Mixing Dynamics of Pulp Suspensions in Cylindrical Vessels: Kwok Wai Leo Hui, 2011

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 165

MIXING DYNAMICS OF PULP SUSPENSIONS

IN CYLINDRICAL VESSELS

by

KWOK WAI LEO HUI

B.A.Sc., The University of British Columbia, Vancouver, Canada, 1996


M.Eng., The University of British Columbia, Vancouver, Canada, 1998

A THESIS SUBMITTED IN PARTIAL FULFILMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

THE FACULTY OF GRADUATE STUDIES

(CHEMICAL AND BIOLOGICAL ENGINEERING)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

April 2011

 Kwok Wai Leo Hui, 2011


ii

Abstract

Cylindrical agitated chests are frequently used to facilitate manufacturing processes in pulp and

paper industry and one of their main functions is to attenuate any process disturbances. However,

owing to the inherited non-Newtonian nature of pulp suspensions, it is not easy to achieve

complete mixing and with the improper chest design, these agitated chests do not always perform

ideally or satisfactorily. The cavern formation in incomplete mixing may induce bypassing and

dead zones, which significantly affect the chest performance. A study of cavern formation in a

cylindrical agitated chest was thus carried out. Also, a dynamic model developed by Soltanzadeh

et al. (2009) was used to quantify the mixing dynamics of the cylindrical chest. In addition, using

computational fluid dynamics (CFD), the simulated results of the flow in the chest were

compared with the experimental results to verify the applicability of the CFD model on the study

of pulp suspension agitation.

To investigate the cavern formation in a lab-scale cylindrical chest, electrical resistance

tomography (ERT) and ultrasonic Doppler velocimetry (UDV) were applied to estimate the

cavern shape and size. Both methods gave satisfactory results and as expected, the cavern size

was found to increase with impeller speeds. The cavern shape was best described as a truncated

right-circular cylinder. Based on this observation, a model considering the interaction between

the cavern and chest walls was developed to calculate the cavern volume.

With the dynamic model, a series of dynamic tests were carried out to characterize the

mixing behavior of the lab-scale cylindrical chest. It was found that the proposed flow

configuration with the outlet close to the cavern could minimize the bypassing which affects

mixing quality. Also, ERT verified the presence of cavern and dead zone when the chest was not

completely agitated in continuous-flow operation.


iii

Numerical simulations using CFD were compared with the experimental results under

different operating conditions. Pulp suspensions are a mixture of water and wood fibres that can

entangle each other to form flocs affecting the mixing flow. Owing to this complex rheology, it

is not easy to model the agitation precisely in CFD using a homogeneous fluid model. The floc

formation and air entrapment observed in experiments were difficult to be numerically taken into

account in the simulations. Although the CFD model could not exactly predict the mixing

situation of pulp suspensions, it still can be used to estimate the mixing flow patterns, e.g., flow

directions, in the proposed chest designs.


iv

Preface

This thesis consists of 6 chapters and there are 2 co-written chapters:

 3 CAVERN MEASUREMENT IN PULP SUSPENSIONS

 4 DYNAMIC TEST STUDY ON LAB-SCALE CHEST

Chapter 3 describes the measurement of cavern formation in pulp suspension agitation.

Proposed by my supervisor, Chad Bennington, electrical resistance tomography (ERT) and

Ultrasonic Doppler velocimetry (UDV) were applied to estimate the cavern shape and size. I

selected the conductive tracer to identify the cavern and designed the measurement method to

determine its size. After carrying out the experiments, I analyzed the data, developed a cavern

model with the help of Chad Bennington and wrote the manuscript. Finally, my two supervisors,

Chad Bennington and Guy Dumont, revised and edited the manuscript. A version of this chapter

has been published as “Cavern formation in pulp suspensions using side-entering axial-flow

impellers” in Chemical Engineering Science Journal (64) on pages 509-519 in 2009.

Chapter 4 illustrates the study of mixing dynamics in a lab-scale cylindrical chest. Using a

pseudo-random binary signal (PRBS) to control the addition of a conductive tracer, I carried out

a series of tests to analyze the mixing dynamics of the chest. With the input-output data and the

dynamic model, I estimated the model parameters using MATLAB and concluded the chest

performance. Finally, I wrote the manuscript which was revised and edited by my two

supervisors. A version of this chapter will be submitted for publication as “Mixing Dynamics in

Cylindrical Pulp Stock Chests” in 2011.


v

Table of Contents

Abstract ........................................................................................................................................... ii
Preface............................................................................................................................................ iv
Table of Contents ............................................................................................................................ v
List of Tables ................................................................................................................................ vii
List of Figures .............................................................................................................................. viii
List of Symbols and Abbreviations............................................................................................... xii
Acknowledgements ..................................................................................................................... xv
1 INTRODUCTION .................................................................................................................. 1
1.1 Thesis background............................................................................................................ 1
1.2 Pulp suspension rheology ................................................................................................. 2
1.3 Agitated chest design ....................................................................................................... 6
1.4 Research objectives .......................................................................................................... 9
2 STUDY METHODS ............................................................................................................. 11
2.1 Electrical resistance tomography (ERT) ........................................................................ 11
2.2 Ultrasonic Doppler velocimetry (UDV) ......................................................................... 14
2.3 Computational fluid dynamics (CFD) ............................................................................ 17
3 CAVERN MEASUREMENT IN PULP SUSPENSIONS ................................................... 23
3.1 Introduction .................................................................................................................... 23
3.2 Experimental set-up and procedure ................................................................................ 26
3.3 Results and discussion.................................................................................................... 30
3.3.1 Selection of cavern measurement technique ........................................................... 30
3.3.2 Effect of impeller speed and pulp mass concentration on cavern volume .............. 37
3.3.3 Effect of pulp type on cavern volume ...................................................................... 38
3.3.4 Impeller offset from the rear wall of the mixing chest ............................................ 39
3.3.5 Comparison of cavern volumes determined by ERT and the cavern models .......... 40
3.3.6 Cavern model including interaction with vessel walls ........................................... 44
3.4 Summary ........................................................................................................................ 52
4 DYNAMIC TEST STUDY ON LAB-SCALE CHEST ....................................................... 53
4.1 Introduction .................................................................................................................... 53
4.2 Experimental set-up and procedures .............................................................................. 56
4.3 Results and discussion.................................................................................................... 61
4.4 Summary ........................................................................................................................ 75
5 CFD SIMULATION OF PULP MIXING ............................................................................ 76
5.1 Introduction .................................................................................................................... 76
5.2 Experimental study of pulp mixing ................................................................................ 79
5.3 CFD modeling of pulp mixing ....................................................................................... 80
5.3.1 Computational geometry ......................................................................................... 80
5.3.2 Modelling suspension rheology .............................................................................. 82
5.3.3 Condition setup for CFD model .............................................................................. 84
5.3.4 Cavern volume determination and dynamic test simulation ................................... 86
5.4 Results and discussion.................................................................................................... 87
5.5 Summary ...................................................................................................................... 108
vi

6 OVERALL CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK ... 109


6.1 Overall conclusions ...................................................................................................... 109
6.2 Recommendation for future work ................................................................................ 112
BIBLIOGRAPHY ....................................................................................................................... 114
APPENDIX A: Details of ERT components .............................................................................. 120
APPENDIX B: MATLAB program for dynamic model parameter estimation ......................... 124
APPENDIX C: Program for cavern volume determination in CFD ........................................... 136
APPENDIX D: Program for tracer analysis in CFD .................................................................. 137
APPENDIX E: ERT data ............................................................................................................ 138
vii

List of Tables
Table 3.1 Comparison of ERT/UDV cavern measurements (Cm = 2% hardwood pulp, E/D = 0.6,
C/D = 0.7 and Z/T = 1.0) ............................................................................................................... 36

Table 4.1 Time delay (Td) measured versus theoretical values for hardwood pulp Cm = 3% (N =
425, 525, 672 and 810 rpm at each flow rate) .............................................................................. 61

Table 4.2 Time constant (2) measured for complete mixing of hardwood pulp at Q = 14 L/min
(Theoretical 2 = 149s) .................................................................................................................. 62

Table 4.3 Dynamic test results ...................................................................................................... 64

Table 4.4 Calculation of impeller speed on scale-up using different criteria ............................... 73

Table 4.5 Power predicted for complete mixing of a hardwood pulp suspension in cylindrical
stock chests using different scale-up criteria (D/T = 0.43, Z/T = 0.8) .......................................... 73

Table 5.1 Rheological parameters of hardwood and softwood pulps ........................................... 84

Table 5.2 Grid independence study (N = 425rpm, 3% softwood pulp (K1 = 135, n = 0.21, y =
155Pa) ........................................................................................................................................... 91

Table 5.3 Comparison of results from CFD and dynamic tests for Cm = 2% hardwood (Q =
14L/min., Z/T = 0.8).................................................................................................................... 103

Table 5.4 Comparison of results from CFD and dynamic tests for Cm = 3% softwood (Q =
14L/min., Z/T = 0.8).................................................................................................................... 103
viii

List of Figures
Figure 1.1 A generalized stress-rate curve for a fiber suspension (Gullichen and Harkonen, 1981). 4

Figure 1.2 Cross-section of a reduced-bottom chest....................................................................... 8

Figure 2.1 Operation of electrical resistance tomography (Malmivuo and Plonsey, 1995). ........ 12

Figure 2.2 Image reconstruction grid (Industrial Tomography Systems, 2007). .......................... 13

Figure 2.3 Principle of UDV showing the measurement of velocity along the measuring line
(Takeda, 1995). ............................................................................................................................. 16

Figure 2.4 Cell types used in computational grid (Ferziger and Peric, 1999). ............................. 20

Figure 2.5 2D structured and unstructured grids (Ferziger and Peric, 1999). .............................. 20

Figure 3.1 Cross-section of cylindrical stock chest showing ERT sensor planes and range of
impeller positions used. ................................................................................................................ 28

Figure 3.2 Suspension yield stress as a function of mass concentration. ..................................... 29

Figure 3.3 Photograph of conductive tracer particles used in tests. Individual particles range in
diameter from 4 to 6 mm (typically 5 mm) and length from 15 to 20 mm (typically 17.5mm). .. 32

Figure 3.4 ERT images obtained for a Cm = 3% hardwood pulp suspension agitated at N = 650
rpm with E/D = 0.4, C/D = 0.7 and Z/T = 1.0. Images on the left were reconstructed using data
obtained with tracer particles (Fig. 3.3) and images on the right were reconstructed by adding a
conductive fluid of 10.9 mS/cm NaCl (15mL) to a pulp suspension having a background
conductivity of 0.2 mS/cm (43.4L). The impeller is located between planes 2 and 3 and
electrodes 5 and 6 as shown for reference purposes in the upper left image. ............................... 33

Figure 3.5 Comparison of ERT images reconstructed using the tracer particles (left) and saline
solution (right). Uppermost sampling plane (P5). The visual surface view is given in the centre
image with the dotted line enclosing the region of active surface motion (determined visually).
(Hardwood pulp suspension at Cm = 3%; N = 650 rpm, E/D = 0.4, C/D = 0.7, Z/T = 1.0). ......... 34

Figure 3.6 Probe locations for UDV measurements. The probe is moved around the chest to
obtain the velocity profiles needed to map the cavern boundary. The arrows show the direction of
the sonic emissions and the numbers give the ERT electrode positions....................................... 34

Figure 3.7 (a) Velocity profile measured by UDV for a representative probe location. The
locations where the velocity reaches zero indicate the cavern boundary. (b) Cavern cross-section
formed by connecting the boundary points measured from all sample locations. ........................ 35
ix

Figure 3.8 Comparison between ERT and UDV caverns made for hardwood pulp at Cm = 2% and
N = 200 rpm (E/D = 0.6, C/D = 0.7 and Z/T = 1.0). ..................................................................... 36

Figure 3.9 Cavern shape (shaded volume) determined for a Cm = 2% hardwood at various
impeller speeds using ERT data (E/D = 0.6, C/D = 0.7, Z/T = 1.0).............................................. 37

Figure 3.10 Effect of impeller speed (N) and pulp consistency (Cm) on cavern volume (Vc ) for
hardwood and softwood pulps (E/D = 0.4, C/D = 0.7 and Z/T = 0.8). ......................................... 39

Figure 3.11 Effect of impeller position (E/D) on cavern volume (Vc) with (a) rotation speed and
(b) power (Cm = 2% hardwood, C/D = 0.7, Z/T = 1.0). ................................................................ 41

Figure 3.12 Comparison of measured cavern volumes and predictions made using the equations
of Amanullah et al. (1998) and Elson (1990). .............................................................................. 42

Figure 3.13 Axial force number (Nf) (a) and power number (NP) (b) for the D = 140mm Maxflo
impeller as a function of impeller speed and suspension mass concentration. ............................. 43

Figure 3.14 Diagram showing intersecting circles with the impeller located at the centre of a
virtual cylinder. The overlapping area between the mixed zone and mixing chest is the cavern
region. ........................................................................................................................................... 46

Figure 3.15 Rotation speed (a) and wall stress calculated (b) for complete mixing in the vessel
with hardwood pulp as a function of suspension mass concentration (E/D = 0.4, C/D = 0.7). .... 47

Figure 3.16 Comparison of wall stresses measured/estimated and suspension yield stress for
hardwood pulp as a function of mass concentration. Estimates for w based on the yield stress (y
- measured), the point at which the cavern first completely filled the mixing vessel, and friction
losses in flowing pulp suspensions. .............................................................................................. 49

Figure 3.17 Cavern volume versus impeller speed for laboratory mixer. Comparison of
experimental data (measured by ERT) with model predictions for hardwood pulp. Z/T = 0.8, E/D
= 0.4. (a) Cm = 2%: y = 9 Pa ;w =12 Pa ; a = 0; Cm = 3%: y = 38 Pa ;w = 42 Pa ; a = 0; Cm =
4%: y = 69 Pa ;w =90 Pa ; a = 0. (b) (y = w = a) Cm = 2%: y = 9 Pa; Cm = 3%: y = 38 Pa; Cm =
4%: y = 69 Pa. .............................................................................................................................. 51

Figure 4.1Dynamic model for non-ideal flow in agitated pulp stock chests (2>>1). .................. 56

Figure 4.2 Cross-section (a) and photograph (b) of cylindrical stock chest. The arrows in (b)
show the direction of suspension flow in continuous operation. .................................................. 59

Figure 4.3 Schematic of apparatus used for the dynamic tests . ................................................... 60

Figure 4.4 Typical signals measured for model parameter identification. ................................... 60
x

Figure 4.5 ERT images of a dynamic test at Cm = 3% and N = 425rpm (E/D = 0.6, C/D = 0.7, Z/T
= 0.8). The suspension inlet and outlet positions are shown in the images of P4 and P1 at time t =
0 s, respectively. ............................................................................................................................ 65

Figure 4.6 Effect of impeller speed (N) and pulp consistency (Cm) on time constant (2) (E/D =
0.6, C/D = 0.7, Z/T = 0.8). ............................................................................................................ 66

Figure 4.7 Effect of flow rate (Q) on time constant (2) with rotation speed (Cm = 3%, E/D = 0.6,
C/D = 0.7, Z/T = 0.8). ................................................................................................................... 67

Figure 4.8 Effect of flow rate (Q) on mixing volume (Vmix) with rotation speed (Cm = 3%, E/D =
0.6, C/D = 0.7, Z/T = 0.8). ............................................................................................................ 67

Figure 4.9 Effect of pulp types (HW: hardwood and SW: softwood) on cavern size obtained in
dynamic and batch (ERT) operations at Cm = 3% (E/D = 0.6, C/D = 0.7, Z/T = 0.8). ................. 69

Figure 4.10 Effect of pulp types (HW: hardwood and SW: softwood) with similar yield stress
(38Pa) on mixing volume (Vmix) with rotation speed (N) (E/D = 0.6, C/D = 0.7, Z/T = 0.8). .... 69

Figure 4.11 Comparison between cavern size obtained in dynamic and batch (ERT) operations at
various hardwood pulp consistencies (Cm) (E/D = 0.6, C/D = 0.7, Z/T = 0.8, Q = 14 L/min for the
dynamic tests). .............................................................................................................................. 70

Figure 5.1 Experimental cylindrical chest and computational domain for (a) batch and (b)
continuous-flow mixing. ............................................................................................................... 82

Figure 5.2 Locations where velocity samples were determined for comparison of three different
meshes. .......................................................................................................................................... 89

Figure 5.3 Velocity profiles calculated for softwood Cm = 3% at two locations (a) and (b) in front
of the impeller (as shown in Figure 2) for three different meshing schemes at N = 425rpm. ..... 90

Figure 5.4 Comparison of cavern size obtained from ERT and numerical simulations for a Cm =
2% hardwood (Z/T = 1.0). ............................................................................................................. 91

Figure 5.5 Comparison of cavern size obtained from ERT and numerical simulations for a Cm =
3% softwood (Z/T = 0.8). .............................................................................................................. 93

Figure 5.6 Flow fields of batch mixing for a) Cm = 2% hardwood at N = 200rpm using Bingham
plastic model and b) Cm = 3% softwood at N = 475 rpm using Herschel-bulkley model. ........... 94

Figure 5.7 a) Top view and b) side view of probe locations for UDV measurements. The arrows
show the direction of the sonic emissions at three heights (12.5mm, 92.5mm and 252.5mm
measured from the chest bottom). ................................................................................................. 96
xi

Figure 5.8 Comparison of measured and computed velocity profiles at three positions for a Cm =
2% hardwood at N = 200rpm. Positive velocity means the velocity direction is away from the
probe. ............................................................................................................................................ 97

Figure 5.9 Comparison of measured and computed velocity profiles at three positions for a Cm =
2% hardwood at N = 250rpm. Positive velocity means the velocity direction is away from the
probe. ............................................................................................................................................ 98

Figure 5.10 Comparison of measured and computed cavern volumes. (a) Cm = 2% hardwood
pulp using Bingham plastic model for cavern determination (Z/T = 1.0); (b) Cm = 3% softwood
pulp using Herschel-Bulkley model for cavern determination (Z/T = 0.8). ............................... 100

Figure 5.11 Experimental and computed dynamic responses for Cm = 2% hardwood pulp. ...... 104

Figure 5.12 Experimental and computed dynamic responses for Cm = 3% softwood pulp. ....... 105

Figure 5.13 Path lines of particles (each particle is represented by a different color) simulated by
CFD using Cm = 2% hardwood pulp at N = 250rpm and Q = 14L/min...................................... 106

Figure 5.14 Experimental and computed determination of mixed volume for (a) Cm = 2%
hardwood pulp and (b) Cm = 3% softwood pulp in continuous operation. ................................. 107
xii

List of Symbols and Abbreviations

a distance between the impeller location and the vessel centre, (m)
A2 surface area of cavern, (m3)
b perpendicular distance from the intersection point of the vessel wall and the
cavern to the distance a, (m)
c speed of sound, (m/s)
C height between the chest bottom and the impeller centre, (m)
C1, C2 constant, dependent on the impeller type and impeller geometry
Cm mass concentration or consistency, (%)
D, D1, D2 impeller diameter, (m)
Dc cavern diameter, (m)
Dp pipe diameter, (m)
Dr, DR rotor diameter, (m)
DT housing diameter, (m)
E distance between the chest wall and the impeller centre, (mm)
f bypassing parameter
f1 factor to correct for temperature and pipe roughness
F total force on the cavern boundary, (N)

F external body forces
Fa axial force, (N)
Fd Doppler shift frequency, (Hz)
Fe transducer frequency, (Hz)

g gravitational acceleration, (m/s2)
H rotor height, (m)
Hc cavern height, (m)
H/L friction loss, (m water/100m pipe)
I unit tensor
k constant in the Herschel-Bulkley model
K numerical coefficient, constant for a given pulp
K1 consistency index
M, M1, M2 torque, (Nm)
Mo, Mo1, Mo2 momentum number (m4/s2)
Mo level momentum (m2/s2)
n power-law index
N, N1, N2 impeller rotation speed, (rpm)
Nf dimensionless axial force number; Fa/N2D4
NP power number; P/N3D5
p static pressure, (Pa)
P, P1, P2 power, (W)
Q pulp flow rate through the chest, (L/min.)

r position vector in the rotating frame
rc cavern radius, (m)
rv vessel radius, (m)
xiii

Rey yield stress Reynolds number; N2D2/y


Sa cavern-air surface, (m2)
Sb cavern-vessel base surface, (m2)
Sm mass added to the continuous phase from the dispersed second phase and any
user-defined sources
Sp cavern-suspension surface, (m2)
Sw cavern-wall surface, (m2)
t time, (s)
td time delay between the start of the pulse burst and its reception in UDV, (s)
T, T1, T2 chest diameter, (m or mm)
Td time delay for the dynamic model, (s)
Tm maximum torque, (Nm)
v velocity, (m/s)

v absolute velocity

vr relative velocity
V, V1, V2 suspension volume in the chest, (m3)
Vc cavern volume, (m3)
Vmix fully mixed volume in the chest, (m3)
Vt suspension volume in the chest, (m3)
x distance from the impeller to the chord connecting the intersection points
between the cavern and vessel wall (m)
X position, (m)
Z stock height, (m)

Greek symbols

 angular velocity of the rotating frame
, ,  indices, constant for a given pulp
F power dissipation per unit volume, (W/m3)
 strain rate (s-1)
y strain rate at yield stress, (s-1)
 Doppler angle, ()
1 angle for determining the cavern-suspension surface, ()
2 angle for determining the cavern-wall surface, ()
 fluid density, (kg/m3)
 dynamic viscosity, (kg/ms)
o yielding viscosity, (kg/ms)
 plastic viscosity, (kg/ms)
 shear stress, (Pa)
 stress tensor
1 time constant for bypassing, (s)
2 time constant for mixing zone, (s)
a friction force per unit area at the air suspension interface, (Pa)
th theoretical time constant of the system, (s)
xiv

w friction force per unit area at the vessel wall, (Pa)


y fluid yield stress, (Pa)

Abbreviations

CFD Computational fluid dynamics


ERT Electrical resistance tomography
PIV Particle image velocimetry
UDV Ultrasonic Doppler velocimetry
xv

Acknowledgements

I would first like to express my profound gratitude and appreciation to my deceased supervisor

Dr. Chad P.J. Bennington for his guidance, encouragement and support throughout the research.

