering
Practice, 64(August 2016):88–101, 2017.

[14] A. H. Zahed, J.-X. Zhu, and J. R. Grace. Modelling and simulation of batch and continuous fluidized bed dryers.
Drying Technology, 13:1-2:1–28, 1995.

[15] Can Ertekin and M. Ziya Firat. A comprehensive review of thin-layer drying models used in agricultural products.
Critical Reviews in Food Science and Nutrition, 57(4):701–717, 2017.

[16] Alexander Krok, Nuno Vitorino, Jianyi Zhang, Jorge Ribeiro Frade, and Chuan Yu Wu. Thermal properties of
compacted pharmaceutical excipients. International Journal of Pharmaceutics, 534(1-2):119–127, 2017.

[17] Helen Tanfara, Todd Pugsley, and Conrad Winters. Effect of particle size distribution on local voidage in a
bench-scale conical fluidized bed dryer. Drying Technology, 20(6):1273–1289, 2002.

[18] Michael Wormsbecker, Todd Pugsley, J. Ruud Van Ommen, John Nijenhuis, and Rob Mudde. Effect of distrib-
utor design on the bottom zone hydrodynamics in a fluidized bed dryer using 1-D X-ray densitometry imaging.
Industrial and Engineering Chemistry Research, 48(15):7004–7015, 2009.

[19] M.A. van der Hoef, M. van Sint Annaland, N.G. Deen, and J.A.M. Kuipers. Numerical Simulation of Dense
Gas-Solid Fluidized Beds: A Multiscale Modeling Strategy. Annual Review of Fluid Mechanics, 40(1):47–70,
2008.

[20] G. A. Bokkers, J. A. Laverman, M. van Sint Annaland, and J. A.M. Kuipers. Modelling of large-scale dense gas-
solid bubbling fluidised beds using a novel discrete bubble model. Chemical Engineering Science, 61(17):5590–
5602, 2006.

[21] J.A.M. Kuipers and W P M Van Swaaij. Computational Fluid Dynamics Applied to Chemical Reaction Engi-
neering. In Advances in Chemical Engineering, Vol. 24, pages 227–319. 1998.

[22] Hao Chen, Subham Rustagi, Emily Diep, Timothy A.G. Langrish, and Benjamin J. Glasser. Scale-up of fluidized
bed drying: Impact of process and design parameters. Powder Technology, 339:8–16, 2018.

[23] Emmanuela Gavi. Application of a mechanistic model of batch fluidized bed drying at laboratory and pilot scale.
Drying Technology, 0(0):1–17, 2019.

[24] D Kunii and Octave Levenspiel. Fluidization Engineering. 1991.

[25] Leszek Stepien. Lecture notes in fluidization. January 2015.

[26] Jannatul Azmir, Qinfu Hou, and Aibing Yu. Discrete particle simulation of food grain drying in a fluidised bed.
Powder Technology, 323:238–249, 2018.

57
[27] Mohammad Mohseni, Alex Kolomijtschuk, Bernhard Peters, and Marc Demoulling. Biomass drying in a vi-
brating fluidized bed dryer with a Lagrangian-Eulerian approach. International Journal of Thermal Sciences,
138(January):219–234, 2019.

[28] Mingzhong Li and Stephen Duncan. Dynamic model analysis of batch fluidized bed dryers. Particle and Particle
Systems Characterization, 25(4):328–344, 2008.

[29] Francis Gagnon, Jocelyn Bouchard, André Desbiens, and Éric Poulin. Development and validation of a batch
fluidized bed dryer model for pharmaceutical particles. Drying Technology, 39(5):620–643, 2021.

[30] Mingzhong Li and Stephen Duncan. Model-based nonlinear control of batch fluidized bed dryers. Particle and
Particle Systems Characterization, 25(4):345–359, 2008.

[31] J. A. Villegas, S. R. Duncan, H. G. Wang, W. Q. Yang, and R. S. Raghavan. Optimal operating conditions for a
batch fluidised bed dryer. IET Control Theory and Applications, 4(2):294–302, 2010.

[32] H. G. Wang, T. Dyakowski, P. Senior, R. S. Raghavan, and W. Q. Yang. Modelling of batch fluidised bed drying
of pharmaceutical granules. Chemical Engineering Science, 62(5):1524–1535, 2007.

[33] N.S. Grewal and S.C. Saxena. Heat transfer between a horizontal tube and a gas-solid fluidized bed. International
Journal of Heat and Mass Transfer, 23(11):1505–1519, 1980.