I also would like to give special thanks to Dr. Guy A. Dumont for his innumerable assistance and

support in the completion of my thesis.

I gratefully acknowledge the valuable advice and helpful suggestions of the members of my

thesis committee, Dr. Fariborz Taghipour and Dr. James Olson.

I am grateful to Mr. Ali Soltanzadeh for his assistance in dynamic model parameter estimation

and Dr. Clara Ford for her advice in CFD modeling.

I acknowledge the assistance of all the staff in the Pulp and Paper Centre and the Chemical and

Biological Engineering Department at UBC. I am especially grateful to Peter Taylor and Tim

Patterson for their valuable help in the fabrication, installation and maintenance of the

experimental setup.

Financial support from the National Sciences and Engineering Research Council of Canada

(NSERC) is highly appreciated.


xvi

To my parents,

For their love, encouragement and support.


1

1 INTRODUCTION
1.1 Thesis background

Pulp suspension mixing in stock chests is an important operation in pulp and paper industry. It

prevents the pulp suspension from dewatering, enhances chemical contact in bleaching stages

and blends pulp streams before papermaking. Improper mixing could lead to non-uniform

delivery of pulp stock and increase process variability, resulting in unstable production and poor

product quality. Although regulatory control loops are employed to minimize the effects of these

process upsets and to maximize product uniformity, they are only effective at regulating out the

disturbances at low frequencies (Bialkowki, 1992). Thus, agitated chests are required to remove

high frequency variability, complementing the action of control loops.

However, these chests do not perform ideally because of the complex rheology of pulp

suspensions and the ineffective design of feed flow in and out of the chests. The suspension

yield stress resulted from the mechanical strength of the wood fibre networks may lead to the

formation of mixing and stagnant zones, creating undesired flows in the agitated chests. Ein-

Mozaffari et al. (2003) analyzed the output signal of an industrial pulp stock chest to a step input

change and found that the non-ideal flows, i.e., bypassing and dead zones, deteriorated the

response signal. To identify what parameters affecting the responses of the agitated pulp chests,

Ein-Mozaffari et al. (2005) studied the mixing dynamics in a rectangular pulp chest using two

different pulp suspensions: a short-fibred hardwood pulp and a long-fibred softwood pulp over a

range of fibre mass concentration and addressed the importance of inlet and outlet locations

relative to the mixing zone.

Other studies have also been carried out to investigate the flow patterns and dynamic

behavior of pulp suspensions in agitated stock chests and most of these studies were conducted
2

in rectangular chests (Bakker and Fasano, 1993; Ein-Mozaffari et al., 2003; Ford, 2004). Very

little research (Wilstrom and Rasmuson, 1998a) about pulp agitation has been carried out in

cylindrical chests, which are widely used, particularly as controlled mixing zone in bleaching

towers and as dilution zone for high density pulp storage chests. Therefore, the purpose of this

study is to investigate the mixing performance of cylindrical chests using electrical resistance

tomography (ERT), ultrasonic Doppler velocimetry (UDV) and dynamic tracer tests. In

addition, computational fluid dynamics (CFD) was used to simulate the mixing behavior of pulp

suspensions in cylindrical chests.

To achieve an effective mixing of a fluid, an understanding of the fluid properties,

especially rheology, is important and essential. Rheology governs fluid motion during agitation.

Pulp suspensions, as a non-Newtonian fluid, have caverns (the turbulent mixed regions) formed

around impellers when they are not in complete mixing. The cavern size would significantly

affect the mixing efficiency of the stock chests which are designed, based on past experience and

semi-empirical techniques, with their shape and size to facilitate the agitation of pulp suspension.

The details of cavern formation are discussed in Chapter 3.

1.2 Pulp suspension rheology

Fluid rheology plays an important role in determining the fluid behavior in stirred vessels. A

pulp suspension is a heterogeneous mixture of water and wood fibres. The fibres interact and

entangle with each other to form a viscoelastic network consisting of fibre aggregates (flocs)

(Gullichsen and Harkonen, 1981). The longer the fibers, the easier the floc formation. This fibre

network can be described with two levels: MACRO (inter-network) and MICRO (intra-network)

scales (Wikstrom and Rasmuson, 1998b). The MACRO scale represents the network between the
3

fibre flocs and the MICRO is described as a 3-D fibre network within a fibre flock. This complex

structure determines the rheological behavior of the pulp suspension, which initially acts as a

non-Newtonian fluid having a significant yield stress (y) (Figure 1.1). When the applied shear

exceeds the suspension yield stress, the fibre network is disrupted and fluid-like motion can be

created within the suspension. At low shear stresses, the MACRO structure of the fibre network

is first broken and the flow of pulp suspension is dominated by the break-up between the fibre

flocs. The fibre flocs flow as fibre flock spheres. When the shear becomes higher and exceeds a

certain value (d) at the fluidization point, the MICRO network is ruptured into individual fibres,

i.e., fluidization and the suspension behaves nearly as a turbulent Newtonian fluid.

The simplest rheological model for a fluid possessing a yield stress is the Bingham model,

which is a linear equation for the shear rate (Silvester, 1985; Barnes et. al, 1989; Chhabra and

Richardson, 1999):

(1.1)

where  is the shear stress,y is the yield stress,  is the shear rate and  is the plastic viscosity or

coefficient of rigidity.

However, based on the generalized stress-rate curve of the pulp suspension in Figure 1.1,

at the low shear rates where the shear stress is between y and d (the flow is laminar), shear-

thinning behavior is observed after the yield stress point. The Herschel-Bulkley model

(Macosko, 1993; Chhabra and Richardson, 1999), which contains a yield stress and a shear-

thinning parameter, can be better suited for describing the non-Newtonian nature of pulp

suspensions.
4

Laminar Turbulent

Shear stress, 
Bingham model

d
y
Newtonian fluid

Shear rate, 

Figure 1.1 A generalized stress-rate curve for a fiber suspension (Gullichen and Harkonen, 1981).

   y  k  n (1.2)

The shear-thinning behavior is obtained when the power index, n, is set to less than 1 and k is a

constant in the model. For pulp suspensions, Ford (2004) estimated k and n to be 0.001 and 0.25

respectively.

To create motion or turbulence in the pulp suspension, the yield stress must be exceeded

and the required stress has been correlated with the suspension mass concentration by the

following equation (Bennington et al. 1990):

 y  Cm  (1.3)

where y is the yield stress, Cm is the suspension mass concentration in percent,  and  are fitted

constants. Ranges of  and  were reported to be 1.18 - 24.5 and 1.25 - 3.02 respectively. For

the mass concentration between 0.6% and 30%, the yield stress will vary from 0.3 Pa to 3104Pa.
5

Pulp suspensions in a turbulent state are often referred as “fluidized”, which implies fibre

motion leading to energy dissipation (Bennington and Kerekes, 1996). Fluidization can be

quantified through power and energy expenditure, and Wahren (1980) applied this concept to

estimate the power dissipation per unit volume of pulp suspension for the onset of fluidization,

F, using the yield stress of the fibre network:

 y2
F   1.2  10 4 C m
5.3
(1.4)

where F is in W/m3 and  is the viscosity of water. However, the use of the water viscosity to

estimate the power required to fluidize a pulp suspension is inappropriate because the apparent

viscosity of the suspension is larger than that of water (Bennington and Kerekes, 1996). Thus

using the criterion chosen by Gullichsen and Harkonen (1981) for the onset of fluidization,

Bennington and Kerekes (1996) developed an expression to estimate the energy dissipation

necessary for fluidization:

 2 .3
 DT 
 F  4.5  10 Cm
4 2.5
  ;
 DR 
(1.5)
D
.  T  3110
13 . ; .  Cm  12.6
DR

where DT is the diameter of the housing and DR is the diameter of the rotor. When the gap size is

extrapolated to zero, the equation becomes:

 F  4.5  10 4 Cm 2.5 (1.6)

Owing to the complexity of the rheological behavior of pulp suspension, it is not easy to

agitate the suspension thoroughly. The shear stress provided by the agitators must exceed the
6

yield stress everywhere in the suspension to create motion through the whole chest. In actual

mixing situations, however, the applied shear stress may not be strong enough to create thorough

mixing in the suspension, leading to the formation of mixing regions surrounded by stagnation

zones.

1.3 Agitated chest design

To facilitate the bulk motion and the cavern generated by the impeller, the chest shape and size

are important in achieving thorough agitation in the suspension. Generally, the agitated pulp

stock chests are designed based on experience. One of the methods for designing these chests

was developed by Yackel (1990). Derived from the concept of “conservation of momentum”,

the method is based on the impeller momentum number and the level momentum. Each

particular impeller has a momentum number (Mo) which is defined as:

Mo  C1 N 2 D 4 (1.7)

where N is impeller rotation speed, D is impeller diameter and C1 is a constant that depends on

the suspension rheology, impeller type and impeller geometry. By referring to sets of diagrams

and tables relating chest diameter or width, consistency, stock level and retention time to obtain a

process number which is converted to the corresponding momentum number, the required

impeller size and horsepower for agitation could be determined. However, there is no direct

reference to suspension rheology in Yackel‟s method. The non-ideal flows like bypassing and

dead zone are not considered in the chest design and the degree of motion generated in the

suspension is not given.


7

In terms of chest geometry, rectangular chests are widely used in pulp mixing because

they save space when they are grouped together using common walls. The optimum shape for a

rectangular chest to achieve complete motion in a pulp suspension is a cube (Yackel, 1990).

Owing to space limitation or restriction, all chests cannot be perfect cubes and so a general rule

of thumb for rectangular chests is that the length to width ratio should not be over 1.5. Besides

rectangular ones, cylindrical chests are used in pulp and paper industry. Since cylindrical chests

take advantage of hoop stress design (tension), they generally have thinner walls than rectangular

chests of the same height (Yackel, 1990; Reed, 1995). Thus, it could be economical to have a

cylindrical chest rather than a rectangular one. To achieve complete motion of a pulp suspension

with the minimum power in a cylindrical chest, the ideal ratio of stock depth (Z) to the chest

diameter (T) should be 0.8 (Z/T = 0.8) (Yackel, 1990).

The reduced-bottom chest is another chest design that is employed to save energy (Figure

1.2). It is like a high-density storage chest placed on top of a dilution chest where controlled

zone agitation is maintained. The top of the chest operates in the range of 5% to 12% stock

consistency and the stock flows slowly down into the bottom of the chest where dilution water is

added to reduce the consistency to 3 - 4%. The lower section is a cylindrical chest with a height

equal to one-half of the diameter (Yackel, 1990). The diameter of the upper section where plug

flow occurs to prevent any stagnant zones is usually designed to be 1.6-1.8 times the diameter of

the lower section. The overall height is three or more times the lower-section diameter.

Depending on the diameter of the lower section, a fillet or a solid horseshoe structure made of

metal or concrete may be used to reduce the volume of the dilution chest. It allows a smaller

mixing zone for decreasing the agitator horsepower without the build-up of unwanted stagnant

pulp zones (dead zones). A general guideline for maximum fillet size is one-third chest diameter
8

but owing to material cost, normally fillets are 25% or less of chest diameter (Devries and Doyle,

1995).

Z
T2

Chest specifications:
- T2 = T1(1.6-1.8);
- Zm = 0.5 T1; Zm Mixing zone
- Z/T1 = or  3
T1

Figure 1.2 Cross-section of a reduced-bottom chest.

Even though these design rules are followed, channeling (bypassing) of pulp through the

reduced-bottom chest may still occur because no consideration for cavern formation in the

controlled mixing zone is included in the design. The flows induced by the cavern may create

channeling in the upper chest portion where the high-density pulp suspension flows downward.

Also, if the cavern is not large enough, the resultant non-mixing regions will create some

shortcuts for the downward pulp stock to flow directly to the exit, without entering the mixing

zone. This channeling can reduce the residence time and create operation problems like poor

consistency control and increased chemical usage.


9

1.4 Research objectives

To analyze the mixing dynamics in cylindrical vessels, the investigation was divided into two

parts: experimental study and numerical simulation. Chapter 2 first describes two methods used

in experimentation: electrical resistance tomography (ERT) and ultrasonic Doppler velocimetry

(UDV), and the numerical simulation method - computational fluid dynamics (CFD). Then the

following two chapters show the experimental results of a lab-scale cylindrical chest. In chapter

3, the mixing behavior of the chest in batch operation was characterized using ERT and UDV,

and the experimental results of the chest operated in continuous mode are given in chapter 4.

After the experimental study, a CFD model was developed to simulate the mixing situations for

verifying its ability to predict the experimental data and the evaluation is described in chapter 5.

Finally, the overall conclusions of the research and recommendations for future work are given

in chapter 6.

The objectives of this research are summarized as:

I. To characterize the batch mixing of pulp suspensions in a cylindrical chest with a side

entering agitator by studying the cavern formation and the flow fields with the aid of

electrical resistance tomography (ERT) and ultrasonic Doppler velocimetry (UDV).

II. To carry out dynamic tests under scaled industrial operating conditions for identifying and

determining the response of an experimental chest. The data will be used to estimate the

channeling parameter and the time constant (for the mixed zone) in the dynamic model of

chest dynamics.

III. To simulate the mixing situation by means of computational fluid dynamics (CFD) and

compare the results with the experimental data.


10

The achievement of the objectives can enhance the understanding of mixing dynamics of

pulp suspensions in cylindrical vessels under batch and continuous operations. The study can

provide information about the properties of cavern (mixing zone) in batch mixing. The cavern

shape and size directly affect the mixing capacity of agitated pulp stock chests. In continuous

operation, the dynamic tests can estimate the chest performance in terms of some quantitative

parameters. These parameters can provide some insights whether the existing chest is operated

at its full capability. Finally, the numerically simulated results can verify the predictability of the

CFD model for pulp suspension agitation, evaluating its potential as a tool for chest design.
11

2 STUDY METHODS

2.1 Electrical resistance tomography (ERT)

With its mathematical concept introduced in the nineteenth century, tomography is a technique

for reconstructing the internal distribution of 2D and 3D objects from multiple external

viewpoints to provide cross-sectional slice images through the object (York, 2001; Hoyle et al.,

2005). Among different types of process tomography, electrical resistance tomography (ERT) is

increasingly used by many researchers to study industrial processes because of its advantages

including non-intrusive flow visualization, high speed measurement and no radiation hazard

(Mann et al., 1997; Ma et al., 2001; Kim et al., 2006). It is used to determine the resistance

distribution in the domain of interest by obtaining a set of measurements using sensors that are

distributed around the periphery without affecting the flow or movement of materials. The main

continuous phase in the domain must be slightly conductive and the other phases (components)

have different values of conductivity. The resistance distribution in a cross region is obtained by

injecting electrical currents on the domain and measuring voltages on it through a number of

spaced electrodes which are mounted on its periphery. The current is injected by using a pair of

neighboring electrodes and the voltage differences are measured by using all other pairs of

neighboring electrodes (Figure 2.1a). This process is then repeated by injecting current using

different pairs of neighboring electrodes and the spatial gradients of electrical conductivity are

measured (Figure 2.1b). Since ERT can detect local changes in resistivity/conductivity, it can be

used to study the mixing dynamics of a system if the fluids to be blended have different

conductivities. An injection of a fluid tracer with high conductivity can help to examine the flow
12

patterns of those non-conductive mixing components. Also, the use of multi-plane electrical

sensors in ERT can provide a pseudo three-dimensional description of the mixing process.

a) b)

Figure 2.1 Operation of electrical resistance tomography (Malmivuo and Plonsey, 1995).

An electrical tomography system is composed of a sensor system, a data acquisition

system (DAS) and an image reconstruction system (Dickin and Wang, 1996). For a cylindrical

chest, the sensor system is usually a set of 16 electrodes arranged in rings, across which an

electrical current is applied and the resulting voltage is measured. The DAS, mainly consisting

of a signal source, voltmeters and a control circuit unit, coordinates the current injection and the

voltage measurement to obtain the required data for image reconstruction. The image

reconstruction system, which is simply an algorithm, is used to determine the distribution of

regions of different conductivities within the mixing vessel by processing the collected data and

displaying the resultant distribution image. Further information about the three ERT components

is given in Appendix A. Although the ERT process is fast for obtaining results, the image

resolution is limited because only 316 pixels (the default square grid has 20  20 = 400 pixels)

are available for representing the cylindrical chest interior cross section (Figure 2.2).
13

Figure 2.2 Image reconstruction grid (Industrial Tomography Systems, 2007).

ERT has been be used as a measurement technique in many applications like level

monitoring in a horizontal pipe (Ma et al., 2001), flow visualization (Bolton et al., 2004; Wang,

2005) and jet mixing within pipelines (Stephenson, 2007). Ma et al. (2001) applied ERT to

monitor air/water two-phase flow in horizontal pipes and developed a Liquid Level Detection

method to check the water surface level. In the analysis of jet mixing within pipelines,

Stephenson et al. (2007) demonstrated the capability of ERT to visualize tracer plume

development exiting a coaxial or side entry jet in mains water and to allow a measure of mixing

progress under different operating conditions. ERT is also a popular analysis method in mixing

processes in agitated vessels. Owing to the slurry conductivity, William et al. (1996) applied

ERT for the three dimensional imaging of the concentration of solids in a slurry mixer as a

function of key process variables (particle size, impeller type and agitation speed) and for

information collection to formulate the improved mixing models and mixer design. Holden et al.

(1998) demonstrated the ability of the ERT system to distinguish the different flow patterns from

two types of impellers and to examine the mixing processes by monitoring the dispersion of a
14

brine pulse tracer. Since the resistivity would be varied by the void spaces due to bubbles in the

fluid, Wang et al. (2000) detected and constructed the pseudo-stationary pattern of gas-liquid

mixing in 3-D using an 8-plane 16-ring elements ERT system. Hosseini et al. (2010) correlated

the ERT measurements in solid-liquid mixing to solid concentration profiles to quantify the

degree of homogeneity and studied the effects of impeller and particle characteristics on the

mixing quality. Similarly, by introducing a saline solution into the region near the impeller, ERT

can be applied to study the cavern size formed in pulp suspension mixing.

2.2 Ultrasonic Doppler velocimetry (UDV)

Ultrasonic Doppler velocimetry (UDV) is a real-time and non-intrusive fluid flow measurement

method to obtain a spatial distribution of the velocity field (Takeda, 1995; Shekarriz and Sheen,

1998). Initially developed for medical applications, it has been increasingly used as a tool to

study the physics and engineering of fluid flow (Takeda, 1999). Besides providing

spatiotemporal information, UDV could be an efficient flow mapping process for comparison

with numerical simulation. Unlike laser Doppler anemometry, it is also applicable to opaque

fluids like pulp suspensions because it is unaffected by the optical properties of the fluid. UDV

is based on pulsed ultrasound echography and relies on the principle of the Doppler frequency

shift of moving particles or scatters within the flowing fluid. An ultrasonic pulse (0.45 -

10.5MHz) is emitted from the transducer along the measuring line and the same transducer

receives the echo reflected form the surface of the particles suspended in the liquid (Figure 2.3).

The velocity field information is contained in the echo. Information on the position, X, from

which the ultrasound is reflected is extracted from the time delay, td, between the start of the

pulse burst and its reception as


15

ct d
X  (2.1)
2

where c is the velocity of sound in the medium. At the same time, the flow velocity (v) is

determined from the Doppler shift frequency at that instant:

Fd c
v (2.2)
2 Fe cos 

where Fd and Fe are Doppler shift frequency and basic frequency of ultrasound respectively.

Thus a velocity profile can be obtained by analyzing the echo signal to derive instantaneous

frequencies at each instant. The measurement accuracy of UDV stated by Takeda (1991) is good:

5% for velocity and 1% for position. In the case of velocity measurement in pulp suspensions,

the sonic velocity in the suspension would be affected by suspension temperature and fibre

concentration. At room temperature (20C), the sound velocity in water is 1482 m/s. When there

are fibres in the suspension, the sound velocity would increase, but the change would not be

significant (Xu, 2003). In a recent study of measuring pulp fiber suspension flow in a rectangular

channel, Xu used a commercial UDV instrument (DOP 1000, Signal Processing, Switzerland)

and determined the error of the measured velocity to be about 2.4%.


16

Figure 2.3 Principle of UDV showing the measurement of velocity along the measuring line
(Takeda, 1995).

Besides the normal instantaneous velocity measurement, UDV has been successfully

applied in some special flow configurations like oscillating pipe flow and T-branching flow of

mercury (Takeda, 1995). Also, in a stirred tank, Bouillard et al. (2001) showed that ultrasonic

techniques can be used to identify fluid recirculation patterns, cavern regions and velocity

fluctuations (turbulence) although their reliability are hampered by the presence of bubbles in the

liquid phase. In addition, Wang et al. (2004) studied how the temperature affected UDV

measurements and concluded that the temperature effect may not be a significant issue for

measurements in aqueous solution. However, corrections are recommended for measurements in

oil flows when non-isothermal condition exists.

With regard to pulp suspensions, a number of studies have been carried out in pipe flows

and stirred tanks. Dietemann and Rueff (2004) analyzed the flow of a pulp suspension at three

consistencies up to 2.1 % in a transparent 80-mm duct using UDV and different flow regimes,

i.e., plug flow, mixed flow and turbulent flow, were clearly observed. Xu and Aidun (2005)
17

employed the pulsed ultrasonic Doppler velocimetry to measure the velocity profiles of fibre

suspension flow in a rectangular channel and a correlation was derived for the velocity profile of

fibre suspension in turbulent flow. In addition, Ein-Mozaffari et al. (2006) identified the flow

characteristics, i.e., flow number and pumping rate, of a Maxflo impeller by using the velocity

profiles measured by UDV. Since the flow velocity inside the mixing cavern is obviously higher

than that in the surrounding regions, UDV can be applied to detect the location and the

dimension of the cavern by measuring the velocity profile across the mixing zone. This method

could provide an accurate and well-defined picture of the cavern but it requires more time to get

the result since a lot of measurements have to be made.

2.3 Computational fluid dynamics (CFD)

In engineering analysis or design, many flows of practical interest cannot be easily described or

solved analytically. The approximate solutions of these flows and related phenomena can only be

obtained numerically. Computational fluid dynamics (CFD) is the numerical simulation of fluid

motion by using computers. It estimates and predicts the flow field by solving a set of

conservation equations for fluid flow, turbulence, chemical species transport and heat transfer.

These equations form a system of coupled non-linear partial differential equations (PDEs) which

describe the continuous movement of the fluid in space and time. The two important governing

equations for fluid flow are the conservation of mass and momentum. For mass, which is neither

created nor destroyed in the flows of engineering interest, the conservation equation is (Ferziger

and Peric, 1999; Bird et al., 2002):

 
   v   0 (2.3)
t
18


where  for fluid density, t for time and v for velocity vector. In the case of momentum

conservation, besides the transport by convection (the left-hand side of the Eqn. (2.4)), several

momentum sources like the pressure gradient, diffusion and other forces are also involved. If

gravity is the only body force with no other forces, the equation can be written as (Ferziger and

Peric, 1999; Bird et al., 2002):

 
t

v     vv   p      g (2.4)


where p is the static pressure,  is the stress tensor, and g is the gravitational body force. The

first term is the rate of change of momentum, the second is the convection of momentum and the

fourth is the diffusion of momentum. The stress tensor  is described by

   v  v T     v I 
   2  
(2.5)
 3 

where  is the dynamic viscosity, I is the unit tensor and the second term on the right hand side is

the effect of volume dilation.

To solve these equations numerically, all aspects of the process must be discretized, i.e., a

change from a continuous to a discontinuous formulation. The domain where the fluid flows

needs to be described by a series of connected control volumes, or computational cells and the

governing equations need to be written in an algebraic form. The process to decompose the

domain into a set of discrete sub-domains, or computational cells is called grid generation. In

two-dimensional domains, the cells are triangular or quadrilateral. In three-dimensional domains,

the elements can be tetrahedra, wedge-shaped prisms, pyramids or hexahedra (Figure 2.4). Also,

the way of dividing the domain into a finite number of cells can be classified into two main

groups: structured and unstructured (Figure 2.5). In structured grids, cells of a single family

occupy the whole domain and the positions of grid points are uniquely identified in a way
19

logically equivalent to a Cartesian grid. This simple and regular cell connectivity creates a

standard matrix of the algebraic equation system, which can be easily solved by a number of

methods. The unstructured grid, on the other hand, can consist of cells of any shape, without any

restriction on the number of neighbor elements or nodes. It is more flexible than the structured

grid and applicable to very complex geometries. However, the node locations and neighbor

connections need to be specified explicitly, so the matrix of the algebraic equation system is

irregular and it takes more time to solve it than that for the structured grid. Generally, the density

of cells in a computational grid needs to be fine enough to capture the flow details, but not so

fine because large numbers of cells require more time to solve.

For discretization of the equations, the well known processes used in CFD are the finite

difference (FD) method and the finite volume (FV) method. In the finite difference method,

derivatives in the PDEs are written in finite differences evaluated at the grid nodes using

Taylor‟s series expansion or polynomial fitting (Ferziger and Peric, 1999). The FD method can

be applied to any grid type but generally it is mostly used in structured grids. It is also simple,

effective and very easy to obtain higher-order schemes on regular grids. In the finite volume

approach, the integral form of the conservation equations is applied to the control volumes (CVs)

defined by cells (Ferziger and Peric, 1999). The variables are approximated at the centroid of

each CV. This method is suitable for complex geometries and conservative by construction

provided that surface integrals are the same for the CVs sharing the boundary.
20

Triangle Tetrahedron Pyramid

Quadrilateral Wedge Hexahedron

Figure 2.4 Cell types used in computational grid (Ferziger and Peric, 1999).

Figure 2.5 2D structured and unstructured grids (Ferziger and Peric, 1999).

The finite set of coupled algebraic equations resulting from the discretization process

should be solved simultaneously in every cell in the solution domain. Owing to the non-linearity

of the equations that govern the fluid flow and related processes, an iterative solution procedure

is required and two methods are commonly used. A coupled solution approach is the one where

all variables, or at a minimum, momentum and continuity, are solved simultaneously in a single

cell before the solver moves to the next cell, where the process is repeated. This approach is

excellent for equations to be linear and tightly coupled. On the other hand, when the equations

are complex and non-linear, a segregated solution approach is preferred. In this method, one
21

variable at a time is solved throughout the entire domain. Each equation is solved for its

dominant variable, temporarily treating the other variables as known, using the best available

values for them. This process is iterated through the equations until all equations are satisfied.

The segregated solution approach is popular for incompressible flows with complex physics,

typical of those found in mixing applications.

CFD is a useful tool for studying the flow dynamics in fluid mixing. Many literatures

show that CFD can predict velocity distribution, evaluate industrial stirred equipments like

fermenters and solid suspension vessels, and optimize impeller design (Armenante et al., 1997;

Bhattacharya and Kresta, 2002; Paul et al., 2004; Li et al., 2005; Kasat et al., 2008; Ankamma

Rao and Sivashanmugan, 2010). Ein-Mozaffari and Upreti (2009) applied CFD to study the

performance of three impellers in mixing of pseudoplastic fluids (xanthum gum solutions) which

were treated as a Herschel-Bulkley fluid. The computed velocities were found to agree

reasonably well with the measured results. The validated model was then used to simulate the

mixing time and the results showed the fluid yield stress and the clearance of the impeller had

considerable effects on the mixing time. Using an experimentally validated CFD model, Bakker

et al. (2010) numerically determined cavern shapes in a pilot-scale flotation cell for a range of

mineral slurries. The combined Herschel-Bulkley and Bingham plastic rheology models were

used to describe the slurries. A model was developed to estimate cavern height, which was found

to be inversely proportional to the slurry yield stress, in the flotation cell and this would be useful

for an engineering approximation of the cell size in the preliminary design. Gomez et al. (2010)

investigated the flow field in a rectangular vessel filled with glycerin solution and equipped with

a side-entering agitator, using CFD and particle image velocimetry (PIV). The CFD predicted
22

velocities agreed very well with the PIV measurements and this confirmed CFD as a useful tool

for optimization and design of mixing systems.

CFD has also been used in the study of pulp mixing. With the rheological data from

Gullichsen and Harkonen (1981) for the pulp suspension, Bakker and Fasano (1993) applied

CFD to model the flow in a rectangular pulp chest with a side-entering impeller and the

simulated results were in satisfactory agreement with the flow field visualizations. Wikstrom and

Rasmuson (1998a) assumed the pulp suspension as a Bingham fluid and studied the effect of

fibre suspension rheology on flow behavior in a laboratory tank equipped with a jet nozzle

agitator. The comparison of experimental results with theoretical CFD-calculations showed that

the applied Bingham model could not fully describe the rheology of the pulp suspension since

the calculated flow field deviated increasingly from the measurements as the distance from the

impeller increased and so a new rheology model must be developed to predict the entire flow

field correctly. In the CFD modeling of a rectangular pulp stock chest with a side-entry agitator,

Ford (2004) described the pulp rheology using a modified Bingham plastic model and showed

that the velocity vectors obtained from the CFD simulations agreed qualitatively with the flow

patterns observed in the experiment. The non-desirable flows like bypassing and recirculation

within the chest could be described by the predicted flow fields. Using a Bingham approximation

for the suspension rheology, Bhole et al. (2009) used CFD to model impeller flow and the

computed power and axial thrust numbers were found to be very close to the experimental

measurements. Also, treating pulp suspension as a modified Herschel-bulkley fluid,

Bhattacharya et al. (2010) studied two industrial pulp chests by simulating the process conditions

to calculate the steady-state flows using CFD. The simulated flows were compared with the

experimental results and found to agree reasonably well.


23

3 CAVERN MEASUREMENT IN PULP SUSPENSIONS1

3.1 Introduction

Agitated pulp stock chests play an important role in pulping and papermaking operations. They

are used to prevent pulp suspensions from dewatering, control pulp consistency (mass

concentration) prior to other processing steps, and for blending pulp streams ahead of

papermaking operations. Pulp suspensions display non-Newtonian rheology, including a yield

stress, which under certain conditions allows creation of caverns (regions of active mixing)

around impellers. Ein-Mozaffari et al. (2003) showed that industrial pulp chests were not ideally

mixed and developed a dynamic model which included the possibility of stock channeling to

quantify the mixing quality attained. Channeling was exacerbated by the size and location of the

cavern relative to the stock exit (Ein-Mozaffari et al., 2005) and consequently to fully understand

pulp mixing, the cavern location, size and shape should be characterized and correlated with the

operating conditions of the chest.

A number of studies have examined cavern formation in a range of non-Newtonian fluids

as a function of fluid properties and mixing conditions (Wichterle and Wein, 1981; Solomon et

al., 1981; Silvester, 1985; Elson, 1990; Amanullah et al., 1998; Wilkens et al., 2005). Solomon et

al. (1981) showed the presence of caverns in both shear thinning and yield stress fluids (CMC,

Carbopol and Xanthan gum) agitated with top-entering radial flow impellers in conventionally

stirred baffled tanks. The caverns formed were spherical and increased in size as impeller speed

was increased. Elson (1990) examined the effect of impeller type and speed on cavern

development using 1% Xanthan gum solutions in water in a similar mixing configuration. Prior

1
A version of this chapter has been published. Leo K. Hui, C.P.J. Bennington and G.A. Dumont (2009) Cavern
formation in pulp suspensions using side-entering axial-flow impellers, Chem. Eng. Sci., 64, 509-519.
24

to the cavern reaching the vessel wall, cavern volume per unit power input was greatest for the

axial flow impeller studied when compared with a Rushton turbine (radial flow impeller).

However, once the cavern interacted with the baffles vertical expansion of the cavern was only

slightly affected by impeller type with the height of the cavern increasing with the impeller

rotational speed raised to a power of about 0.7.

Several mathematical models have been developed to predict cavern size as a function of

mixing conditions and fluid properties. In general, these models specify a shape for the cavern

(based on observation) and then balance the shear force transported to the cavern surface with

the yield stress of the fluid. In all cases the models were developed for isolated caverns that do

not interact with the vessel walls. Soloman et al. (1981) assumed a spherical cavern with the

torque induced by the impeller acting tangentially at the cavern boundary. This gives an equation

for the cavern diameter, Dc

 Dc   4   N D 
3 2 2
    3 NP  (3.1)
 D       y 

where , N, D and y are the fluid density, impeller rotation speed, impeller diameter and fluid

yield stress, respectively. The impeller power number, NP = P/(N3D5), must be known or

measured and the yield stress Reynolds number, Rey = N2D2/y, calculated.

Elson et al. (1986) used an x-ray technique to image caverns produced in Xanthan gum

solutions with Rushton impellers. The cavern was described by a right circular cylinder of height

Hc and diameter Dc with the ratio of Hc/Dc set between 0.35 and 0.45 for the Ruston impeller.

 Dc     N 2 D 2 
3
1
    NP   (3.2)
 D   H c Dc  1 3    y 
2
25

Amanullah et al. (1998) considered the total force transported to the cavern boundary by

an axial flow impeller as being the sum of the tangential and axial force components. The model

was applied to a spherical cavern, giving

1  N 2 D 2 
2 2
 Dc   4N P 
   N 2
   (3.3)
   y   3 
f
 D

where Nf =Fa/(N2D4) is the dimensionless axial force number that must be measured or

determined from correlations. Amanullah et al. extended the model to shear thinning power-law

fluids, and located the cavern boundary as the surface having a specified (and low) velocity.

The model by Elson et al. (1986) predicted the cavern (right-circular cylindrical cavern)

diameter within 9% for disc turbines. The model by Amanullah et al. (1998) predicted the

cavern (spherical) diameter for an axial flow impeller with Re > 20 to within  5 – 15%. As

volume scales with the cube of diameter, the error expressed on a volume basis ranged from 16

to 53%.

Application of the above equations to pulp mixing operations is anticipated to be difficult.

First, in most industrial pulp mixer configurations the impellers are side-entering. This will

impose interaction between the cavern and the vessel floor and side walls, which has not been

considered in past work. The yield stress of pulp suspensions can be measured and characterized,

but the preferred method of Amanullah (which requires fitting the rheological curve to a power

law relationship) would be difficult as the flow curve of a fibre suspension beyond the yield

point is not well characterized or readily measured.

In this chapter, cavern size and shape was measured as a function of mixer operating

conditions (impeller speed, impeller offset from the wall, suspension height in the chest) for

several pulp fibre suspensions (hardwood and softwood pulp with suspension mass
26

concentrations from 1 to 5%). A scaled version of a commercial axial flow impeller designed for

pulp suspension agitation was used in the standard side-entering configuration in a cylindrical

chest. Due to the opaque nature of the suspensions it was impossible to use direct optical

techniques to measure cavern size. Two methods, electrical resistance tomography (ERT) and

ultrasonic Doppler velocimetry (UDV) were evaluated to obtain the needed data. Measured

cavern sizes and volumes were then compared against the cavern models available in the

literature, and a new model that accounts for interaction with the vessel boundaries was

developed.

3.2 Experimental set-up and procedure

A transparent 1/10 scale-model of a cylindrical stock chest (T = 38.1 cm) was built for studying

mixing in pulp suspensions. The chest was equipped with a side-entering Maxflo impeller

(Chemineer Inc., Dayton, OH) having D = 16.5 cm mounted 12.0 cm (C) from the bottom (i.e.,

C/D = 0.7. C/D is typically in the range of 0.5 to 1.5 (AGIMIX International, Uddevalla,

Sweden)). The impeller assembly could be moved horizontally to study the effect of impeller

position on cavern formation and two offsets, E = 7.0 and 9.6 cm (giving E/D = 0.4 and 0.6),

were used. The laboratory chest was fitted with 6 equally-spaced sensor planes for electrical

resistance tomography (ERT) with each plane having 16 square (2.52.5 cm2) stainless-steel

electrodes flush-mounted around the vessel periphery at 22.5o intervals (Fig. 3.1). Three stock

height to chest diameter ratios (Z/T = 0.8, 1.0 and 1.2) were examined (based on

recommendations by Yackel, 1990).

Equations 3.1 through 3.3 demonstrate the predicted dependences of cavern diameter on

yield stress, which for a pulp suspension depends primarily on fibre (pulp) type and suspension
27

mass concentration. Two bleached kraft pulps (hardwood and softwood) from Domtar Inc.

(Windsor, ON) were used in the study. The hardwood pulp had a length-weighted mean fibre

length of 1.28 mm and the softwood pulp had a length-weighted mean fibre length of 2.96 mm.

For the hardwood pulp, detailed cavern development was studied for three different suspension

mass concentrations (Cm = 2, 3 and 4%) while for the softwood pulp, only Cm = 3% was used.

The point at which the cavern completely filled the vessel was determined over a wider

concentration range for both pulps. The suspension yield stress was measured using a Haake

RV12 Rotovisco concentric cylinder viscometer (Thermo Fisher Scientific, Waltham, US). A

four-bladed vaned rotor (with a diameter of 19 mm and a height of 38 mm) was immersed in the

suspension and the rotor accelerated slowly from rest to a maximum speed of 0.2 min-1. The peak

value of torque was measured and used to calculate the yield stress

Tm
y  (3.4)
 3 H 1 
Dr   
2  Dr 3 

where Tm, Dr, H and y are maximum torque, rotor diameter and height, and suspension yield

stress, respectively (Macosko, 1994). This calculation assumes that the yielding surface is

defined by the outer surface of the rotor blades. Figure 3.2 summarizes the yield stress as a

function of suspension mass concentration for the pulps studied in this work. For the hardwood

pulp suspension, the yield stress varied to the power 3.0  0.1 of suspension mass concentration;

for the softwood suspension the dependence was to the power 2.0  0.1. These dependences are

similar to those reported in the literature (Bennington et al., 1990).

Two methods were evaluated for measuring the cavern size in our tests: electrical

resistance tomography (ERT) and ultrasonic Doppler velocimetry (UDV). ERT is a non-invasive

technique that can reconstruct the conductivity distribution within a region of interest from
28

electrical measurements made through a series of electrodes placed uniformly around the

periphery of a vessel. Electrical current is injected through a pair of neighboring electrodes with

voltage measurements taken between all remaining pairs of neighboring electrodes. The process

is repeated by rotating current injection around all the electrodes. The aggregate data is then used

to reconstruct the original conductivity distribution within the vessel (Mann et al., 1997; York,

2000). A P2000 ERT tomography instrument (ITS, Manchester, UK) was used to obtain the data

with the image reconstructed using the linear back projection algorithm. A suitable conductive

tracer was added to the cavern, with ERT used to locate the region of higher conductivity and

thus the cavern size and shape. Data collection using the ERT process is computerized and rapid,

with about 45 minutes required to obtain cavern dimensions for a given test. However, the spatial

resolution of the cavern boundary is limited due the characteristics of the electric field and the

algorithm and grid size used for image reconstruction. The measurement error of the cavern

boundary (based on the resolution of the image grid) is typically given as 5 to 10% of the chest

diameter – about 1.9 to 3.8 cm for our vessel.

w = 5.1mm

381mm

Electrodes Sensor plane

P5
50.8mm
80mm
625mm
P4

P3

P2
120mm

P1

Figure 3.1 Cross-section of cylindrical stock chest showing ERT sensor planes and range of impeller
positions used.
29

1000

100
y (Pa)

10

hardwood
softwood
0.1
0.1 1 10
Cm(%)

Figure 3.2 Suspension yield stress as a function of mass concentration.

Ultrasonic Doppler velocimetry (UDV) was used to measure velocity profiles across the

mixer and has been used successfully on pulp suspensions in other flow configurations

(Dietemann and Rueff, 2004; Xu and Aidun, 2005; Ein-Mozaffari et. al., 2006). A single probe is

used to transmit and receive the ultrasonic signals. An ultrasonic pulse, emitted at a set frequency,

travels through the suspension and is reflected back to the probe by the fibres suspended in the

flow. The time delay (t) and frequency shift (Fd) measured by the probe allow the velocity (v) of

the particles (fibres) a distance (X) along the probe sampling direction to be determined

ct
X  (3.5)
2

Fd c
v (3.6)
2 Fe cos
30

where c is the speed of sound in the suspension, Fe is the frequency emitted from the transducer

and  is the Doppler angle (the angle the particle trajectory makes with the propagation direction

of the ultrasonic wave). A commercial UDV instrument (DOP2000, Signal Processing,

Switzerland) was used in our study. The probe was placed at various locations around the vessel

periphery to locate the points at which the suspension velocity fell to zero and hence the cavern

boundaries. The measurement accuracy of the UDV technique is good: 5% for velocity and 1%

for position (Takeda 1991) which is equivalent to a spatial resolution of about 4 mm in this study.

However, reconstruction of the cavern boundary for one mixing condition takes about 2 days

because data collection and analysis are time consuming. Results obtained for selected tests

made with UDV and ERT were compared to determine the most suitable technique for our study.

The point at which the cavern completely filled the vessel was readily determined by

visual inspection as the laboratory chest is fabricated from Plexiglas and is transparent. In

addition, video clips were taken of selected tests to confirm the cavern shapes/dimensions by

locating the cavern boundary at the vessel wall and suspension surface. Image orientation

between visual, ERT and UDV data was accomplished easily using the ERT electrodes as

reference points.

3.3 Results and discussion

3.3.1 Selection of cavern measurement technique

Electrical resistance tomography (ERT) was able to image caverns formed in the pulp

suspensions once a conductive tracer was added to the well-mixed region around the impeller

and dispersed throughout the cavern. Initially, a concentrated saline solution was used as the
31

tracer. However, the reconstructed images did not agree with other measurements (visual

inspection of the cavern surface in contact with the chest wall or with UDV measurements).

Other difficulties occurred. For example, when the saline solution was injected into the mixed

region, although the signal (the conductivity contrast between traced and non-traced regions) was

initially strong, the reconstructed image became blurred or completely disappeared as tracer

spread and was diluted through the cavern. To overcome the difficulties with suboptimum

conductivity contrast, we attempted to continuously add saline solution to the cavern or to

increase the initial concentration/dosage. In both cases the imaged cavern was larger than that

observed by other techniques which may be due to convection or diffusion of the aqueous tracer

beyond the cavern boundary.