[34] M.I. Osman, S.N. Upadhyay, and S.C. Saxena. Heat transfer investigations for in-bed and freeboard vertical
u-tubes in a fluidized bed at moderate temperatures. Energy, 7(5):465–472, 1982.

[35] Stuart W. Churchill and Humbert H.S. Chu. Correlating equations for laminar and turbulent free convection from
a vertical plate. International Journal of Heat and Mass Transfer, 18(11):1323–1329, 1975.

[36] T. E. Broadhurst and H. A. Becker. Onset of fluidization and slugging in beds of uniform particles. AIChE
Journal, 21(2):238–247, 1975.

[37] Tayyeb Nazghelichi, Mohammad Hossein Kianmehr, and Mortaza Aghbashlo. Thermodynamic analysis of flu-
idized bed drying of carrot cubes. Energy, 35(12):4679–4684, 2010.

[38] Md Sazzat Hossain Sarker, Mohd Nordin Ibrahim, Norashikin Abdul Aziz, and Mohd Salleh Punan. Energy and
exergy analysis of industrial fluidized bed drying of paddy. Energy, 84:131–138, 2015.

[39] S. Syahrul, F. Hamdullahpur, and I. Dincer. Thermal analysis in fluidized bed drying of moist particles. Applied
Thermal Engineering, 22(15):1763–1775, 2002.

[40] M. R. Assari, H. Basirat Tabrizi, and E. Najafpour. Energy and exergy analysis of fluidized bed dryer based on
two-fluid modeling. International Journal of Thermal Sciences, 64:213–219, 2013.

58
A PPENDIX A
M ATLAB C ODE

A.1 Model

function dy = f (t ,y , i nl et Ga sPa ra me te rs )
%{
Two - phase fluidized bed drying model for pharmaceutical powder .
Model to be used to validate experimental data .

This file contatins the model system of equations and all other
necessary calculations .

Marcus Lau
mll668 @ mail . usask . ca
%}
%% Inlet Gas Specifications and number of bubble layers
global X_in
global nLayers
T_in = i nl et Ga sP ar am et er s (1) ;
v_in = i nl et Ga sP ar am et er s (2) ;
% end of Inlet Gas

%% Optimized variables
phi = 0.9; % Particle Sphericity
c_s = 1200; % Heat Capacity of particle
d_p = 1E -3; % particle diameter
d_b = 0.035; % bubble diameter
% end of optimized variables

%% Constant Variables

59
P_v = 101389; % Vessel Pressure
g = 9.81; % Gravity constant
c_g = 1005; % Heat Capacity of air
c_v = 1996; % Heat Capacity of water vapour
c_w = 4200; % Heat Capacity of water
rho_g = 1.27; % Density of air
rho_s = 1560.5; % Density of dry particle
rho_w = 992.3; % Density of water in particle
lambda_g = 0.2735; % thermal conductivity of air
u_g = 0.00001907; % dynamic viscosity of air
T_a = 293.15; % ambient temperature
T_wv = 293.15; % Temperature of water vapor
D_wa = 0.00002733; % Diffusivity of water
X_pc = 0.08; % Critical Moisture point
rho_pd = rho_s * rho_w / ( rho_s * X_pc + rho_w ) ; % Dry particle density

% Vessel Dimensions ( currently not conical )


d_v = 0.115; % vessel diameter
d_vo = 0.135; % outer vessel diameter
d_vr = 0.25; % freeboard vessel diameter
d_vro = 0.28; % freeboard outer vessel diameter
m_v = 10; % mass of vessel
c_v = 900; % vessel heat capacity
z_v = 1.53; % total vessel height
m_st = 2.1; % -( kg ) mass of solids in vessel
S_f = pi * d_v ^2 / 4; % -( m ^2) Cross - sectional area of FBD
% end of Constant Variables

%% System of Equations
dy = zeros (2* nLayers +8 ,1) ;

% Variable List for DAE system


X_p = y (1) ; % Particle Moisture
T_p = y (2) ; % Particle Temperature
T_v = y (3) ; % Vessel Temperature
X_d = y (4) ; % Dense phase humidity