To create reliable cavern images and avoid concerns about potential tracer diffusion,

conductive tracer particles were used. These particles were made by covering small pieces of

sponge with aluminum foil, with a typical particle being cylindrical, 5 mm in diameter and 17.5

mm in length, and having a volume of approximately 0.29 cm3 – a bit larger than that of a typical

fibre floc. The tracer particle density ranged from 0.90 to 1.06 g/cm3 (they could absorb water

and change density with prolonged use) and would move inside the cavern with the fibre

suspension. The number of particles deployed in a test depended on the volume of the cavern to

be measured and was optimized to obtain a good ERT image. Typically 100 to 120 particles

were introduced into the mixing zone near the impeller – corresponding to about 0.2% of the

cavern volume. The particles are shown in Figure 3.3. Figure 3.4 compares ERT images

reconstructed for a one set of test conditions (hardwood pulp at Cm = 3% with N = 650 rpm)

made using both the saline solution and the conductive particles as tracers. The cavern shape in

each measurement plane is represented (P1 is at the base of the chest and the impeller is located
32

between electrodes 5 and 6 and imaging planes 2 and 3, see Figure 3.1). In all cases the cavern

size imaged is larger when the saline tracer solution is employed. Figure 3.5 shows that the ERT

image obtained using tracer particles (on the right) agreed best with visual inspection of the top

surface of the suspension (note that P5 is 4.85 cm below the suspension surface).

Figure 3.3 Photograph of conductive tracer particles used in tests. Individual particles range in
diameter from 4 to 6 mm (typically 5 mm) and length from 15 to 20 mm (typically 17.5mm).

UDV was used to measure fibre suspension velocity profiles at eight different positions

spaced evenly between the ERT sensors on each sample plane as illustrated in Figure 3.6. The

cavern boundary was located at the points where suspension velocity fell to zero along this

measurement chord. By rotating the probe around the vessel, an image of the cavern was created

(Figure 3.7). Repeating these measurements for the each imaging plane allowed a three

dimensional reconstruction of the cavern to be created that was directly comparable with the

ERT measurements and from which the cavern volume could be calculated.
33

Tracer particle Saline solution

Plane 1 (the lowest plane)

Plane 2

Plane 3

Plane 4

Low conductivity High conductivity

Figure 3.4 ERT images obtained for a Cm = 3% hardwood pulp suspension agitated at N = 650 rpm
with E/D = 0.4, C/D = 0.7 and Z/T = 1.0. Images on the left were reconstructed using data obtained
with tracer particles (Fig. 3.3) and images on the right were reconstructed by adding a conductive
fluid of 10.9 mS/cm NaCl (15mL) to a pulp suspension having a background conductivity of 0.2
mS/cm (43.4L). The impeller is located between planes 2 and 3 and electrodes 5 and 6 as shown for
reference purposes in the upper left image.
34

10
1

9
2

8 3

7 4
6 5

Tracer particle Saline solution

Figure 3.5 Comparison of ERT images reconstructed using the tracer particles (left) and saline
solution (right). Uppermost sampling plane (P5). The visual surface view is given in the centre
image with the dotted line enclosing the region of active surface motion (determined visually).
(Hardwood pulp suspension at Cm = 3%; N = 650 rpm, E/D = 0.4, C/D = 0.7, Z/T = 1.0).

13
12 14

11 15

UDV
10 16 probe

9 1

8 2

measuring line
7 3

6 4
5

Figure 3.6 Probe locations for UDV measurements. The probe is moved around the chest to obtain
the velocity profiles needed to map the cavern boundary. The arrows show the direction of the sonic
emissions and the numbers give the ERT electrode positions.
35

70

a) 60 b) 13
12 14
50 11 15

40 10 16
Velocity (mm/s)

30
x1
9 1
x2
20
8 2
10
Cavern region 7 3
0 6 4
x1 x2 5
-10
0 100 200 300 400
Distance (mm)

Figure 3.7 (a) Velocity profile measured by UDV for a representative probe location. The locations
where the velocity reaches zero indicate the cavern boundary. (b) Cavern cross-section formed by
connecting the boundary points measured from all sample locations.

The cavern areas for each sample plane and the total cavern volume are summarized in

Table 3.1 for a representative set of tests where both the ERT (using tracer particles) and UDV

techniques were used. A visual comparison of the cavern images measured in each plane for one

test condition is also presented in Figure 3.8. There is no comparison of images for P3 because

the design of the chest could not allow the velocity measurement using UDV probe at that plane.

Cavern images reconstructed using both ERT and UDV were similar, although the ERT data

consistently gave a larger cavern volume which we attribute to the image reconstruction

technique used. The UDV data underestimates the location of the cavern boundary as it measures

only the suspension velocity component in the direction of the probe. For example, in planes P4

and P5 (Figure 3.8), the cavern area measured by UDV does not extend to the vessel wall

although suspension flowed vertically (observed visually) there. Despite these differences, the

measured cavern volumes are within 10% of each other, with the ERT technique predicting
36

slightly larger cavern volumes. The uncertainty in cavern volume is estimated to be 16% for

ERT and 7% for UDV measurements (based on the accuracy to which the cavern boundary can

be located by each technique). ERT, with conductive particles used as tracer particles, was used

for all subsequent tests due to the considerable reduction in time required for data acquisition and

analysis.

Table 3.1 Comparison of ERT/UDV cavern measurements (Cm = 2% hardwood pulp, E/D = 0.6,
C/D = 0.7 and Z/T = 1.0)
Impeller Speed, N (rpm)
Cavern area or
250 200 150
volume
ERT UDV ERT UDV ERT UDV
2
P1 (dm ) 6.42 6.39 5.59 4.83 5.19 4.32
P2 (dm2) 6.28 7.08 5.84 4.94 5.44 5.01
P4 (dm2) 7.11 5.62 5.37 3.88 3.67 3.48
P5 (dm2) 6.02 6.24 2.21 4.06 2.21 2.14
Volume (dm3) 24.6 24.1 18.1 16.9 15.7 14.2

12 13 14 12 13 14
11 15 11 15
10 16 10 16

9 1 9 1

8 2 8 2
7 3 7 3
6 4 6 4
5 5
P1 P2
12 13 14 12 13 14
11 15 11 15
10 16 10 16
9 1 9 1

8 2 8 2
7 3 7 3
6 4 6 4
5 5
P4 P5

Figure 3.8 Comparison between ERT and UDV caverns made for hardwood pulp at Cm = 2% and N
= 200 rpm (E/D = 0.6, C/D = 0.7 and Z/T = 1.0).
37

The cavern geometry produced in a 2% hardwood pulp suspension is reconstructed in

Figure 3.9 from ERT data at three impeller speeds. The cavern is cylindrical in shape, although a

pronounced taper is evident towards the suspension surface at lower rotational speeds. As the

impeller speed is increased, the cavern enlarges, particularly at the suspension surface, with the

cavern extending to the vessel surface.

line of
rotation

N = 150 rpm 200 rpm 250 rpm

Figure 3.9 Cavern shape (shaded volume) determined for a Cm = 2% hardwood at various impeller
speeds using ERT data (E/D = 0.6, C/D = 0.7, Z/T = 1.0).

3.3.2 Effect of impeller speed and pulp mass concentration on cavern volume

The cavern volume increased with increasing impeller speed as shown in Figure 3.10. The

growth in cavern volume is fairly uniform until the cavern boundary begins to interact

significantly with the vessel walls (about half the suspension volume is agitated at this point).

Any further increase in impeller speed rapidly causes the cavern to fill the vessel. As suspension

mass concentration (and consequently the suspension yield stress) was increased, higher impeller
38

speeds were required to create a given cavern volume (in agreement with Eqns. 3.1 through 3.3).

However, in all cases once the cavern had grown to about half the vessel volume a more

dramatic increase in cavern volume was created for a modest change in impeller speed. This step

change is similar to the dramatic change in channeling that was observed for small changes in

impeller speed in dynamic mixing tests made previously (Ein-Mozaffari et al., 2005). Perhaps we

have found a physical reason for the sudden transition to channeling.

3.3.3 Effect of pulp type on cavern volume

Figure 3.10 also includes a plot of cavern development for a softwood pulp suspension at

Cm = 3%. Here the suspension yield strength is significantly larger than that of the corresponding

hardwood pulp, which results in a reduced cavern volume for a given impeller speed. Pulp fibre

suspensions display unique flow behavior that depends on many aspects of the suspension,

primarily the suspension mass concentration but also the average fibre length and fibre length

distribution, fibre surface properties and the suspending medium. Although the suspension yield

stress is readily characterized, suspension behavior in flow has not yet been adequately described

by a rheological model.
39

0.040

3
0.035 suspension volume = 0.0347m

0.030

Vc ( m ) 0.025
3

0.020

0.015

Cm(%) y(Pa)
0.010
HW 2 9
HW 3 38
0.005 HW 4 69
SW 3 119
0.000
0 200 400 600 800 1000
N (rpm)

Figure 3.10 Effect of impeller speed (N) and pulp consistency (Cm) on cavern volume (Vc ) for
hardwood and softwood pulps (E/D = 0.4, C/D = 0.7 and Z/T = 0.8).

3.3.4 Impeller offset from the rear wall of the mixing chest

The location of the impeller in an industrial vessel is fixed, with the typical offset clearance

from the rear wall being from E/D = 0.3 to 1.0 (AGIMIX International, Uddevalla, Sweden).

Proximity to the wall can restrict suspension flow to the impeller and hence reduce the pumping

ability of the impeller. The impact of impeller clearance on cavern volume is shown in Figure

3.11 for our laboratory vessel. Here a Cm = 2% hardwood kraft pulp suspension was agitated

using two different impeller clearances. When the off-wall clearance was small (E/D = 0.4) the

cavern volume was smaller than when the off-wall clearance was larger (E/D = 0.6) for a fixed

impeller speed. As the power required by the impeller (at a fixed speed) was about the same

(Figure 3.11b) the larger offset resulted in slightly more energy-efficient mixing. However, any
40

increase in impeller offset may increase stresses on the impeller shaft and a balance between

increased pumping action and the mechanical installation must be achieved.

3.3.5 Comparison of cavern volumes determined by ERT and the cavern models

The predictions for cavern volume made with the available literature correlations are

compared with our experimental data in Figure 3.12. Here cavern volume is plotted against the

yield stress Reynolds number, with predictions made using the Amanullah et al. (1998) model

for a spherical cavern and the Elson (1990) model for a right circular cylinder. In both cases the

specific power number and/or axial force number measured for the impeller under the test

conditions (identical impeller geometry, impeller speed, pulp type and suspension mass

concentration, see data in Figure 3.13) was used.


41

a)

0.05

3
suspension volume = 0.0434m

0.04

0.03
Vc ( m )
3

0.02

0.01
E/D = 0.6
E/D = 0.4

0.00
100 200 300 400
N (rpm)

b)
0.05
3
suspension volume = 0.0434m

0.04

0.03
Vc ( m )
3

0.02

0.01
E/D = 0.6
E/D = 0.4

0.00
0 2 4 6 8 10
Power (W)

Figure 3.11 Effect of impeller position (E/D) on cavern volume (Vc) with (a) rotation speed and (b)
power (Cm = 2% hardwood, C/D = 0.7, Z/T = 1.0).
42

0.040

3
0.035 suspension volume = 0.0347m

0.030
Amanullah et al. (1998)

0.025
Vc ( m )
3

0.020
Measured Predicted
HW 2%
0.015 HW 3%
HW 4%
0.010 SW 3%

0.005
Elson (1990)

0.000
0 20 40 60 80 100
Rey

Figure 3.12 Comparison of measured cavern volumes and predictions made using the equations of
Amanullah et al. (1998) and Elson (1990).

The predictions of the Amanullah et al. (1998) model show the best agreement with the

experimental data, although the cavern formed is clearly cylindrical in geometry (Figure 3.9).

Amanullah‟s model predicts more rapid cavern growth with increasing Rey than found. Also, the

rotational speed required for the cavern to fill the vessel is under predicted. The model by Elson

(1990) significantly under-predicts cavern volumes for all test conditions despite modeling the

cavern as a cylinder. The main explanation for this is that axial thrust is not explicitly included in

the model. Elson also noted that once the cavern reached the baffles a further increase in impeller

speed increased cavern size at a slower rate than prior to contact with the vessel boundary,

opposite to that found here. As both models assume no interaction with vessel walls the

successful prediction of cavern volume in a side-entering geometry is not expected.


43

Consequently, we sought to develop a model where cavern interaction with the vessel was

accounted for.

a)

100

10

Hardwood Cm(%)
Nf

0
1
1 2
3
4

0.1
0 200 400 600 800 1000 1200
N (rpm)

b)

100

10

Hardwood Cm(%)
NP

0
1
1 2
3
4

0.1
0 200 400 600 800 1000 1200
N (rpm)

Figure 3.13 Axial force number (Nf) (a) and power number (NP) (b) for the D = 140mm Maxflo
impeller as a function of impeller speed and suspension mass concentration.
44

3.3.6 Cavern model including interaction with vessel walls

A model for cavern development with interaction between the cavern and vessel walls due

to impeller location is developed below. We begin with the equation developed by Amanullah et

al. (1998) to estimate the force transferred by an axial flow impeller to a spherical surface

2
 4 Po 
F  N D 2 4
N 
2
 (3.7)
 3 
f

This force is assumed to be balanced with the total force resisting motion at the cavern surface

regardless of the cavern geometry. For unbounded caverns this is the suspension yield stress at

the cavern-suspension interface. For bounded caverns, as considered here, the total resistive

force can arise from interactions with three surfaces: the cavern-suspension surface, Sp, the

cavern-wall surface, Sw, and the cavern-air surface, Sa. Equating the force transferred by the

impeller (Eqn. 3.7) with the total contribution from these three surfaces gives

F   y SP   w Sw   a Sa (3.8)

where y is the pulp suspension yield stress (measured as discussed previously), w is the friction

force at the vessel wall (which must be measured or estimated as detailed later), and a is the

friction force per unit area at the air suspension interface, taken as zero.

To proceed further we must specify the shape of the cavern. To a good approximation the

cavern is a truncated (due to interaction with the side of the mixing chest) right circular cylinder

(see Fig. 3.9) of radius rc (here the cavern “radius” is measured from the centre of the impeller to

the cavern boundary in front of the impeller) and height Z (extending from the base of the vessel

to the suspension surface). Once the cavern shape is specified the appropriate surface areas SP, Sw

and Sa must be calculated to use in Eqn. 3.8.

Analytical equations for the area and arc lengths of two intersecting circles are given by

Martin (1874) and adapted to the cavern-vessel geometry as shown in Figure 3.14. The cavern-
45

vessel surface area, Sw, is given by two components – the interaction at the base of the vessel, Sb,

and interaction along the side walls. The surface area where motion occurs at the base of the

vessel is given by

2
 r 2  a 2  rc 2   a 2  rc 2  rv 2   a 2  rc 2  rv 2 
S b  rv cos  v
1   rc 2 cos 1    a rc 2   
2
(3.9)
 2arv   2arc   2a 
     

where the impeller is located at the centre of a virtual cavern. Here rv is the vessel radius (rv =

T/2), rc is the cavern radius and a is the distance between the centre of a virtual cylindrical

cavern (the impeller location) and the centre of the mixing vessel (a = (T/2-E)). To find the

cavern-side wall and cavern-suspension surface areas, the distance, b, perpendicular to the line

running from the impeller to the centre of the vessel and the point where the cavern intersects the

vessel wall must be determined. We calculate this using Pythagoras‟ theorem, giving

rc  x 2  b 2 (for the cavern) and rv  (a  x) 2  b 2 (for the chest)


2 2

rc  rv  a 2
2 2

b  rc  x 2 where x 
2
(3.10)
2a
allowing the total cavern-wall surface area, Sw,

 b b
S w  2 2 rv Z  Sb  2  sin 1  rv Z  Sb where  2  sin 1   (3.11)
 rv   rv 
and the cavern-suspension surface area, Sp

 b b 
S p  21rc Z  2  sin 1  rc Z where 1  sin 1   (3.12)
 rc   rc 
to be determined. Z is the height of suspension in the chest (the height of the cavern).

The cavern volume was calculated using an iterative procedure. The three surface areas,

Sa = Sb, Sp and Sw were determined as a function of rc for the chosen mixer geometry (given D, T,

E). The resistive force contributed by these interfaces (Eqn. 3.8) was then balanced with the
46

force generated by the impeller (Eqn. 3.7) to determine the cavern radius associated with a given

impeller speed. The cavern volume was then computed by multiplying the floor surface area Sb at

this point by the stock height, Z.

Chest wall

rc b rv

1 2
x
a

Impeller

Cavern region

Figure 3.14 Diagram showing intersecting circles with the impeller located at the centre of a virtual
cylinder. The overlapping area between the mixed zone and mixing chest is the cavern region.

To use Eqn. 3.8 we must determine the resistive force at the suspension-wall interface, w,

which we did by measuring the rotational speed at the point where the cavern first completely

filled the vessel. The impeller power number and flow number were determined using the data in

Figure 3.13 which permitted the force at the interface to be calculated using Eqn. 3.7. (The

power numbers measured during the cavern tests and those determined using the data from

Figure 3.13 agreed within 3%). At this point Sp = 0, a = 0. Sw is given by the vessel geometry

and w was obtained by dividing the total applied impeller force by Sw (shown in Figure 3.15(b)

for hardwood pulp as a function of suspension mass concentration).


47

a)

1600

N (rpm) 1200

800

400
Z/T
0.8
1.0
1.2
0
0 1 2 3 4 5
Cm(%)

b)

200

150
w (Pa)

100

50
Z/T
0.8
1.0
1.2
0
0 1 2 3 4 5
Cm(%)

Figure 3.15 Rotation speed (a) and wall stress calculated (b) for complete mixing in the vessel with
hardwood pulp as a function of suspension mass concentration (E/D = 0.4, C/D = 0.7).
48

w can also be estimated from published pipe friction data. We used the correlation

presented in the TAPPI (1998) technical information sheets to do this. Here the head loss in a

piping system is correlated by

H/L = f1 K V Cm T (3.13)

where the factors f1, K,,  and  are given for a range of pulp types  for a hardwood pulp: f1 =

1.15, K = 236,  = 0.27,  = 1.78 and  = -1.08 (TAPPI TIS 0410-14). The chest diameter is T =

381mm. The velocity at which the head loss is determined must be specified and we chose

values of V = 0.001, 0.01 and 0.1 m/s for our calculations although the suspension velocity along

the cavern-wall interfaces will vary with position throughout the chest and with impeller speed.

Cm is the suspension mass concentration (%). Figure 3.16 compares these „pipe flow‟ estimates

of w with those determined using the point at which the cavern first filled the vessel and the

measured values of the suspension yield stress, y. The pipe flow estimates of w were much

lower than the other values.

We are now in a position to predict cavern development as a function of impeller speed in

our mixing vessel. The axial force and power number at a given impeller speed and suspension

consistency were determined using the data in Figure 3.13. The force generated by the impeller

was iteratively balanced with the forces at the cavern interfaces, as described previously. We

examined two scenarios. In the first (Figure 3.17a) we use the suspension yield stress as

measured in the rheometer for y (data in Figure 3.2), w determined at the point where the cavern

first filled the vessel (using the procedure described above), and a = 0. Figure 3.17a compares

model predictions (solid lines) with the experimental data (ERT) for hardwood pulp at three

mass concentrations. The dashed lines show the predictions with the suspension yield stress

varied by 20% (the average coefficient of variation for the yield stress measurements was 22%,
49

with a 20% variation in y giving a confidence interval of approximately 95%). The error bars

shown for selected experimental data represent 16% uncertainty in the cavern volume

measurements made using ERT. The figure shows that the model predicts the trend in cavern

development well, particularly for the suspensions at Cm = 3 and 4%. In general the experimental

points are reasonably close to those predicted by the model (volume predicted to an average of 

21%). Note that because the model was used to evaluate w each curve intersects the

experimental point where the cavern first filled the vessel.

200

150

Measured y (HW)
w, y (Pa)

Experimental 
100
Calculated  TAPPI)
v=0.001m/s
v=0.01m/s
v=0.1m/s
50

0
0 1 2 3 4 5
Cm (%)

Figure 3.16 Comparison of wall stresses measured/estimated and suspension yield stress for
hardwood pulp as a function of mass concentration. Estimates for w based on the yield stress (y -
measured), the point at which the cavern first completely filled the mixing vessel, and friction losses
in flowing pulp suspensions.
50

From Figure 3.16 we see that the values for w measured using the onset of complete

motion in the vessel are almost identical to the values of y measured by rheometry. Indeed, it

seems that this estimate of w evaluates y. In any case, an argument can be made for requiring

that the suspension yield stress must be exceeded at all cavern interfaces. This is because the

suspension yield stress must be exceeded at all suspension-suspension interfaces just prior to the

cavern reaching the vessel wall or suspension-air interface. Thus for the second scenario, given

in Figure 3.17(b) the computations were repeated with a = w = y with y determined by

rheometry. These predictions agree better with the experimental data, and require that only one

suspension measurement – the suspension yield stress – be made. For both scenarios, cavern

development (shape of the volume vs. impeller speed curves) is largely determined by cavern

geometry and its interaction with the vessel walls. However, the model (using a = w = y) was

found to predict the cavern volume to an average of 13%, which is better than that found for

previous models. This accuracy depends primarily on the yield stress which was determined to

22%. In the future, computational fluid dynamics may improve cavern estimation by

calculating the cavern geometry directly; however, the specification of suspension rheology is

still an issue and subject to measurement inaccuracies.


51

a)

0.04
3
suspension volume = 0.0347m

0.03
Vc ( m )
3

0.02

Cm(%)
2
0.01 3
4
Model
+/-20% y
0.00
0 200 400 600 800 1000
N (rpm)

b)

0.04
3
suspension volume = 0.0347m

0.03
Vc ( m )
3

0.02

Cm(%)
2
0.01
3
4
Model
+/-20% y
0.00
0 200 400 600 800 1000
N (rpm)

Figure 3.17 Cavern volume versus impeller speed for laboratory mixer. Comparison of
experimental data (measured by ERT) with model predictions for hardwood pulp. Z/T = 0.8, E/D =
0.4. (a) Cm = 2%: y = 9 Pa ;w =12 Pa ; a = 0; Cm = 3%: y = 38 Pa ;w = 42 Pa ; a = 0; Cm = 4%: y =
69 Pa ;w =90 Pa ; a = 0. (b) (y = w = a) Cm = 2%: y = 9 Pa; Cm = 3%: y = 38 Pa; Cm = 4%: y = 69
Pa.
52

3.4 Summary

Since pulp suspensions are opaque, cavern size must be imaged using indirect techniques. We

evaluated two methods, ERT and UDV, and found that both techniques gave satisfactory

measurements from which cavern shape and volume could be determined. However, ERT was

chosen due to the marked decrease in the time required to acquire data. The shape of the cavern

was best approximated as a truncated right-circular cylinder, and as expected, increasing impeller

speed increased cavern volume. However, development of cavern volume with increasing

impeller speed was not uniform, which was attributed to interaction between the cavern and the

vessel walls.

Current models for predicting cavern development in yield stress fluids treat the cavern

as being isolated with no interaction with the vessel walls and do not describe our experimental

results well. This prompted us to develop a model that included interaction with vessel

boundaries which required specification of cavern shape (based on ERT imaging) and

development of equations to balance the force provided by the impeller with the net resistive

force at the cavern boundaries. The model proposed predicts the trend in cavern volume with

increasing impeller speed well although the absolute cavern volume is only predicted to within

13%. These predictions are sensitive to the values of suspension yield stress which was measured

to 22%.
53

4 DYNAMIC TEST STUDY ON LAB-SCALE CHEST2

4.1 Introduction

Agitated stock chests are used in conjunction with process control algorithms to maintain process

and product uniformity during the manufacture of pulp and paper products. Process control is

effective at attenuating process disturbances below the cut-off frequency of the control loop,

which is determined by process dynamics and the type of control algorithm and tuning employed

(Bialkowki, 1992). Aggressive tuning increases the cut-off frequency, although over aggressive

tuning can create oscillations which amplify variability. The agitated stock chest acts as a low-

pass filter to remove the high-frequency variability and extend the range of disturbance

attenuation.

In most cases stock chests are designed based on past experience and semi-empirical

techniques (Yackel, 1990; Reed, 1995). Two common chest geometries are employed for pulp

suspension mixing: rectangular and cylindrical. Rectangular chests save space when grouped

together using common walls (as in the basement of a papermachine) while cylindrical chests

can take advantage of hoop stress design which allows the use of thinner walls for vessels of the

same height. For cylindrical pulp stock chests, complete motion across the suspension surface is

most readily achieved when the stock height (Z) to chest diameter (T) ratio is 0.8 (Z/T = 0.8)

(Yackel, 1990). An axial flow impeller mounted in a side-entering mode is typically used in this

application. Guidelines for impeller installation, including impeller offset from the rear chest

wall (E) and height from the chest bottom (C) are typically E/D = 0.3 – 1.0 and C/D = 0.5 – 1.5

(where D is the impeller diameter) (AGIMIX International, Uddevalla, Sweden). A common

2
A version of this chapter will be submitted for publication. Leo K. Hui, C.P.J. Bennington and G.A. Dumont (2011)
Mixing Dynamics in Cylindrical Pulp Stock Chests.
54

design strategy matches the momentum required to agitate a pulp suspension in a given

application with that generated by the impeller (Yackel, 1990). Cylindrical chests are widely

used in the pulp and paper industry, particularly for creating controlled mixing zones in

bleaching towers and in high-density pulp storage chests.

Agitated stock chests often do not behave ideally (Ein-Mozaffari et al., 2003). When a

non-Newtonian fluid, such as a pulp suspension, is agitated, a cavern (a region of active motion

around the impeller) is created (Wichterle and Wein, 1981; Solomon et al., 1981; Silvester, 1985;

Elson, 1990; Amanullah et al., 1998; Wilkens et al., 2005; Hui et al., 2009). In batch operations

there is limited or no fluid flow outside the cavern; in continuous operations, suspension flow but

limited mixing occurs outside the cavern. Ein-Mozaffari et al. (2003) measured the efficiency of

several industrial pulp chests and found significantly reduced attenuation efficiency. The non-

ideal flows identified included bypassing or channeling (where part of the feed directly flowed to

the exit without passing through the impeller zone) and the formation of dead zones (regions of

stagnant suspension that were disconnected from or moved significantly slower than the bulk

flow). These phenomena were exacerbated by the non-Newtonian nature of the pulp suspensions,

reducing chest mixing capacity and the ability to attenuate process disturbances.

To identify the factors that affected mixing quality, Ein-Mozaffari et al. (2005) studied

mixing dynamics in a laboratory-scale rectangular stock chest using two pulp suspensions, a

shorter-fibred hardwood pulp and a longer-fibred softwood pulp, over a range of fibre mass

concentrations. It was found that the cavern volume, which was affected by the suspension yield

stress and the impeller rotational speed, played a significant role in determining mixing dynamics.

As the yield stress decreased or impeller speed increased, the extent of bypassing and dead zone

formation was reduced and the degree of disturbance attenuation improved. In addition, the
55

imposed flow configuration, determined by the position of the stock inlet and outlet points

relative to the cavern, affected mixing performance. Bypassing was minimized when the pulp

outlet was located within the cavern zone. Other studies with pulp suspensions (the majority in

rectangular chests) have confirmed these findings (Bakker and Fasano, 1993; Ein-Mozaffari et

al., 2003; Ford, 2004; Wilstrom and Rasmuson, 1998a).

In this chapter, the dynamic response of a cylindrical chest was examined as a function of

impeller speed for several pulp suspensions. The results were also compared with previous

findings for a rectangular chest configuration. A dynamic mixing model (Soltanzadeh et al., 2009)

was used to quantify mixing dynamics. Figure 4.1 shows the continuous-time dynamic model

used to represent the observed behavior of a pulp chest. The model incorporates two parallel

suspension flow paths through the vessel: one that enters the cavern created by the impeller and

one that bypasses this mixed zone and exits the chest with minimal mixing. f is the fraction of

pulp stock that channels through the chest while the remaining stock (1-f) enters the mixing zone.

1 and 2 are the time constants for each zone, respectively (with 1  2). Note that it is possible

to have transport delay, Td, in the step response signal (due primarily to flow delay in the process

piping prior to and following the chest). Using input-output data obtained when the chest is

excited with an appropriately designed signal allows the model parameters to be identified. For

the model chosen (Fig. 4.1), ideal mixing is approached when the extent of bypassing approaches

zero and the time constant for the mixing zone approaches the theoretical time constant of the

system, i.e., th = Vt/Q, where Vt and Q are the suspension volume and flow rate.
56

1
f
1+1s
+
u y
+

1-f 1
1+2s

Figure 4.1Dynamic model for non-ideal flow in agitated pulp stock chests (2>>1).

4.2 Experimental set-up and procedures

The transparent 1/10 scale-model of a cylindrical stock chest (T = 38.1 cm) with a laboratory-

scale (D = 16.5 cm) side-entering axial-flow impeller (Maxflo, Chemineer Inc., Dayton, OH)

used in this study is shown in Fig. 4.2. The impeller was located 7.0 cm from chest wall (E/D =

0.4) and 12.0 cm from the bottom of the chest (C/D = 0.7). The pulp suspension was pumped

from a feed tank, through the chest and then to a discharge tank. The suspension was added to

the pulp chest through an inlet pipe placed at the suspension surface near the wall opposite the

impeller (a thin plastic sheet was attached to the pipe opening to avoid pulp splashing as it

dropped into the chest). The pulp exit was located at the wall behind and below the impeller.

This stock entrance/exit configuration is optimal for forcing the suspension feed through the

cavern (Ein-Mozaffari et al., 2005).

The mixing dynamics were measured by adding a conductive tracer (saline solution) to

the pulp feed using a pseudo-random binary signal (PRBS) to control tracer addition (Fig. 4.3)
57

and excite the chest dynamics at the appropriate frequencies. The conductivity of the suspension

entering and leaving the chest was measured (Fig. 4.4 shows one typical data set) with the first

portion of the experimental data used to estimate the model parameters and these parameters

were then used to model the second part of the data. The MATLAB program used for this model

parameter estimation is shown in Appendix B. The model output for the latter portion of the test

was compared with the measured data and the extent of agreement expressed as:

 1 n  y computedi  y measuredi 
% fit  100%  1     (4.1)
 n i 1  y measuredi  mean( y measured )  
  

where ycomputed is the output calculated from the model and ymeasured is the actual output. The

effectively mixed volume (Vmix) was estimated from the identified parameters using Vmix = Q2(1-

f).

Dynamic tests were performed with hardwood and softwood bleached kraft pulps from

Domtar Inc. (Windsor, QC). The hardwood pulp had a length-weighted mean fibre length of 1.28

mm and the softwood pulp had a length-weighted mean fibre length of 2.96 mm. Suspension

concentrations of Cm = 2, 3 and 4% were used for the hardwood pulp and Cm = 1.8 and 3% for

the softwood pulp. The suspension yield stress for the pulps were measured using a Haake RV12

Rotovisco concentric cylinder viscometer (Thermo Fisher Scientific, Waltham, US) and a vaned

measuring bob, as described by Hui et al. (2009). The pulp flow rates through the chest were 7,

14 and 20 L/min, with the mean pulp residence time greater than 4 times the blend time (Yackel,

1990). The batch blend time was determined for a typical test condition (Cm = 3% hardwood

pulp at N = 500 rpm (the median impeller rpm used for the tests)) by measuring the response to a

tracer addition above the impeller using a conductivity probe located on the opposite wall, giving
58

a blend time of 6 s. The residence times at the highest flow rate of 20 L/min and at the lowest

flow rate of 7 L/min were 104 s and 298 s respectively (both > 24 s). The net power required to

create complete motion in the chest was determined using P = 2NM, where the impeller speed

(N) and net torque (M) were measured using an inductive-rotary torque transducer (model

0411IE50, Staiger Mohilo, Germany) mounted on the impeller shaft, with complete motion

determined by visually inspecting the vessel through the clear walls (i.e., the entire suspension

was moving at all the vessel boundaries, not just on the suspension surface).

For selected tests, electrical resistance tomography (ERT) was used to image flow

through the chest under conditions where a cavern existed. ERT is a non-invasive imaging

technique that reconstructs the conductivity distribution inside the mixing vessel using voltage

measurements made at the chest periphery. A step change of saline concentration was introduced

to the suspension feed with the vessel imaged using a P2000 ERT tomography instrument (ITS,

Manchester, UK) as a function of time. Four sensor planes (P1 - P4), each with 16 stainless steel

electrodes installed on the chest periphery, were used to measure the conductivity changes within

the chest and allowed the temporal reconstruction of the tracer‟s progress through the vessel.

ERT was also used under batch mixing conditions to compare the agitated cavern volume with

that measured using the dynamic tests (Hui et al., 2009).


59

a)

381mm (15”) w = 5.1mm

Inlet

50.8mm
625mm
(24.6”) P4

P3
70mm

P2
120mm
P1

Outlet

b)

Feed

Impeller
shaft

Exit

Figure 4.2 Cross-section (a) and photograph (b) of cylindrical stock chest. The arrows in (b) show
the direction of suspension flow in continuous operation.
60

9
Liquid
Tracer
7

1 1
1 1
5
8
5 6
Feed Discharge
Tank 3 4 tank
Stock
chest 6

7
2
7
1 : Transmitter
2 : Indicator
3 : Tachometer
4 : Torque transducer
5 : Magnetic flowmeter
6 : Conductivity sensor
7 : Pump
8 : Solenoid valve
9 : Pulsation dampeners

Figure 4.3 Schematic of apparatus used for the dynamic tests.

7
Estimation Verification

6
Conductivity (mS)

3
Input
Output
2
0 600 1200 1800 2400
Time (s)

Figure 4.4 Typical signals measured for model parameter identification.


61

4.3 Results and discussion

The accuracy of the dynamic model was verified by measuring the time delay (Td) and mixing

time constant (2) under conditions of complete vessel agitation and comparing them with

theoretical values. Under conditions of ideal mixing, the volume of the suspension in the piping

between the locations of conductivity measurement and the chest inlet and outlet should account

for the delay. The time delay (and its uncertainty) at several flow rates was estimated by

averaging the results of four independent tests made using hardwood pulp at Cm = 3% under

conditions where different extents of mixing existed in the vessel. Different impeller speeds

were used, but no trend in delay time with impeller speed was observed. The delay estimated

using the model was within experimental error of that calculated, as shown in Table 4.1 (the

uncertainty in the theoretical time delay was estimated using the accuracy of pipe and flow rate

measurements). Also, it was found that the 2 estimate (for complete mixing) was always

slightly larger than the theoretical value (by 6% on average, which is close to the error (7%) in

parameter estimation) (Table 4.2). As the preparation for each test is time-consuming, only one

trial was conducted at each test condition given in Table 4.3.

Table 4.1 Time delay (Td) measured versus theoretical values for hardwood pulp Cm = 3% (N = 425,
525, 672 and 810 rpm at each flow rate)
Flow rate (Q) Td estimated from Td based on flow
(L/min.) dynamic tests (s) and system geometry (s)
7 79.6  9.5 81.9  6.4
14 38.4  2.0 40.9  2.9
20 31.7  4.4 28.7  2.1
62

Table 4.2 Time constant (2) measured for complete mixing of hardwood pulp at Q = 14 L/min
(Theoretical 2 = 149s)
Percentage of fitting
Cm (%) N (rpm) 2 (s) Eqn. (4.1)
(%)
380 1557 91
2
460 14210 91
672 1634 92
3
810 16121 78
830 16311 93
4
1000 16413 91

The data from all the tests are reported in Table 4.3. On average, the identified model

parameters were found to fit the experimental data to 88%, with all cases fitted above 78% (Eqn.

4.1). The flow configuration used in the tests forces the pulp feed through the cavern before

leaving the chest and consequently, the parameters f and 1 were expected to be low or negligible.

This was found in all but six cases. In the cases where the bypassing was identified it was found

to involve less than 10% of the flow.

To verify the presence of a cavern and a poorly agitated zone in a partially agitated chest

under continuous operation conditions, ERT was used to image the mixing of a pulse of tracer

(220 s in duration) through the chest under typical operating conditions – agitation of a

hardwood pulp suspension at Cm = 3%, with the standard vessel geometry (E/D = 0.6, C/D = 0.7

and Z/T = 0.8) using an impeller speed that would lead to cavern formation (N = 425 rpm and Q

= 14 L/min). Under the test condition 2 was found to be 70s (Table 4.3), implying that the

cavern occupied about 47% of the vessel volume since the theoretical 2 of complete vessel

agitation was 149s.


63

The step change in tracer was followed as it entered the vessel at the suspension surface

and flowed towards the side exit located between ERT sensors 3 and 4 above plane 1 (P1), as

shown in the t = 0 image in Fig. 4.5. Note that before the injection of tracer, the conductivity in

the chest was low (0.6 mS/cm) as indicated by the blue color in the tomographic images. As the

saline solution entered the chest the local conductivity began to increase (to green in the images).

At t = 30 s the tracer moved to the impeller-side of the chest along the suspension surface by the

upper circulation loop and began to fill the top part of the vessel. The pulse then moved

downward and spread out to the lower planes (t > 30s). As more tracer accumulated in the chest,

the conductivity increased (indicated by the red colour in the images for t = 100s and 170s) and

the truncated cylindrical shape of the cavern can be clearly seen. Tracer injection stopped at t =

220 s, and the tracer began to be washed from the chest (t = 255 to 480s). Towards the end of the

test, the presence of tracer was observed to linger in regions of the vessel (note the green regions

in the top image plane, P4, of the t = 360 s image) which indicates regions of slower mixing. No

obvious sign of direct passing of tracer to the exit could be seen, which supports the “negligible

bypassing” results obtained from the dynamic tests. In addition, the ERT images show that

outside the cavern there is a stagnant region the tracer does not enter. This dead zone was also

observed visually during the experiment (lack of flow adjacent to the vessel wall). This suggests

that a fillet (the insertion of solid material in this region) could be used to reduce this unwanted

dormant volume.
64

Table 4.3 Dynamic test results


Hardwood Pulp Suspension
Mass Flow rate Rotation f 1 (s) 2 (s) Percentage of
Concentration (L/min) speed (rpm) fitting (%)
(%)
2.0 14 125 0 0 47 89
250 0 0 106 85
380 0 0 155 91
(%) 460 0 0 142 91
3.0 7 425 0.031 0 168 92
525 0 0 256 92
672 0 0 301 92
810 0.065 1.07 304 86
14 425 0.087 5.93 70 89
525 0.017 0 113 89
672 0 0 163 92
810 0 0 161 78
20 425 0 0 56 88
525 0 0 93 85
672 0 0 115 89
810 0 0 106 86
4.0 14 550 0 0 68 91
675 0 0 104 90
830 0 0 163 93
1000 0 0 164 91
Softwood Pulp Suspension
Mass Flow rate Rotation f 1 (s) 2 (s) Percentage of
Concentration (L/min) speed (rpm) fitting (%)
(%)
1.8 14 425 0 0 116 85
525 0.099 0 133 83
680 0 0 151 92
750 0.013 0 171 87
3.0 14 425 0 0 69 86
525 0 0 97 82
1020 0 0 160 91
1100 0 0 156 80
65

Side view along


Time(s) P1 P2 P3 P4
the dotted line
9 9 9 inlet
0 7
8 9 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
outlet 3 15 3 15 3 15 3 15
2 1 16 2 1 16 2 1 16 2 1 16
9 9 9 9
30 7
8 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 1 16 2 1 16 2 1 16 2 1 16
9 9
40 7
8 9 10
11 7
8 10
11 7
8 9 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 16 2 1 16 2 16 2 1 16
1 1
9 9 9
100 7
8 9 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 16 2 1 16 2 1 16 2 1 16
1
9 9 9
170 7
8 9 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 2 16 2 16 2 1 16
1 16 1 1
9 9 9 9
220 7
8 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 1 16 2 1 16 2 1 16 2 1 16
9 9 9
255 7
8 9 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 16 2 1 16 2 1 16 2 1 16
1
9 9
290 7
8 10
11 7
8 9 10
11 7
8 10
11 7
8 9 10
11
6 12 6 12 6 12 6 12
5 13 5 13 5 13 5 13
4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 1 16 2 16 2 1 16 2 16
1 1
9 9 9 9
360 7
8 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 1 16 2 1 16 2 1 16 2 1 16
9 9 9
480 7
8 9 10
11 7
8 10
11 7
8 10
11 7
8 10
11
6 12 6 12 6 12 6 12

5 13 5 13 5 13 5 13

4 14 4 14 4 14 4 14
3 15 3 15 3 15 3 15
2 16 2 1 16 2 1 16 2 1 16
1
low conductivity high conductivity

Figure 4.5 ERT images of a dynamic test at Cm = 3% and N = 425rpm (E/D = 0.6, C/D = 0.7, Z/T = 0.8). The
suspension inlet and outlet positions are shown in the images of P4 and P1 at time t = 0 s, respectively.
66

Figures 4.6 and 4.7 show that 2, and hence the mixing quality, increases with increasing

impeller speed until a plateau is attained. The error in determining 2 was ~7% (the average

standard deviation calculated in model parameter estimation using MATLAB for three hardwood

mass concentrations at Q = 14L/min). As impeller speed increased, the cavern enlarged

increasing the active mixed volume and hence 2. As suspension mass concentration increased,

the suspension yield stress increased which decreased cavern volume at a fixed impeller speed.

Thus a higher impeller speed was needed to maintain the same 2. At a fixed impeller speed, 2

decreased as the flow rate was increased because it is inversely proportional to Q (Fig. 4.7).

When Vmix was calculated using 2 (Fig. 4.8), it was found that for a fixed suspension

concentration, changing the flow rate did not significantly affect the mixed volume which

depended mainly on the impeller speed. The uncertainty of Vmix was estimated to be 10% based

on the error in measuring 2 and Q.

200

th =149s (14L/min)


150

100
 s

Hardwood Cm (%)
50
2
3
4

0
200 400 600 800 1000
N (rpm)

Figure 4.6 Effect of impeller speed (N) and pulp consistency (Cm) on time constant (2) (E/D = 0.6,
C/D = 0.7, Z/T = 0.8).
67

th =298s (7L/min)


300
Q(L/min)
7
14
 s 200 20

th =149s (14L/min)

th =104s (20L/min)


100

0
400 500 600 700 800 900
N (rpm)

Figure 4.7 Effect of flow rate (Q) on time constant (2) with rotation speed (Cm = 3%, E/D = 0.6, C/D
= 0.7, Z/T = 0.8).