60
X_bave = y (5) ; % Bubble phase humidity
X_b1 = y (6) ; % Bubble phase humidity (1 st layer )
T_b1 = y (7) ; % Bubble phase temperature (1 st layer )
T_d = y (2* nLayers +6) ; % Dense phase temperature
T_out = y (2* nLayers +7) ; % Outlet temperature
X_out = y (2* nLayers +8) ; % Outlet humidity

rho_p = rho_pd * (1 + X_p ) ; % Particle density ( with water )

% velocity and voidage space due to fluidization and bubbbles


e_mf = 0.586 * phi ^( -.72) * ( u_g ^2 / ( rho_g * ( rho_p - rho_g ) * g * d_p ^3)
) ^(0.029) * ...
( rho_g / rho_p ) ^(0.021) ;
v_mf = d_p ^2 * ( rho_p - rho_g ) * g * e_mf ^3 * phi ^2 / (150 * u_g * (1 -
e_mf ) ) ;
v_b = ( v_in - v_mf ) + 0.711 * ( g * d_b ) ^(1/2) ;
delta = ( v_in - v_mf ) / v_b ;
e_f = (1 - delta ) * e_mf + delta ;

z_f = m_st / ( S_f * rho_pd * (1 - e_f ) ) ;


delta_z = z_f / nLayers ; % bubble phase descretization

% Heat and Mass Transfer Coefficients


% heat - particle stagnant film to dense phase
h_epf = ( lambda_g / d_p * (2 + 0.6 * ( d_p * v_in * rho_g / u_g ) ^(1/2) * (
c_g * u_g / lambda_g ) ^(1/3) ) ) ;
% mass - particle stagnant film to dense phase
k_pfe = h_epf / c_g ;

K_bc = 4.5 * v_mf / d_b + 5.85 * D_wa ^(0.5) * g ^(0.25) / d_b ^(1.25) ;
K_ce = 5.71 * ( e_mf * D_wa * g ^(0.5) / d_b ^(2.5) ) ^(0.5) ;
% mass - dense to bubble phase
K_eb = 1 / (1/ K_bc + 1/ K_ce ) ;
H_bc = 4.5 * v_mf * rho_g * c_g / d_b + 5.85 * ( lambda_g * rho_g * c_g )
^(0.5) * g ^(0.25) / d_b ^(1.25) ;
H_ce = 5.71 * ( e_mf * lambda_g * rho_g * c_g * g ^(0.5) / d_b ^(2.5) ) ^(0.5) ;

61
% heat - bubble to dense phase
H_be = 1 / (1/ H_bc + 1/ H_ce ) ;

VesselMag = 1; % set to 0 for adiabatic system


% heat - dense phase to vessel wall
h_ev = (47 * lambda_g * (1 - e_f ) / d_v * ( v_in * d_v * u_g / ( rho_pd * d_p
^3 * g ) ) ^(0.325) * ...
( rho_s * c_s * d_v ^(1.5) * g ^(0.5) / lambda_g ) ^(0.23) *...
( c_g * u_g / lambda_g ) ^(0.3) ) * VesselMag ;
% heat - freeboard region to vessel wall
h_rv = (3.95 * lambda_g / d_vr * ( d_vr * v_in * rho_g / u_g ) ^(0.55) * ( c_g
* u_g / lambda_g ) ^(1/3) ) * VesselMag ;

P_sat = 100000 * exp (13.869 - (5173 / ( T_p ) ) ) ;

a = 2.604; % dimensionless variables for expirical equation


kappa = 0.01; % for surface moisture in falling rate period

% Stagnant film moisture content


if ( X_p >= X_pc )
X_s = 0.622 * ( P_sat / ( P_v - P_sat ) ) ;
elseif ( X_p < X_pc )
X_s = X_p ^ a * ( X_pc ^ a + kappa ) / ( X_pc ^ a * ( X_p ^ a + kappa ) ) * 0.622 *
( P_sat / ( P_v - P_sat ) ) ;
end

Gr = g * z_v ^3 * abs ( T_v - T_a ) * rho_g ^2 / (0.5 * ( T_v + T_a ) * u_g ^2) ;
% heat - vessel wall to ambient air
h_va = lambda_g / z_v * (0.825 + (0.387 * ( Gr * c_g * u_g / lambda_g )
^(1/6) /...
(1 + (0.492 * lambda_g / ( c_g * u_g ) ) ^(9/16) ) ^(8/27) ) ) ^2;

H_evap = (3168 - 2.4364 * T_p ) * 1000;

% X_p differential
dy (1) = -6 * k_pfe * rho_g * ( X_s - X_d ) / ( phi * d_p * rho_pd ) ;

62
% T_p differential
dy (2) = 6 / ( phi * d_p * rho_pd * ( c_s + X_p * c_w ) ) * ( h_epf * ( T_d - T_p
) - k_pfe * rho_g * ( X_s - X_d ) ...
* ( H_evap + ( c_v - c_w ) * ( T_p - T_wv ) ) ) ;

% T_v differential
dy (3) = pi / ( m_v * c_w ) * ( d_v * z_f * h_ev * ( T_d - T_v ) + d_vr * ( z_v -
z_f ) * h_rv * ( T_out - T_v ) ...
- d_vro / d_vr * z_v * h_va * ( T_v - T_a ) ) ;
% X_d
dy (4) = 6 * k_pfe * (1 - e_f ) / ( phi * d_p ) * ( X_s - X_d ) - delta * K_eb * (
X_d - X_bave ) - v_mf * ...
( X_d - X_in ) / z_f ;

% X_bAverage ( analytical solution )


dy (5) = X_d - X_bave + ( X_d - X_in ) * v_b / ( K_eb * z_f ) * ( exp ( - K_eb *
z_f / v_b ) - 1) ;

% initial layer for X_b and T_b


dy (6) = X_in - X_b1 + delta_z * K_eb * ( X_d - X_b1 ) / v_b ;
dy (7) = T_in - T_b1 + delta_z * ( T_d - T_b1 ) * ...
( H_be / rho_g + K_eb * c_v * ( X_d - X_b1 ) ) / ( v_b * ( c_g + X_b1 * c_v )
);
T_btot = T_b1 ;
X_btot = X_b1 ;

% for additional layers of X_b ( z ) and T_b ( z )


if nLayers >= 2
for a = 8:2:2* nLayers +4
dy ( a ) = y (a -2) - y ( a ) + delta_z * K_eb * ( X_d - y ( a ) ) / v_b ;
dy ( a +1) = y (a -1) - y ( a +1) + delta_z * ( T_d - y ( a +1) ) * ...
( H_be / rho_g + K_eb * c_v * ( X_d - y ( a ) ) ) / ( v_b * ( c_g + y ( a ) *
c_v ) ) ;
T_btot = T_btot + y ( a +1) ;
X_btot = X_btot + y ( a ) ;

63
end
end
T_bAverage = T_btot / nLayers ;
X_bAverage = X_btot / nLayers ;
% T_d
dy (2* nLayers +6) = delta * H_be / rho_g * ( T_bAverage - T_d ) + 6 * (1 - e_f )
/ ( phi * d_p ) * ...
( k_pfe * c_v * ( X_s - X_d ) + h_epf / rho_g ) * ( T_p - T_d ) - v_mf / z_f *
( c_g + X_in * c_v ) * ( T_d - T_in ) - ...
4 * h_ev /( d_v * rho_g ) * ( T_d - T_v ) ;
% T_out
dy (2* nLayers +7) = delta * v_b * ( c_g + X_out * c_v ) * y (2* nLayers +5) +
v_mf * ( c_g + X_d * c_v ) * T_d ...
+ 4 * ( z_v - z_f ) * h_rv * T_v / ( d_vr * rho_g ) - ( v_in * ( c_g + X_out
* c_v ) + (4 * ( z_v - z_f ) * h_rv /( d_vr * rho_g ) ) ) * T_out ;
% X_out
dy (2* nLayers +8) = v_mf * X_d + delta * v_b * y (2* nLayers +4) - v_in * X_out
;
% end of System of Equations

A.2 Model Controller and Post Calculations

%{
Two - phase fluidized bed drying model for pharmaceutical powder .
Model to be used to validate experimental data . Currently set up
accomodates one step change to the inlet gas velocity , v_in , and
temperature , T_in , during the process .
Marcus Lau
mll668 @ mail . usask . ca
%}

load ( ' Experiments . mat ')

global X_in ; % Inlet gas moisture content


global nLayers ; % Number of bubble phase layers

64
X_in = 0.0001; nLayers =40;

RevenueRatio = 1.508 e +06;


c = [ 3.43 e +02 , 1.97 , 3.06 e +02 , 1.95 , 5.23 e +03];