0.04
Vt = 0.0347m3

0.03
Vmix (m )
3

0.02

Q (L/min)
0.01 7
14
20

0.00
400 500 600 700 800 900
N (rpm)

Figure 4.8 Effect of flow rate (Q) on mixing volume (Vmix) with rotation speed (Cm = 3%, E/D = 0.6,
C/D = 0.7, Z/T = 0.8).
68

Figure 4.9 shows the change in mixed volume (cavern size) with impeller speed for the

hardwood and softwood pulps at Cm = 3%. Vmix for the hardwood suspension is larger than that

for the softwood for all impeller speeds. However, comparison on the basis of suspension

concentration is not appropriate as cavern size is determined primarily by the yield stress of the

suspension (Hui et al., 2009). A better comparison can be made using hardwood and softwood

suspensions of the same yield stress. In Fig. 4.10, the cavern volume of a Cm = 1.8% softwood

suspension is compared with that of the Cm = 3% hardwood suspension (both have a yield stress

of 38 Pa). We see that the mixed volume is very similar, except at the lowest impeller speeds

tested where the softwood pulp had the higher mixed volume. This same trend was also reported

by Ein-Mozaffari et al. (2005) for tests in rectangular chests. Thus, in addition to the suspension

yield stress other suspension factors (which we can only categorize in the broadest sense of „pulp

type‟) also affect mixing dynamics. As impeller speed is increased to approach the fully mixed

volume, the differences in cavern volume are reduced within our ability to measure them. Other

factors may account for these observations, including the uncertainty with which the yield stress

can be determined experimentally or the effect of the fibre type on the impeller flow and power

number, which have not been widely studied.

Finally, mixing volumes calculated using the dynamic tests were compared with those

measured using ERT under batch conditions as shown in Fig. 4.11. For the batch tests, the

cavern was imaged using conductive particle tracers added to the impeller region (see Hui et al.,

2009). It was found that the mixed volume determined using the dynamic model, Vmix, was

slightly larger than that determined by ERT under batch conditions, Vc. This may be due to the

location of the feed pipe and the momentum added to the mixed suspension by the entering flow.
69

However, changes in flow rate (Q = 7, 14, 20 L/min) did not significantly change the mixed

volume.

0.045

0.040

0.035 Vt = 0.0347m3

0.030
Vmix, Vc (m )

0.025
3

0.020

0.015

0.010 Continuous Batch


(Dynamic) (ERT)
0.005 HW 3%
SW 3%
0.000
400 600 800 1000 1200
N (rpm)

Figure 4.9 Effect of pulp types (HW: hardwood and SW: softwood) on cavern size obtained in
dynamic and batch (ERT) operations at Cm = 3% (E/D = 0.6, C/D = 0.7, Z/T = 0.8).

0.04
Vt = 0.0347m3

0.03
Vmix (m )
3

0.02

0.01
HW 3%
SW 1.8%

0.00
400 500 600 700 800
N (rpm)

Figure 4.10 Effect of pulp types (HW: hardwood and SW: softwood) with similar yield stress
(38Pa) on mixing volume (Vmix) with rotation speed (N) (E/D = 0.6, C/D = 0.7, Z/T = 0.8).
70

0.04
Vt = 0.0347m3

0.03
Vmix, Vc (m )
3

0.02

Continuous Batch
0.01 (Dynamic) (ERT)
HW 2%
HW 3%
HW 4%
0.00
0 200 400 600 800 1000
N (rpm)

Figure 4.11 Comparison between cavern size obtained in dynamic and batch (ERT) operations at
various hardwood pulp consistencies (Cm) (E/D = 0.6, C/D = 0.7, Z/T = 0.8, Q = 14 L/min for the
dynamic tests).

Scaling procedures are used to design mixers based on data obtained in geometrically

identical mixers at a different, usually smaller, scale. Yackel‟s (1990) design procedure for pulp

and paper chests uses the concept of level momentum. The momentum number (Mo) and level

momentum ( Mo ) are defined as:

Mo  C1 N 2 D 4 (4.2)

Mo C1 N 2 D 4
Mo  2
 2
(4.3)
3 3
V V

where C1 is a constant that depends on impeller type and geometry. The V2/3 term is due to the

proportionality between chest surface area and chest volume, which also fits the experimental

criteria Yackel uses (of having complete motion across the surface) to determine good mixing.

According to Yackel, designing for equal level momentum produces equal bulk velocities in the

scaled vessel and thus similar mixing results:


71

2 4 2 4
C1 N1 D1 C 2 N 2 D2
2
 2
(4.4)
V1 3 V2 3

where N1, D1 and V1 are the impeller speed, impeller diameter and fluid volume used in the

smaller chest and V2 is the volume of the larger chest having an impeller diameter of D2.

Assuming C1 = C2 (geometric similarity), the corresponding impeller speed (N2) for a cylindrical

chest having a larger fluid volume (V2) is calculated using

1
 N 2 D 4V 2 2 3  2

N 2   1 41 2  (4.5)
 D V 3 
 2 1 

Since V1 and V2 are proportional to T13 and T23, respectively, and that D1/T1 = D2/T2 (geometric

similarity), N2 is given as:

1
 N1 2 D1 4T2 2  2
D 
N 2   4 2

  N1  1  (4.6)
 D2 T1   D2 

Other scale-up methods are used in literature, including constant power/volume (P/V) and

constant torque/volume (M/V). Constant P/V is the most commonly recommended scaling

criteria for mixing operations (Oldshue, 1983; Wilkens et al., 2003; Paul et al., 2004). This

method correlates well with mass-transfer characteristics in a mixer and is suitable for

calculating the power needed for gas dispersion or dispersion of immiscible liquids (Zlokarnik,

2006). Equal M/V is also a practical and common scale-up criterion because it relates directly to

the overall size, torque capability and cost of the mixer (Paul et al., 2004). It is the same as

scaling for constant tip speed and is usually applied when flow velocities in the impeller region

need to be the same in both vessels. The cavern model developed by Hui et al. (2009) can also be

used to scale pulp mixers. This model takes into account the interactions between the actively

mixed cavern and the vessel walls by balancing the forces produced by the impeller with those
72

acting at the cavern boundary. To create complete mixing in the vessel, the cavern boundaries

must extend to fill the mixer volume. Table 4.4 gives the dependence of N2 on N1 for four scale-

up methods considered and the calculation of N2 using the cavern model.

Table 4.5 gives the power predicted by scaling three- and ten-fold the volume of our

laboratory cylindrical chest using the criteria described above. Geometric similarity was

maintained (D/T = 0.43, Z/T = 0.8) with the base conditions measured in the 34.7L vessel for

complete motion. Of the methods considered, the constant P/V criterion gives the highest power

consumption. The constant M/V and level momentum scaling criteria give identical predictions

because, for geometric similarity, they both scale as N2=N1(D1/D2) as shown in Table 4.4.

Determining N2 using the cavern model does not require knowing N1, just the impeller

characteristics (Nf and Np) and suspension yield stress (y). The model predicts power levels close

to those calculated using the constant M/V and level momentum scaling methods. This is

expected as the cavern model also scales as N2=N1(D1/D2) for geometric similarity and constant

impeller behavior.

In industrial processes agitated pulp chests are usually rectangular or cylindrical. To

compare which shape is more efficient, the specific power to create complete mixing in a

rectangular chest (measured by Ein-Mozzaffari, 2002) was compared with that for a cylindrical

chest. As the cylindrical chest used in our study had a lower volume than that used by Ein-

Mozaffari, the base data was scaled to that of a chest having equal volume and allowing for the

different suspension height in the rectangular chest, Z/T = 1.2, and the fact that D/T changed

between chests. The power for the scaled cylindrical chest was found to be 1.4 W/kg and 4.8

W/kg for Cm = 1.8% and 3% softwood pulp suspensions, respectively. Both these predictions are

close to the values measured for the rectangular chest (1.4 W/kg at Cm = 1.8% and 5.4 W/kg at
73

Cm = 3%). Due to the use of scaling and the corrections needed to make this comparison we can

only state that vessel geometry, cylindrical or rectangular, requires approximately the same

amount of energy to create a well-mixed state.

Table 4.4 Calculation of impeller speed on scale-up using different criteria


Criteria Relationship Scaled impeller speed
P1 P2
Constant power  D 
2
3

per unit volume*


V1 V2 N 2  N1  1 
P  N p N 3 D 5  D2 
M1 M 2

Constant torque V1 V2 D 
N 2  N1  1 
per unit volume* P  D2 
M 
2N
Mo1 Mo2 D 
Level momentum* 2
 2 N 2  N1  1 
V1 3 V2 3  D2 
D 
Constant tip speed N1 D1  N 2 D2 N 2  N1  1 
 D2 
F   y A2
1
 
2
 4N p 
2

F  N 2 D2 N f  
2 4 2
   y A2 
Cavern model N 
 3  2 
 D2 4 N f 2  4 N p 3 2


(A2: surface area of cavern;  
Nf: axial force number of impeller)
*Assumptions: D/T = constant; Np = constant, geometric similarity

Table 4.5 Power predicted for complete mixing of a hardwood pulp suspension in cylindrical stock
chests using different scale-up criteria (D/T = 0.43, Z/T = 0.8)
Power (W)
Cm Scale-up Constant Constant Cavern
(%) y (Pa) volume ratio Level
power/volume torque/volume model
momentum
(Wilkens et al., (Wilkens et al., (Leo et al.,
(Yackel, 1990)
2003) 2003) 2009)
1:3 22 18 18 12
2 9
1 : 10 76 41 41 28
1:3 142 98 98 105
3 38
1 : 10 497 227 227 243
1:3 374 264 264 258
4 69
1 : 10 1310 593 593 579
74

Other factors should be considered when selecting a pulp mixer. Rectangular chests are

more susceptible to the creation of stagnant or dead zones, especially near the chest corners

(Ford, 2004). Also, a sudden transition to non-ideal behavior has been observed in a pilot-scale

rectangular chest under certain operating conditions (e.g., a dramatic increase of bypassing could

occur when the impeller speed was decreased slightly, all other operating conditions remaining

fixed) (Ein-Mozaffari, 2005). This was attributed to the location of the suspension feed and exit

relative to the impeller. This undesirable flow bifurcation could create fluctuations and/or

instability in the performance of an agitated pulp chest but was not observed in the cylindrical

chest. It can be avoided by ensuring that entrance and exit locations are optimized.
75

4.4 Summary

Dynamic tests were carried out in a laboratory-scale cylindrical chest to identify and characterize

the mixing quality. The dynamic model chosen, which included an actively mixed region with

the possibility of bypassing, was found to accurately describe the dynamics measured. The

insignificant bypassing measured indicated that the flow configuration used (with the exit located

within the active cavern, close to and below the impeller) is an effective configuration for

avoiding stock bypassing. Imaging, using electrical resistance tomography (ERT), verified the

size and shape of the caverns created when the chest was not fully agitated. Cavern size could be

increased at a given suspension concentration by increasing impeller speed, which improves

mixing quality. The power required to completely agitate a cylindrical chest was similar to that

required for a rectangular chest of the same volume.


76

5 CFD SIMULATION OF PULP MIXING

5.1 Introduction

Agitated stock chests are important in pulp and paper processing. They keep pulp suspensions in

motion to avoid dewatering and blend different pulp stocks or blend pulp with other chemicals.

They also act buffers between processes and act as low-pass filters to remove high frequency

variability and to complement the action of control loops. However, pulp chests do not always

function as expected. Ein-Mozaffari et al. (2003) studied the mixing performance of industrial

chests and noticed the non-ideal dynamic response of rectangular pulp stock chests. This

abnormality was attributed to improper chest design and the non-Newtonian rheology of pulp

suspension, which has a yield stress promoting the formation of caverns (regions of active

motion) around impellers. Under certain flow configurations, these caverns can induce

undesirable flow, like bypassing (part of the feed flows directly to the exit without entering the

cavern) and dead zones (stagnant regions outside the cavern) which degrade mixing and reduce

the ability of the chest to attenuate variability. Poor chest performance can lead to unstable

downstream processes and low product quality, resulting in high production costs.

Pulp stock chests are typically rectangular or cylindrical. Their design is based on the use

of semi-empirical methods and experience with a design strategy of matching the momentum

generated by an impeller with that required to provide complete suspension motion across the

surface of the chest (Yackel, 1990). This method does not specifically take the locations of pulp

inlet and outlets into consideration, although heuristics are provided. Laboratory studies using

tracer analyses, ultrasonic Doppler velocimetry (UDV) and electrical resistance tomography

(ERT) have investigated mixing and suspension flow through agitated stock chests. Ein-
77

Mozaffari et al. (2005) used dynamic test results to show that cavern size and its position relative

to inlet and outlet locations played an important role in determining the amount of feed entering

the cavern which affected the mixing dynamics of the chests. Bypassing was minimized when

the stock outlet was located within the cavern zone. In addition, Ein-Mozaffari et al. (2007)

measured the flow characteristics (impeller flow number and pumping rate) of a Maxflo impeller

and located the cavern using the velocity profiles measured by UDV. Hui et al. (2009) applied

ERT and UDV to measure the cavern volume in a cylindrical chest. A model was developed to

estimate the cavern size using a force balance with consideration of interaction between the

cavern and chest walls.

Although experimental studies provide some information about the mixing dynamics

inside stock chests, obtaining this data is very difficult due to the opaqueness of the suspensions.

Computational fluid dynamics (CFD) has been used to obtain information about the velocity

fields inside these and other vessels, although there are difficulties associated with modeling the

suspension and validating the simulations made. Many studies have shown that CFD can predict

the velocity distribution, evaluate industrial stirred equipment, and optimize impeller design

(Armenante et al., 1997; Bhattacharya and Kresta, 2002; Paul et. al., 2004; Li et al., 2005;

Ankamma Rao and Sivashanmugan, 2010). For pulp suspensions, a number of studies using

CFD have been carried out to investigate the flow in agitated pulp chests. Bakker and Fasano

(1993) applied CFD and used the rheological data from Gullichsen and Harkonen (1981) for pulp

suspension to model the flow in a rectangular pulp chest with a side-entering impeller. The

numerically simulated results were found to agree with visual observation of the suspension

surface. In the study of the impact of pulp suspension rheology on flow field in an agitated pilot

tank with a jet nozzle mixer, Wikstrom and Rasmuson (1998a) compared the CFD-calculations
78

with the measured results and concluded that a more sophisticated rheology model should be

used instead of the Bingham model because of the increasing deviation between the computed

and measured flow field as the distance from the impeller increased. Ford et al. (2006) developed

a CFD model of a rectangular pulp mixing chest and used the Bingham plastic model as the pulp

suspension rheology. The model captured the mixing dynamics of the chest fairly well but it

overestimated the extent of mixing in the bypassing flow, especially in flow situations with

significant bypassing. This deviation was attributed to the inability of the Bingham model to

fully describe the suspension rheology. In the CFD modeling of a cylindrical tank agitated with

an axial-flow impeller, Saeed et al. (2007) treated pulp suspension as a Herschel-Bulkley fluid.

The CFD simulations picked up the features of the flow field and the computed velocities agreed

satisfactorily with the measured results. The discrepancy between the computed and measured

velocities was believed to be due to the suspension rheology. Bhattacharya et al. (2010) modeled

two industrial pulp stock chests using CFD and described the pulp rheology as a modified

Herschel-Bulkley model. The computed dynamic response of the chests was found to agree

reasonably well with the experimental measurements and the simulated flow fields provided

insight about cavern formation, stagnant regions and bypassing zones created in the chests.

However, these studies have not provided a detailed examination of the cavern flow, which is

important in determining the non-ideal flow in pulp mixing chests. The extent of cavern growth

could determine the degree of bypassing and the size of dead zone in agitated chests.

In this chapter, the commercial CFD software (FLUENT 6.2, Fluent Inc, Lebanon, NH)

was used to model a cylindrical chest equipped with a scaled version of a commercial axial-flow

Maxflo impeller (Chemineer Inc., Dayton, OH) in the standard side-entering configuration for

simulating the mixing of two pulp suspensions (hardwood and softwood) over a range of
79

impeller rotational speeds (N) and suspension concentrations (Cm). Ford et al. (2006) have

presented a computational approach to simulate pulp suspension in a rectangular agitated chest

equipped with a side-entering agitator for a range of operating conditions. This similar approach

was used to simulate the pulp mixing in the cylindrical chest. Two non-Newtonian fluid models

were used to describe suspension rheology (based on rheological measurements). The simulation

results were then used to calculate the cavern size in batch mixing mode and to estimate the

dynamic response of the chest in continuous operation. The computed results were then

compared with the experimental results obtained in the studies of Hui et al. (2010). By

determining the degree to which the CFD model can predict experimental data, its usefulness as

a tool for chest design can be evaluated. If the evaluation is positive, the CFD model can be used

for optimization and economical design of pulp chests for a range of objectives, including

mixing efficiency, energy consumption and equipment manufacturing costs.

5.2 Experimental study of pulp mixing

A number of experimental methods were used to study the mixing behavior of pulp stock

chests. For batch operation, electrical resistance tomography (ERT) and ultrasonic Doppler

velocimetry (UDV) were applied to study the cavern shape and size. ERT is a non-invasive

imaging technique that can image the conductivity distribution inside a region of interest using

voltage measurements made through a series of electrodes placed at the chest periphery. With

the addition of a conductive tracer to the cavern, a P2000 ERT tomography instrument (ITS,

Manchester, UK) was used to locate the region of higher conductivity and thus the cavern shape

and size (Hui et al., 2009). The measurement error of the cavern boundary is typically about 5-
80

10% of the chest diameter. Also, the cavern results obtained from ERT were validated by visual

observations of motion at the vessel walls and the suspension surface. Another method used to

study caverns was UDV, a real-time fluid flow measurement based on the principle of Doppler

frequency shift. An ultrasonic pulse is emitted from the UDV probe (DOP2000, Signal

Processing, Switzerland) and the echoes reflected from the particles in the fluid are used to

determine the velocity profile along the measurement line with the velocity direction parallel to

the pulse path. The locations where the suspension velocity falls to zero are the cavern

boundaries. The measurement accuracy of UDV is about 2.4% for velocity (Xu, 2003).

For continuous operation, a dynamic mixing model (Fig. 4.1) was used to quantify the

mixing dynamics in the cylindrical chest. f is the fraction of pulp stock that bypasses the mixed

zone with minimal mixing and (1-f) is the remaining stock entering the mixed zone. 1 and 2 are

the time constants for each portion, respectively and Td is the transport delay due to the flow

delay in the process piping prior to and following the chest. The model parameters were

determined by adding a conductive tracer (saline solution) to the pulp feed using a pseudo-

random binary signal (PRBS) to control tracer addition and excite the chest dynamics. The

conductivity of the suspension entering and leaving the chest was measured and this data set was

used to determine the parameters in the dynamic model.

5.3 CFD modeling of pulp mixing

5.3.1 Computational geometry

The configuration of the mixing domain, identical to the lab-scale cylindrical chest was modeled

using GAMBIT, a geometry and grid generation software bundled with FLUENT (Figure 5.1(a)).
81

It was used for both batch and continuous-flow systems; except that for the dynamic simulations,

an inlet was created at the top of the surface and an outlet at 45 with respect to the impeller as

shown in Fig. 5.1(b). The flow in the mixing chest is not steady in an inertial frame of reference

because the impeller blades sweep the domain periodically. However, since no baffles or stators

were involved in the mixing chest, the flow could be viewed as steady relative to the rotating

(non-inertial) frame (Ford, 2004). Thus, a multiple reference frame approach, including a

rotating frame and a stationary frame, was used to model the mixing domain. The rotating frame

is a cylindrical region comprising the impeller with the impeller shaft acting as its axis. The rest

of the vessel is in the stationary frame. To obtain reliable computational results, it is desirable to

balance a fine mesh structure with the need of having a reasonable computational time. A mesh

of 700,000 cells took 2 to 3 days to converge (on a Pentium(R) 4 CPU 3.00GHz) while meshes

of 200,000-300,000 cells took about one day. To compromise between the running time and

computational reliability, a finer mesh was only used in the rotating domain and a check of mesh

independency was also carried out.

For the batch simulation, the mixing domain was divided into six blocks for meshing.

Unstructured mesh, using tetrahedral elements connected one other, was applied to the rotating

domain and its neighbouring block because of their irregular geometries. To capture flow details

near the impeller, a finer mesh was created in the vicinity of the impeller blades. A conformal

mesh (having element nodes at identical locations at the block interfaces) was created using a

size function in the rotating domain to allow mesh elements to grow slowly as a function of the

distance from the impeller hub and blades so that the fine mesh in the rotating domain were

logically linked to the surrounding stationary domain which was not highly meshed (Ford, 2004).

The other four regular blocks were discretized using a structured mesh, comprising hexahedral
82

elements logically connected. To ensure the simulated results were not dependent on the grid

density, three different mesh sizes (198,084, 407,592 and 633,013 cells) were used to check grid

independence under the same simulation conditions (i.e., N = 425 rpm and Cm = 3% softwood).

Stationary frame
a)

Rotating frame

Impeller

Impeller

b) Inlet
Inlet Stationary frame

Impeller Rotating frame


shaft

Impeller

Outlet

Outlet

Figure 5.1 Experimental cylindrical chest and computational domain for (a) batch and (b)
continuous-flow mixing.

5.3.2 Modelling suspension rheology

A pulp suspension is a heterogeneous fluid (fibres and water) that is not easily modeled. As it

possesses a yield stress, the first comparable model that can be used to describe its rheology is

the Bingham model. Wikstrom and Rasmuson (1998) used this approach to model the agitation
83

of pulp suspensions with a jet nozzle agitator and found that the flow field far away from the

boundary conditions was underestimated due to the shear-thinning property of pulp suspension.

The Herschel-Bulkley model, which contains a shear-thinning parameter in addition to a yield

stress, can also be used to model the pulp suspension rheology. Both models are expressed in

Fluent, as

 n  y  n 
 y  K1      
  o 

 (5.1)

where y, K1, n, o and  are yield stress threshold, consistency index, power-law index, yielding

viscosity and strain rate respectively. All these parameters depend on the mass concentrations

(Cm) of pulp suspensions and can be determined from rheological measurements of suspensions.

The rheological parameters for the Herschel-Bulkley model were determined by fitting the flow

curves of pulp suspensions in the study of Gomez et al. (2010) to the model equation. For the

Bingham approximation, the pulp suspension is described with K1 = 0.001kg/ms (equal to the

viscosity of water) and n = 1 since the pulp and paper industry often uses the water viscosity to

design equipment where the suspension is exposed to high shear rates (Ford, 2004). Both

rheological models were applied in the CFD simulations of pulp agitation. The rheological

parameters for hardwood pulp (Cm = 2%) and softwood pulp (Cm = 3%) are summarized in Table

5.1.
84

Table 5.1 Rheological parameters of hardwood and softwood pulps


Cm = 2% hardwood Cm = 3% softwood
Model Herschel-Bulkley Bingham Herschel-Bulkley Bingham
K1
4.97 0.001 135 0.001
(kgsn-2/m)
n 0.328 1 0.21 1
y (Pa) 9.44 9.44 155 155
o
9.6 9.6 10400 10400
(kg/ms)

5.3.3 Condition setup for CFD model

In the rotating domain, an angular velocity was specified equal to the impeller rotation rate used

in the experimental test. The rotation axis was set at the centre of the impeller hub, with the

positive Y direction extending into the vessel (Figure 5.1). The impeller hub and blades were

modeled as rotating walls and moved with zero relative velocity with respect to the rotating

frame. The continuity and momentum equations governing the rotating domain are given by:

   
     vr   S m (5.2)
 t 


where  is the fluid density, t is time, vr is the relative velocity and Sm is the mass added to the

continuous phase from the dispersed second phase and any user-defined sources.

 
t


  
v     vr v      v  p      g  F (5.3)

 
where v is the absolute velocity, p is the static pressure,  is the stress tensor, g is the
 
gravitational body force and F is the external body forces. The relative velocity, vr , is related to

the absolute velocity, v , using the following equation:
85



  
vr  v    r  (5.4)

 
where  is the angular velocity of the rotating frame, r is the position vector in the rotating

frame.

In the stationary frames, the chest walls were set with a velocity of zero in the absolute

reference frame. The upper surface of the suspension was treated as a stationary wall with zero

shear stress. Since no rotation is involved in the stationary domain, the continuity and

momentum equations applied in this domain are written as follows:

 
   v   S m (5.5)
t

 
t
 
v     vv   p      g  F (5.6)

At the boundaries between the two domains, the continuity of absolute velocity is enforced to

provide the correct neighbor values of velocity for the corresponding domain.

In the pulp mixing experiments, the observed suspension flow was generally laminar, so

the continuity and momentum equations were solved in laminar regime using a second-order

upwind scheme. Near the impeller, the flow might be turbulent, but the fluctuation velocities

would fade out quickly in the regions outside of the impeller zone because of the fibrous

structure of the heterogeneous suspensions. Also, the non-Newtonian Reynolds number used by

Gibbon et al. (1962) and Blansinski et al. (1972) for mixing pulp suspensions was calculated and

found to be below 103 (i.e., in the laminar regime), so treating the suspension flow as laminar in

the entire domain would be fine for CFD simulation.


86

When continuous operation was simulated, inlet and outlet pipes were added to the chest

geometry. The lengths (L) and diameters (Dp) of the pipes in the simulation followed the

dimensions of the experimental set-up, i.e., L = 215 cm and Dp = 5cm for the inlet and L =

257cm and Dp = 5cm for the outlet. The inlet was defined in the absolute frame as having a flat

profile with a velocity magnitude ( v  4Q D p ). Since L/Dp for the outlet was greater than 10,
2

the exit was defined as an outflow boundary condition under the fully-developed flow

assumption. The boundary conditions at the outflow boundary are a zero diffusion flux for all

flow variables and an overall mass balance correction. The zero diffusion flux condition means

that the conditions of the outflow plane are extrapolated from within the domain and have no

effect on the upstream flow.

5.3.4 Cavern volume determination and dynamic test simulation

Several researchers have determined the cavern boundary by assuming that the shear stress at the

cavern surface is equal to the yield stress of the suspension (y) (Solomon et al., 1981; Elson,

1990; Amanullah et al., 1998; Hui et al., 2009). Thus, this criterion was used to determine the

cavern boundary in the CFD simulations. Since the shear stress was not given in Fluent, the

strain rate at the yield stress (y) was used to define the cavern boundary

y
y  (5.7)
o

A User Defined Function (UDF) was used to determine those cells included in the cavern criteria

chosen to calculate the total cavern volume (Appendix C).

The flow results calculated from the CFD simulations were also used to determine virtual

dynamic responses to the experiments conducted by Hui et al. (2010). Since the properties of the
87

tracer and the modeled pulp suspension are similar, the converged results for the flow field of the

suspension were used to perform transient species calculation for the tracer (Ford, 2004). Here

the species conservation equation was solved in time-dependent form, discretized in both space

and time. A UDF (Appendix D) specifying the tracer injection sequence at the inlet was linked to

the FLUENT solver for the continuous-flow system simulation and the concentration of the

tracer „measured‟ at the outlet. The inlet-outlet concentration profiles were then analyzed using

the dynamic model of Soltanzadeh et al. (2009) to quantify the mixing and compare them

directly with the experimental results. The model parameters identified included the transport

delay (Td) arising mainly due to flow in the process piping before and after the chest, the

bypassing (channeling) fraction (f) of suspension which flows through the chest with minimal

mixing, and the time constant for the mixing zone (2). The data were divided into two parts: one

for model parameter estimation and one for parameter validation. The extent of agreement

between the model output and the simulated data was expressed using

 1 n  y computedi  y simulatedi 
% fit  1     100% (5.8)
 n i 1  y simulatedi  mean( y simulated )  
  

The time constant (2) was then used to estimate the effectively mixed volume (Vmix) in the chest

using the equation Vmix = Q2(1-f).

5.4 Results and discussion

The effect of the meshing scheme used in the CFD calculations was first examined. The velocity

profiles calculated at two locations in front of the impeller (as shown in Figure 5.2) for the three
88

meshing schemes are given in Figure 5.3. There is no significant difference among the three

schemes. Thus, the lowest density mesh was used for the simulations because of the shorter

computer running time required. Table 5.2 shows cavern volumes determined using the three

meshes described earlier, with cavern size decreasing slightly as the cell density increases (and

consequently the average cell size decreases). The volume difference is small, less than 2.5%,

and is attributed to the finer resolution of the cavern as cell density increases.

CFD simulations were made for a range of operating conditions (two pulp types, a range

of suspension concentrations and a number of impeller speeds). Figures 5.4 and 5.5 compare

cavern shape and volume measured using ERT and computed using CFD for Cm = 2% hardwood

and Cm = 3% softwood at three different impeller speeds. For the hardwood pulp (Figure 5.4),

when the stock height (Z) was equal to the chest diameter (T), all the simulations were unable to

predict the suspension surface motion observed visually during the experiments. The possible

reason for this deviation could be that the yield stress may not be consistent throughout the pulp

suspension during agitation. Since pulp suspension is a two-phase fluid, the pulp fibres would be

redistributed in the suspension by the vigorous action of the impeller in actual situation. Fewer

fibres would be in the region around the impeller and so the yield stress in the cavern could be

slightly lower than the surrounding, promoting the cavern growth. However, in CFD simulations,

the pulp suspension is assumed to be homogenous, with no localized difference in yield stress

and so the predicted cavern volume could be smaller. The best way to justify this explanation is

to repeat the experiments using a homogeneous fluid with known rheology (possibly similar to

the properties of pulp suspension) and then compare with the results of heterogeneous pulp

suspensions. Also, in actual mixing situation, the top surface of the agitated chest was in motion

and open to the air, allowing air to enter the chest and promoting the cavern growth which
89

enhances the surface motion. However, in CFD simulations, the top surface was modeled as a

flat stationary wall with no air entering and so the actual situation may not be easily simulated.

The two rheology models examined did not make a significant difference in the cavern

shape determined for hardwood pulp. Compared with the ERT results, the average percentage

differences in cavern volumes for both models are also similar: 53% for the Herschel-Bulkley

model and 49% for the Bingham model. However, the Bingham model predicted the cavern

volume better at higher impeller speeds. When the impeller speed was increased to the situation

close to complete mixing (i.e., 397rpm - experimental value for 2% hardwood pulp), the

Bingham plastic model using the viscosity of water for K1 and n = 1 gave the cavern volume

more comparable to the experimental result. This observation agrees with the study of Gullichsen

and Harkonen (1981), showing that pulp suspensions behave more like water at high rotational

speeds.

Chest wall

Impeller

(a) (b)
Locations where velocity profiles are calculated
at z = 12cm (measured from the chest bottom)

Figure 5.2 Locations where velocity samples were determined for comparison of three
different meshes.
90

a)

low mesh
1.0 medium mesh
high mesh

Velocity magnitude (m/s)


0.8

0.6

0.4

0.2

0.0
-0.20 -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 0.20
Distance from the shaft centre line (m)

b)
0.00035

0.00030
low mesh
medium mesh
Velocity magnitude (m/s)

0.00025 high mesh

0.00020

0.00015

0.00010

0.00005

0.00000
-0.20 -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 0.20
Distance from the shaft centre line (m)

Figure 5.3 Velocity profiles calculated for softwood Cm = 3% at two locations (a) and (b) in
front of the impeller (as shown in Figure 2) for three different meshing schemes at N =
425rpm.
91

Table 5.2 Grid independence study (N = 425rpm, 3% softwood pulp (K1 = 135, n = 0.21, y =
155Pa)
No. of element cells Vc (dm3)
198,084 9.2
407,592 9.0
633,013 8.9

Method ERT CFD


Rheology
Herschel-Bulkley Bingham
Model
K1 (kgsn-2/m) 4.965 0.001
n 0.328 1
y (Pa) 9.443 9.443

Cavern
shape
(150rpm)

Vc (dm3) 15.7 6.2 5.4

Cavern
shape
(200rpm)

Vc (dm3) 18.1 9.3 9.2

Cavern
shape
(250rpm)

Vc (dm3) 24.6 12.2 16.7

Figure 5.4 Comparison of cavern size obtained from ERT and numerical simulations for a
Cm = 2% hardwood (Z/T = 1.0).
92

The CFD simulations for softwood pulp (Cm = 3%) are similar to those made for the

hardwood pulp. The two rheology models did not make significant differences in cavern shape

although mild surface motion was predicted in the cases of Herschel-Bulkley model, which also

provided cavern volumes closer to the experimental results, with an average of 13% lower than

the ERT results (Figure 5.5). Since the fibres of softwood pulp are longer than those of hardwood

pulp, they easily tangle together, especially at low impeller speeds, but they gradually align more

substantially than shorter hardwood fibres in the direction of increasing shear, producing less

resistance to flow. This makes softwood pulp behave more like a shear-thinning fluid, which can

be further supported by the value of the power-law index (n = 0.21 for Cm = 3% softwood

whereas 0.328 for Cm = 2% hardwood with n < 1 for shear-thinning fluid and n = 1 for

Newtonian fluid). Also, the impeller speeds used in the comparison are not close to the point of

complete mixing (i.e., 1020rpm for 3% softwood pulp) and so it is unlikely that the suspension

would behave like water (Bingham model with n = 1). With favorably tangled long-fibred

structure, the 3% softwood pulp would be better described by Herschel-Bulkley model in mixing

simulation at low impeller speeds.

Figure 5.6 shows the flow field computed at the measurement location along the cross-

section of the impeller hub (Fig. 5.2) for the simulations of two pulp types using their most

comparable rheology models mentioned above. Both simulated flow fields of hardwood and

softwood pulps were found to match the flows expected for an axial-flow impeller. The impeller

discharges the suspension along the rotation axis and up to the surface for return to the suction

side of the impeller, creating a large upper circulation loop which was observed during

experiments. The smaller loop generated under the impeller is also shown in the two simulations.
93

Method ERT CFD


Rheology
Herschel-Bulkley Bingham
Model
K1 (kgsn-2/m) 135 0.001
n 0.21 1
y (Pa) 155 155

Cavern
shape
(425rpm)

Vc (dm3) 9.8 9.2 6.5

Cavern
shape
(475rpm)

Vc (dm3) 10.6 9.6 7.0

Cavern
shape
(525rpm)

Vc (dm3) 13.3 10.0 7.4

Figure 5.5 Comparison of cavern size obtained from ERT and numerical simulations for a
Cm = 3% softwood (Z/T = 0.8).
94

a)

b)

Figure 5.6 Flow fields of batch mixing for a) Cm = 2% hardwood at N = 200rpm using
Bingham plastic model and b) Cm = 3% softwood at N = 475 rpm using Herschel-bulkley
model.
95

In addition to the observed flow from the sides of the vessel, the velocity profiles

calculated from CFD simulations were compared with the ones measured using UDV at the

specified location shown in Figure 5.2 at three different heights (Figure 5.7). The velocity profile

along the measurement line (the dotted arrow lines) was measured with the velocity direction

parallel to the pulse path. The positive velocity means the velocity direction is away from the

UDV probe. Figures 5.8 and 5.9 compare the measured velocities with the ones calculated using

two rheology models for Cm = 2% hardwood at two impeller speeds. The sudden irregular ups

and downs shown on the UDV-measured velocity curves are possibly due to the interference of

bubbles introduced by air entrapment at the free suspension surface during agitation. The bubble

movement and coalescence-breakup behavior induce turbulent fluctuations to the flow field in

both direction and amplitude, and this would distort the ultrasonic signals and affect the velocity

measurement (Bouillard et al., 2001; Wang et al. 2003). Since UDV measured the velocity of

pulp fibres whereas CFD computed the fluid velocity, it would be expected that the computed

values would be slightly different from the measured values. At N = 200 rpm, neither model can

exactly predict the measured velocities but the velocity profiles obtained from the Bingham

plastic model are closer to the measurements, e.g., at position 2, the velocity profile computed

using the Bingham model follows the direction change of the measured profile (from positive to

negative) at about +0.05m from the chest centre. At N = 250rpm, the velocity profile shape of

the Bingham model is very close to the measured one. The peaks and dips of both profiles occur

almost at the same locations and so it is obvious that the Bingham model would be preferred for

hardwood pulp mixing simulation.


96

a)

Impeller

UDV probe
Line of the impeller
shaft

b)

UDV probe

Position 3

Position 2

Position 1

Figure 5.7 a) Top view and b) side view of probe locations for UDV measurements. The
arrows show the direction of the sonic emissions at three heights (12.5mm, 92.5mm and
252.5mm measured from the chest bottom).
97

a) Position 1 0.04

0.03 UDV
Herschel-Bulkley
0.02 Bingham plastic

0.01

Velocity (m/s)
0.00

-0.01

-0.02

-0.03

-0.04
-0.2 -0.1 0.0 0.1 0.2
Distance from the chest centre (m)

b) Position 2
0.05

0.04 UDV
Herschel -Bulkley
0.03 Bingham plastic

0.02
Velocity (m/s)

0.01

0.00

-0.01

-0.02

-0.03
-0.2 -0.1 0.0 0.1 0.2
Distance from the chest centre (m)
c) Position 3
0.00

-0.02
UDV
Herschel-Bulkley
Velocity (m/s)

Bingham plastic
-0.04

-0.06

-0.08

-0.2 -0.1 0.0 0.1 0.2


Distance from the chest centre (m)

Figure 5.8 Comparison of measured and computed velocity profiles at three positions for a
Cm = 2% hardwood at N = 200rpm. Positive velocity means the velocity direction is away
from the probe.
98

a) Position 1
0.10

0.08 UDV
Herschel-Bulkley
0.06 Bingham plastic

Velocity (m/s)
0.04

0.02

0.00

-0.02

-0.04
-0.2 -0.1 0.0 0.1 0.2
Distance from the chest centre (m)

b) Position 2
0.20

0.15
UDV
Herschel-Bulkley
0.10 Bingham plastic
Velocity (m/s)

0.05

0.00

-0.05

-0.10

-0.15
-0.2 -0.1 0.0 0.1 0.2
Distance from the chest centre (m)

c) Position 3 0.01

0.00

-0.01

-0.02
Velocity (m/s)

-0.03

-0.04
UDV
-0.05
Herschel-Bulkley
Bingham plastic
-0.06

-0.07

-0.2 -0.1 0.0 0.1 0.2


Distance from the chest centre (m)

Figure 5.9 Comparison of measured and computed velocity profiles at three positions for a
Cm = 2% hardwood at N = 250rpm. Positive velocity means the velocity direction is away
from the probe.
99

CFD was also used to predict the point where the suspension was in complete motion

throughout the mixing chest (i.e., where the cavern volume equaled the suspension volume). The

impeller speeds for complete mixing predicted by CFD for both pulps are about twice the

experimental one. This deviation may be due to the complex pulp suspension rheology, which is

not easily described by CFD, e.g., floc formation easing the flow of pulp suspension. Also, for

both the hardwood and softwood pulps, the cavern volumes determined using CFD and ERT

increased with impeller speed (Fig. 5.10). The increasing trend of cavern volume observed in

ERT results was predicted satisfactorily for the hardwood pulp but not for the softwood pulp,

which has CFD results increasing more steadily and slowly than those of ERT. The possible

reason for this difference is the difficulty of using CFD to describe the complex behavior of pulp

fibres in agitation, i.e., softwood pulp forms flocs more readily than hardwood pulp in actual

situation. With the same amount of fibres, the total surface area of flocs is smaller than the sum

of individual fibre surface area. Thus, the resistance to flow in softwood pulp suspension

becomes lower than the expected because flocs would move more easily than long pulp fibres. In

CFD, no simulation of floc formation was involved and so the resistance to flow would not be

changed, leading to relatively smaller cavern volume when compared with the experimental

results. In addition, during experiments, it was observed that air entrapment was significant in

agitation of softwood pulp. This would enhance the cavern growth, which cannot be easily

simulated in CFD.
100

a) 50

45 3
Suspension volume = 43.4dm
40

35

30

Vc ( dm )
3
25

20
ERT
15
CFD
10

0
100 200 300 400 500 600
N (rpm)
b)
40

3
Suspension volume = 34.7dm
35

30

25
Vc ( dm )
3

20

15
ERT
CFD
10

400 800 1200 1600 2000 2400


N (rpm)

Figure 5.10 Comparison of measured and computed cavern volumes. (a) Cm = 2%


hardwood pulp using Bingham plastic model for cavern determination (Z/T = 1.0); (b) Cm =
3% softwood pulp using Herschel-Bulkley model for cavern determination (Z/T = 0.8).

Figures 5.11 and 5.12 compare the responses simulated from CFD with those obtained

from experiments for dynamic tests. It seems that both the CFD and experimental results are

reasonably similar and increasingly close to each other with increasing impeller speed. Both

simulations and experiments show no bypassing was detected in the studied mixing situations.

Figure 5.13 shows the simulated path lines of particles released into the chest using CFD. It is
101

clear that the particles enter the region (cavern) around the impeller before exiting the chest and

this agrees with the experimental result of no channeling. Tables 5.3 and 5.4 shows the dynamic

parameters (Td – time delay and 2 – time constant for mixing zone) obtained from CFD and

experimental responses. For both pulps, Td predicted by CFD is larger than that of dynamic test,

especially for softwood pulp. The possible explanation for this is the air entrapped in the pipe

flow in actual situation. During experiments, the feed tank was open and agitated by a top-entry

mixer. Air could then be drawn into the pulp suspension and flowed with the suspension. Thus,

besides pulp suspension, there was air in the piping. This air occupancy would create “shortcut”

in the pipes and shorten the time delay. However, no such phenomenon was considered in CFD

simulations and so the CFD Td was found to be larger than the experimental one. Since air

entrapment is more significant for softwood pulp than for hardwood pulp as observed during

experiments, the deviation in Td is higher for softwood pulp.

For the 2 of 2% hardwood pulp, excluding the value at low impeller speed (i.e., 125rpm),

the average difference between simulated and experimental values is about 16%. This result is in

accord with that of batch mixing in Figure 5.10a where CFD predicted fairly well at high speeds.

For 3% softwood pulp, the difference in 2 becomes higher as the impeller speed increases. The

predicted time constants of the last two impeller speeds (i.e., 1020 and 1100rpm), which show

complete mixing in experiments, are much lower than the experimental ones. The reason is that

under CFD simulations, these two speeds are not at complete mixing, as shown in Figure 5.10.

The impeller speed calculated from CFD for complete mixing of 3% softwood pulp is over 2300

rpm and so that‟s why the predicted time constants are lower than the experimental ones. Figure

5.14 also shows the effectively mixed volume calculated from 2 for Cm = 2% hardwood and Cm

= 3% softwood. It is clear that CFD predicted better for hardwood pulp than softwood pulp and
102

the reason would possibly be the higher tendency of long-fibred softwood pulp to form flocs,

which offer less resistance to flow and this phenomenon could not be easily simulated by CFD.

This also agrees with the CFD results of complete mixing in batch operation (Figure 5.10),

showing that CFD approximated better for hardwood than softwood. In addition, the percentages

of fitting for CFD are generally higher than those for dynamic tests because unlike the

experimental outputs, the simulated responses generally do not contain any noises that would

affect the fitting process.

To conclude the study, the possible reasons for the difference between experimental

results and numerical simulations are the heterogeneous nature of pulp suspensions and the air

entrapment during experiments. To verify these explanations, this study should be repeated with

a homogeneous fluid with known rheology (possibly similar to the properties of pulp

suspensions) in a closed system. The approach using glycerin solution in the study of the flow

field in a rectangular vessel by Gomez et al. (2010) can be used and then the outcome can be

compared with the results of this study. In addition, their study showed very good agreement

between the experimental and computational results. Since the CFD modeling in this study is

similar to that of Gomez‟s study, the deviation between the modeling and experimental results is

most likely due to the special rheology of pulp suspension, which is not easy to be formulated

and implemented in the CFD model.