T_in = c (1) ;
v_in = c (2) ;
T_in2 = c (3) ;
v_in2 = c (4) ;
tChange = c (5) ;
tFinal = 12000;

in le tG as Pa ra me te rs = [ T_in v_in ];
% end of Inlet Gas Parameters
%% Constant Variables ( may not be used )
T_ref = 273.15; % Reference Temperature
P_v = 101389; % Vessel Pressure
phi = 1; % Particle Sphericity
c_g = 1005; % Heat Capacity of air
c_v = 1996; % Heat Capacity of water vapour
c_s = 1098; % Heat Capacity of particle
c_w = 4200; % Heat Capacity of water
A_orifice = 0.01038689; % Area of distribution plate
rho_g = 1.27; % Density of air
rho_s = 1560.5; % Density of dry particle
rho_w = 992.3; % Density of water in particle
X_pc = 0.08; % Critical Moisture point
rho_pd = rho_s * rho_w / ( rho_s * X_pc + rho_w ) ; % ( kg / m ^3) Dry particle
density
% end of Constant Variables
%% Initial Conditions
X_pinit = 0.45; % Particle moisture content
T_pinit = 293.15; % Particle temperature
T_vinit = 293.15; % Vessel temperature
X_init = 0; % General initial moisture content

65
T_init = 293.15; % General initial temperature

% Initial conditions for DAE system


Init = zeros (2* nLayers +8 ,1) ;
Init (1) = X_pinit ; % Particle moisture content
Init (2) = T_pinit ; % Particle temperature
Init (3) = T_vinit ; % Vessel temperature
Init (4) = X_init ; % Dense phase moisture content
Init (5) = X_init ; % Average bubble phase moisture content
Init (2* nLayers +6) = T_init ; % Dense phase temperature
Init (2* nLayers +7) = T_in ; % Outlet temperature
Init (2* nLayers +8) = X_init ; % Outlet moisture content

for j = 6:2:2* nLayers +4


Init ( j ) = X_init ; % Bubble phase layers moisture content
Init ( j +1) = T_init ; % Bubble phase layers temperature
end
%
%% Compute DAE until inlet gas step change
options = odeset ( ' RelTol ' ,1e -5 , ' AbsTol ' ,1e -5 , ' MassSingular ' , ' yes ') ;

[t , y ] = ode15s (@( t , y ) fFBDEnergy (t ,y , i nle tG as Pa ra me te rs ) , [0:1: tChange ] ,


Init , options ) ;
% end of 1 st compute
%% Compute DAE after inlet gas step change
in le tG as Pa ra me te rs = [ T_in2 v_in2 ];

% Initial conditions ( from the end of the 1 st ode15s solution )

Init2 = y ( end ,:) ;

[ t2 , y2 ] = ode15s (@( t , y ) fFBDEnergy (t ,y , i nle tG as Pa ra me te rs ) , [ tChange :1:


tFinal ] , Init2 , options ) ;
% end of 2 nd compute
%% Consilidating solution data
y2 (1 ,:) = [];

66
t2 (1 ,:) = [];
yall = cat (1 ,y , y2 ) ;
tall = cat (1 ,t , t2 ) ;

for i = 1:1: round ( tChange ) +1


T ( i ) = T_in ;
v ( i ) = v_in ;
end
for i = round ( tChange ) +2:1: tFinal +1
T ( i ) = T_in2 ;
v ( i ) = v_in2 ;
end
% end of Considilating
%% Post - Calculations
% Relative Humidity
for i = 1:1: size ( yall ,1)
RH_out ( i ) = yall (i ,2* nLayers +8) *1000* yall (i ,2* nLayers +7) / (6.112 *
2.1674 * exp (17.67 * ( yall (i ,2* nLayers +7) - 273.15) / ( yall (i ,2*
nLayers +7) - 29.65) ) ) ;
end

% Energy Analysis
for i = 1:1:4
energy (1 , i ) = 0;
end

for i = 2:1: size ( yall ,1)


energy (i ,1) = ( yall (i ,1) - yall (i -1 ,1) ) * 1.74 ; %
moisture removal rate
energy (i ,2) = energy (i ,1) * (1000 * (3168 - 2.4364 * yall (i ,2) ) ) ; %
energy used for vaporization
energy (i ,3) = rho_g * v ( i ) * c_g * ( T ( i ) - T_ref ) * A_orifice ; %
energy inputted into system
energy (i ,4) = - energy (i ,2) / energy (i ,3) ; %
energy efficiency
end

67
% Average energy efficiency until particle reaches 0.02 moisture content
for i = 1:1: size ( yall ,1)
if yall (i ,1) < 0.03
effectiveStop = i ;
break
end
effectiveStop = size ( yall ,1) ;
end

averageEnEff = mean ( energy (2: effectiveStop , 4) ) ;

% Convert to ratio of energy not consumed by drying


energyLossRatio = 1 - averageEnEff ;

% Exergy Analysis
for i = 1:1:3
exergy (1 , i ) = 0;
end

for i = 2:1: size ( yall ,1)


exergy (i ,1) = energy (i ,2) * (1 - T_ref / yall (i ,2) ) ; % exergy
used for vaporization
exergy (i ,2) = energy (i ,3) - T_ref * rho_g * v ( i ) * ... % exergy
inputted into system
A_orifice * log ( T ( i ) / yall (i ,2) ) ;
exergy (i ,3) = - exergy (i ,1) / exergy (i ,2) ; % exergy
efficiency
end

% Average exergy efficiency until particle reaches 0.02 moisture content


averageExEff = mean ( exergy (2: effectiveStop , 3) ) ;

% Convert to ratio of exergy lost


exergyLossRatio = 1 - averageExEff ;

68
% Energy efficiency of gas in dense phase
for i = 1:1: size ( yall ,1)
Q_gas (i ,1) = rho_g * v ( i ) * c_g * ( T ( i ) - yall (i ,2* nLayers +7) ) ...
* A_orifice ; % energy available in
dense phase
Q_gas (i ,2) = - energy (i ,2) / Q_gas (i ,1) ; % energy used for
vaporization
end

% Average Q_gas efficiency until particle reaches 0.02 moisture content


averageQgasEff = mean ( Q_gas (2: effectiveStop , 2) ) ;

% Convert to dense phase energy lost to outlet


QgasLossRatio = 1 - averageQgasEff ;

TimeStop = effectiveStop ;

RunEnergy = sum ( energy (2: effectiveStop ,3) ,1) ;

RunEnergyDay = RunEnergy * 86400 / TimeStop ; % Total Energy / day ( J )


DryMassDay = 3 * 86400 / TimeStop ; % Total powder / day ( kg )
EnergyCost = 0.0000000028; % $/ J
PowderProfit = EnergyCost * RevenueRatio ; % $/
kg

OF = ( - DryMassDay * PowderProfit + RunEnergyDay * EnergyCost ) *100;


fprintf ( '% f \ n ' , OF ) ;
fprintf ( ' ExergyLoss = % f \ n ' , exergyLossRatio ) ;
fprintf ( ' RunEnergy = % f \ n ' , RunEnergy ) ;
fprintf ( '% f % f % f % f % f % f % f % f % f \ n ' , T_in , v_in , energyLossRatio , ...
exergyLossRatio , QgasLossRatio , effectiveStop , tChange , y2 (1 ,1) , yall (
effectiveStop ,2) ) ;

A.3 Optimization

69
%{
Two - phase fluidized bed drying model for pharmaceutical powder .
Model to be used to validate experimental data .

This file contains the globalsearch optimization using the model .

Marcus Lau
mll668 @ mail . usask . ca
%}
% create empty text file
fid = fopen ( ' OptimResults ' , 'w ') ;
fclose ( fid ) ;

for i = 1:1:1
RevenueRatio = 1534000 + (i -1) * 500;

model = @( c ) fbDrying (c , RevenueRatio ) ;

x0 = [333.15 1.8 313.15 1.4 1200];


lb = [303.15 0.9 303.15 0.9 240];
ub = [343.15 2.0 343.15 2.0 3000];

% set up optimization and run problem


options = optimoptions (" fmincon " , ' MaxFunEvals ' , 3 e6 , ' MaxIter ' , 1 e6 ,
' O pt i m al i t yT o l er a n c e ' , 1e -10 , ' StepTolerance ' , 1e -10 ,...
' C on s t ra i n tT o l er a n ce ' , 5e -8 , ' Algorithm ' , ' sqp ') ;
problem = cr ea te Opt im Pr ob le m ( ' fmincon ' , ' x0 ' , x0 , ' objective ' , model ,
' lb ' , lb , ' ub ' , ub ,...
' options ' , options ) ;
gs = GlobalSearch ( ' Func tionTo leranc e ' , 1e -6 , ' XTolerance ' , 1e -6) ;
[c , min_sum , flag , ~] = run ( gs , problem ) ;

line = [i , RevenueRatio , x0 , c , min_sum , flag ];


print_line ( ' OptimResults ' , line ) ;
end

70

You might also like