103

Table 5.3 Comparison of results from CFD and dynamic tests for Cm = 2% hardwood (Q =
14L/min., Z/T = 0.8)
Td (s) 2 (s) Percentage of fitting (%)
N (rpm) CFD Dynamic CFD Dynamic CFD Dynamic test
test test (Eqn (5.8)) (Eqn. (4.1))
125 65.0 58.8 124 47 94 89
250 55.0 40.0 116 106 95 85
380 49.0 34.8 178 155 90 91
460 41.6 38.7 177 142 89 91

Table 5.4 Comparison of results from CFD and dynamic tests for Cm = 3% softwood (Q =
14L/min., Z/T = 0.8)
Td (s) 2 (s) Percentage of fitting (%)
N (rpm) CFD Dynamic CFD Dynamic CFD Dynamic test
test test (Eqn.(5.8)) (Eqn.(4.1))
425 60.0 41.4 70 69 88 86
525 56.0 40.0 72 97 85 82
1020 50.0 36.2 68 160 90 91
1100 55.0 40.3 75 156 83 80
104

12
12.0

11.5
11

11.0

Conductivity (mS)
Conductivity (mS)

10.5
10

10.0

9 9.5

9.0 CFD
CFD Experimenta
Experimental
8 8.5
0 240 480 720 960 1200 1440 1680 1920 2160 2400 0 240 480 720 960 1200 1440 1680 1920 2160 2400
Flow time (s) Flow time (s)

125rpm 250rpm

7.5 10

7.0

9
6.5
Conductivity (mS)

Conductivity (mS)

6.0
8
5.5

5.0
7

CFD CFD
4.5
Experimental Experimental

4.0 6
0 240 480 720 960 1200 1440 1680 1920 2160 2400 0 240 480 720 960 1200 1440 1680 1920 2160 2400
Flow time (s) Flow time (s)

380rpm 460rpm

Figure 5.11 Experimental and computed dynamic responses for Cm = 2% hardwood pulp.
105

3.5 5.0

3.0 4.5

2.5 4.0
Conductivity (mS)

Conductivity (mS)
2.0 3.5

1.5 3.0

1.0 2.5

CFD CFD
0.5 Experimental 2.0 Experimental

0 240 480 720 960 1200 1440 1680 1920 2160 2400 0 240 480 720 960 1200 1440 1680 1920 2160 2400
Flow time (s) Flow time (s)

425rpm 525rpm

5.5

2.5
5.0

2.0
4.5
Conductivity (mS)

Conductivity (mS)

1.5
4.0

1.0
3.5

0.5 3.0
CFD CFD
Experimental Experimental
0.0 2.5
0 240 480 720 960 1200 1440 1680 1920 2160 2400 0 240 480 720 960 1200 1440 1680 1920 2160 2400
Flow time (s) Flow time (s)

1020rpm 1100rpm

Figure 5.12 Experimental and computed dynamic responses for Cm = 3% softwood pulp.
106

Inlet

Outlet

Figure 5.13 Path lines of particles (each particle is represented by a different color)
simulated by CFD using Cm = 2% hardwood pulp at N = 250rpm and Q = 14L/min.
107

a) 45

40

35

30

25

Vmix ( dm )
3
20

15

10
Dynamic test
CFD
5

0
100 200 300 400 500
N (rpm)

b) 40

35

30

25
Vmix ( dm )
3

20

15

10
Dynamic test
CFD
5

0
400 600 800 1000 1200
N (rpm)

Figure 5.14 Experimental and computed determination of mixed volume for (a) Cm = 2%
hardwood pulp and (b) Cm = 3% softwood pulp in continuous operation.
108

5.5 Summary

A CFD model of a lab-scale cylindrical mixing chest was developed using FLUENT 6.2. To

verify the ability of the CFD model for flow estimation in agitated pulp stock chests, a number of

numerical simulations were carried out to compare with the experimental results in terms of

cavern volumes and dynamic model parameters. Although the CFD model cannot precisely

depict the cavern shape and the surface motion, it can predict the increasing trend of cavern size

with impeller speed. Based on the cavern comparison, it seemed that Bingham plastic model

would be preferred for the hardwood pulp at impeller speeds close to complete mixing whereas

Herschel-Bulkley model predicted better for the softwood pulp at low impeller speeds. The

suitability of the Bingham model for the hardwood pulp was further supported by the UDV

measurement. The reason why the CFD model cannot fully predict the actual mixing situations is

that the actual behavior of pulp fibres, like floc formation of softwood pulp in agitation, and the

air entrapment during agitation cannot be easily described by CFD. The best way to justify the

explanation is to repeat the mixing experiments using a homogeneous fluid with no known

rheology similar to that of pulp suspensions in a closed system and then compare the outcome

with the results of this study. Actually, a major difficulty in using CFD for simulation of pulp

mixing is the lack of good description of rheology in CFD code. However, the agreement of

simulated results of particle path lines with the experiment result of no passing suggests that the

CFD model can provide valuable information about the flow patterns in agitated pulp chests.
109

6 OVERALL CONCLUSIONS AND RECOMMENDATIONS FOR


FUTURE WORK

6.1 Overall conclusions

Since pulp suspensions possess a complex rheology, it is not easy to achieve an effective

agitation in cylindrical chests. Caverns form when the chests are not completely agitated and this

may lead to bypassing and dead zone formation. In order to avoid undesirable flows affecting

chest performance, a study of pulp mixing in batch and continuous operations was carried out.

Electrical resistance tomography (ERT) and Ultrasonic Doppler velocimetry (UDV) were used to

study the cavern shape and size in batch operation. Both methods gave comparable results but

ERT was preferred due to less time required to acquire data. The cavern shape was best

described as a truncated right-circular cylinder and its volume was found as expected to increase

with impeller speed but the increase was non-uniform due to interaction between the cavern and

the vessel walls.

There are models for estimating cavern volumes in fluids with yield stresses but without

including the interaction between the cavern and the vessel walls. Like the situations in the study,

they do not describe the experimental results well. Thus, based on the observed cavern shape in

ERT imaging, a model was developed to estimate the cavern volume by balancing the force

provided by the impeller with the resistive force at the cavern boundaries. Although the proposed

model could not accurately calculate the cavern volume, it could predict the increasing trend in

cavern volume with impeller speed.

In continuous operation, dynamic tests were used to characterize the mixing quality in a

cylindrical chest. A model including a mixed region (cavern) and the possibility of bypassing
110

was shown to precisely describe the mixing dynamics of the chest because the time delay

estimated using the model agreed closely with the calculated result based on flow and system

geometry. The measured bypassing was found to be insignificant, indicating that the flow

configuration used in the study (with the exit within the cavern, close to and below the impeller)

is effective for avoiding feed bypassing. Also, ERT further confirmed the presence of cavern

when the chest was not fully agitated in continuous operation with feed and the cavern volume

increased with impeller speed. In addition, the power for complete agitation in a cylindrical chest

was similar to that in a rectangular chest of the same volume.

Besides the experimental study of pulp mixing in a cylindrical chest, the ability of

computational fluid dynamics (CFD) to estimate the flow in agitated pulp chests was examined.

Numerical simulations were carried out to compare with the experimental results. Owing to the

complex properties of pulp suspension, it is not easy to describe its unique behavior like floc

formation in agitated chests using CFD. The application of both Bingham and Herschel-Bulkley

models could not exactly portray the cavern, but the increasing trend of cavern volume with the

impeller speed was still predicted. The CFD simulations also picked up the features of the flow

field measured by UDV at the high impeller speed. The calculated path lines of the simulated

particle feed matched with the experimental results of no bypassing. Therefore, in spite of its

limitation to model pulp suspensions, the CFD model can still provide useful flow information in

pulp agitated chests.

The outcome of this research can aid the design of industrial cylindrical chests. The

cavern model can estimate the mixing volume in proposed agitated pulp chests, aiding the sizing

of chests. The dynamic test study provides insight about the positions of in/out piping of new

chests, e.g., the outlet should be placed close to the cavern, to avoid non-ideal flows which affect
111

the chest performance. Besides new chests, this test method can also be used to estimate the

performance of existing pulp chests, trying to improve or enhance their mixing efficiency. In

addition, the flow configuration of the proposed chest design can be checked using CFD. Saving

the need of building a lab-scale chest and doing experiments, the CFD model can provide

information about the flow pattern in the proposed chest, helping chest design and optimization.

Finally, in general, the contributions of this research are:

 The experimental results about the cavern geometry in cylindrical vessels can provide

insights for the design of chest geometry to facilitate pulp mixing, e.g., installation of

fillets is needed to minimize dead zone. The proposed cavern model could be used to

estimate the mixing volume in existing and proposed pulp chests to see whether they meet

the requirement.

 The results of dynamic tests show that the flow configuration with the exit installed near

the cavern can avoid bypassing. This information confirms the importance of exit location

in an agitated cylindrical chest and helps the design of efficient-mixing chests.

 The results of the CFD model can provide valuable information about the flow pattern,

e.g., the presence of bypassing, inside the existing and proposed chests. This also helps

the design of flow configuration of agitated pulp chests, like the locations of the impeller,

inlet and outlet.


112

6.2 Recommendation for future work

The results of this study provided some insights for future consideration as follows:

 The presence of dead zone under incomplete mixing implies the possibility of disturbance

to normal operation (e.g., the oozing of undesirable materials from the dead zone into the

cavern). The installation of fillets can avoid this hazard and its effect on the cavern

formation should be examined.

 Based on the experimental results of ERT and UDV, a model for determining the cavern

volume in a cylindrical chest was developed. Similar modeling should be carried out in

rectangular chests because they are also widely used in the industry. This model can be

modified to extend its application on rectangular chests.

 For the agitation of softwood pulp, entrapment of air into the suspension is significant at

high impeller speeds and this will affect the cavern size. Thus, the effect of air entrapment

on the cavern formation should be studied.

 Owing to the poor knowledge of pulp suspension rheology, it was proposed that the

heterogeneous structure of pulp suspensions could cause the discrepancies between the

experimental results and CFD simulations. To validate this proposition, a study of the

same mixing system using a homogeneous fluid with known rheology should be carried

out.

 In this study, the simulated results of CFD in cavern size are not in complete agreement

with the experimental results and one of the possible reasons is the incomplete quantitative

description of pulp suspension rheology, e.g., the interactions of flocs. A better numerical

description of pulp suspension rheology, possibly in two stages (MACRO and MICRO),

should be developed for CFD simuation.


113

 Besides cylindrical chests, reduced bottom chests are frequently used in pulp and paper

industry. They do not always perform satisfactorily because of the undesirable flows such

as channeling induced by the cavern. Thus, cavern formation and dynamic behavior of

these chests should be investigated to minimize these non-ideal flows.


114

BIBLIOGRAPHY
Amanullah, A., S.A. Hjorth and A.W. Nienow, “A New Mathematical Model to Predict Cavern
Diameters in Highly Shear Thinning, Power Law Liquids Using Axial Flow Impellers”,
Chemical Engineering Science, 53(3), 455-469 (1998).

Ankamma Rao, D. and P. Sivashanmugam, “Experimental and CFD Simulation Studies on


Power Consumption in Mixing using Energy Saving Turbine Agitator”, Journal of Industrial and
Engineering Chemistry, 16, 157-161 (2010).

Armenante, P.M., C. Luo, C. Chou, I. Fort and J. Medek, “Velocity Profiles in a Closed,
Unbaffled Vessel: Comparison between Experimental LDV Data and Numerical CFD
Predictions”, Chemical Engineering Science, 52(20), 3483-3492 (1997).

Bakker, A. and J.B. Fasano, “A Computational Study of the Flow Pattern in an Industrial Paper
Stock Chest with a Side-entering Impeller”, AIChE Symposium Series, 89(293), 118-124 (1993).

Bakker, C.W., C.J. Meyer and D.A. Deglon, “The Development of a Cavern Model for
Mechanical Flotation Cells”, Minerals Engineering, in press (2010).

Barnes, H.A., J.F., Hutton, K. Walters, An Introduction to Rheology, Elsevier, Amsterdam


(1989).

Bennington, C.P.J., R.J. Kerekes and J.R. Grace, “The Yield Stress of Fibre Suspensions”, The
Canadian Journal of Chemical Engineering, 68(10), 748-757 (1990).

Bennington, C.P.J. and R.J. Kerekes, “Power Requirements for Pulp Suspension fluidization”,
TAPPI Journal, 79(2), 253-258 (1996).

Bhattacharya, S. and S.M. Kresta, “CFD Simulations of Three-dimensional Wall Jets in a Stirred
Tank”, The Canadian Journal of Chemical Engineering, 80, 1-15 (2002).

Bhattacharya, S., C. Gomez, A. Soltanzadeh, F. Taghipour, C.P.J. Bennington and G.A. Dumont,
“Computational Modelling of Industrial Pulp Stock Chests”, The Canadian Journal of Chemical
Engineering, 88, 295-305 (2010).

Bhole, M., C. Ford and C.P.J. Bennington, “Characterization of Axial Flow Impellers in Pulp
Fibre Suspensions”, Chemical Engineering Research and Design, 87, 648-653 (2009).

Bialkowski, W.L., “Newsprint Variability and Its Impact on Competitive Position”, Pulp and
Paper Canada, 93(11), T299-T306 (1992).

Bird, R.B., W.E. Stewart and E.N. Lightfoot, Transport Phenomena, John Wiley & Sons, Inc.
(2002).
115

Blasinski, H. and E. Rzyski, “Mixing of non-Newtonian Liquids. Power Consumption for


Fibrous Suspensions”, Inz. Chem., 2(1), 169-182 (1972).

Bolton, G.T., C.W. Hooper, R. Mann and E.H. Stitt, “Flow Distribution and Velocity
Measurement in a Radial Flow Fixed Bed Reactor Using Electrical Resistance Tomography”,
Chemical Engineering Science, 59, 1989-1997 (2004).

Bouillard, J., B. Alban, P. Jacques and C. Xuereb, “Liquid Flow Velocity Measurements in
Stirred Tanks by Ultra-sound Doppler Velocimetry”, Chemical Engineering Science, 56, 747-
754 (2001).

Chhabra, R.P. and J.F. Richardson, Non-Newtonian Flow in the Process Industries, Butterworth-
Heinemann (1999).

Devries, T.C. and M.D. Doyle, “Dilution Zone Agitation Critical in Blowtanks, High-density
Towers”, Pulp and Paper, 87-95 (1995).

Dickin, F. and M. Wang, “Electrical Resistance Tomography for Process Applications”, Meas.
Sci. Technol., 7, 247-260 (1996).

Dietemann, P. and M. Rueff, “A Study of Fibre Suspension Flow by means of Doppler


Ultrasound Velocimetry and Image Analysis”, Annual Meeting - Technical Section, Pulp and
Paper Technical Association of Canada (PAPTAC), Preprints, v. A, 225-230 (2004).

Ein-Mozaffari, F., “Macroscale Mixing and Dynamic Behavior of Agitated Pulp Stock Chests”,
PhD Thesis, University of British Columbia, Vancouver, Canada (2002).

Ein-Mozaffari, F., L.C. Kammer, G.A. Dumont and C.P.J. Bennington, “Dynamic Modeling of
Agitated Pulp Stock Chests”, TAPPI Journal, 2(9), 13-17 (2003).

Ein-Mozaffari, F., C.P.J. Bennington and G.A. Dumont, “Suspension Yield Stress and the
Dynamic Response of Agitated Pulp Chests”, Chemical Engineering Science, 60, 2399-2408
(2005).

Ein-Mozaffari, F., C.P.J. Bennington, G.A. Dumont and D. Buckingham, “Using Ultrasonic
Doppler Velocimetry to Measure Flow Velocity in Pulp Suspension Mixing”, 12th European
Conference on Mixing, Bologna, 27-30 June 2006.

Ein-Mozaffari, F., C.P.J. Bennington, G.A. Dumont and D. Buckingham, “Measuring Flow
Velocity in Pulp Suspension Mixing Using Ultrasonic Doppler Velocimetry”, Chemical
Engineering Research and Design, Trans IChemE, Part A, 85(A5), 591-597 (2007).

Ein-Mozaffari and S.R. Upreti, “Using Ultrasonic Doppler Velocimetry and CFD Modeling to
Investigate the Mixing of non-Newtonian Fluids Possessing Yield Stress”, Chemical Engineering
Research and Design, 87, 515-523 (2009).
116

Elson, T.P., D.J. Cheesman, A.W. Nienow, “X-ray Studies of Cavern Sizes and Mixing
Performance with Fluids Possessing a Yield Stress”, Chemical Engineering Science, 41(10),
2555-2562 (1986).

Elson, T.P., “The Growth of Caverns Formed Around Rotating Impellers During the Mixing of a
Yield Stress Fluid”, Chemical Engineering Communication, 96, 303-319 (1990).

Ferziger, J.H. and M. Peric, Computational Methods for Fluid Dynamics, Springer-Verlag Berlin
Heidelberg (1999).

Ford, C., “CFD Simulation of Mixing Dynamics in Agitated Pulp Stock Chests”, MASc Thesis,
The University of British Columbia, Vancouver, Canada (2004).

Ford, C., F. Ein-Mozaffari, C.P.J. Bennington and F. Taghipour, “Simulation of Mixing


Dynamics in Agitated Pulp Stock Chests using CFD”, AIChE Journal, 52(10), 3562-3569 (2006).

Gibbon, J.D. and D. Attwood, “Prediction of Power Requirements in the Agitation of Fibre
Suspensions”, Trans. Inst. Chem. Engrs., 40, 75-82 (1962).

Gomez, C., C.P.J. Bennington and F. Taghipour, “Investigation of the Flow Field in a
Rectangular Vessel Equipped with a Sider-entering Agitator”, Journal of Fluids Engineering,
132(5), 051106-1 – 051106-13 (2010).

Gomez, C., B. Derakhshandeh, S.G. Hatzikiriakos and C.P.J. Bennington, “Carbopol as a model
fluid for studying mixing of pulp fibre suspensions”, Chemical Engineering Science, 65, 1288-
1295 (2010).

Gullichsen, J. and E. Harkonen, (1981). “Medium Consistency Technology”, TAPPI Journal,


64(6), 69-72 (1981).

Hossseini, S., D. Patel, F. Ein-Mozaffari and M. Mehrvar, “Study of Solid-Liquid Mixing in


Agitated Tanks through Electrical Resistance Tomography”, Chemical Engineering Science, 65,
1374-1383 (2010).

Holden, P.J., M. Wang and R. Mann, “Imaging Stirred-vessel Macromixing Using Electrical
Resistance Tomography”, AIChE Journal, 44(4), 780-790 (1998).

Hoyle, B.S., H. McCann and D.M. Scott, “Process Tomography”, Process Imaging for
Automatic Control, Taylor & Francis Group, LLC (2005).

Hui, L.K., C.P.J. Bennington and G.A. Dumont, “Cavern Formation in Pulp Suspensions Using
Side-entering Axial-flow Impellers”, Chemical Engineering Science, 64, 509-519 (2009).

Hui, L.K., C.P.J. Bennington and G.A. Dumont, “Mixing Dynamics in Cylindrical Pulp Stock
Chests”, 2010 EXFOR & PAPTAC Annual Meeting, Feb. 2-4, 2010, Montreal, QC (2010).
117

Industrial Tomography Systems, ITS System 2000 v6 Manual (2007).

Kasat, G.R., A.R. Khopkar, V.V. Ranade and A.B. Pandit, “CFD Simulation of Liquid-phase
Mixing in Solid-liquid Stirred Reactor”, Chemical Engineering Science, 63, 3877-3885 (2010).

Kim, S., A.N. Nkaya and T. Dyakowski, “Measurement of Mixing of Two Miscible Liquids in a
Stirred Vessel with Electrical Resistance Tomography”, International Communications in Heat
and Mass Transfer, 33, 1088-1095 (2006).

Li, M., G. White, D. Wilkinson and K.J. Roberts, “Scale Up Study of Retreat Curve Impeller
Stirred Tanks using LDA Measurements and CFD Simulation”, Chemical Engineering Journal,
108, 81-90 (2005).

Ma, Y., Z. Zheng, L. Xu, X. Liu and Y. Wu, “Application of Electrical Resistance Tomography
System to Monitor Gas/Liquid Two-phase Flow in a Horizontal Pipe”, Flow Measurement and
Instrumentation, 12, 259-265 (2001).

Macosko, C.W., Rheology, Principles, Measurement and Applications, VCH Publisher, New
York (1994).

Malmivuo, J. and R. Plonsey, Bioelectromagnetism, Principles and Applications of Bioelectric


and Biomagnetic Fields, Oxford University Press (1995).

Mann, R., F.J. Dickin, M. Wang, T. Dyakowski, R.A. Williams, R.B. Edwards, A.E. Forrest and
P.J. Holden, “Application of Electrical Resistance Tomography to Interrogate Mixing Processes
at Plant Scale”, Chemical Engineering Science, 52, 2087-2097 (1997).

Martin, A., “To Find the Area Common to Two Intersecting Circles”, The Analyst, 1(2), 33-34
(1874).

Oldshue, J.Y., Fluid Mixing Technology, McGraw-Hill Publications Co., New York (1983).

Paul, E.L., V.A. Atiemo-Obeng and S.M. Kresta, Handbook of Industrial Mixing: Science and
Practice, John Wiley & Sons, Inc. (2004).

Reed, C.S., “Selecting the Right Equipment for Agitation and Blending, Part 2”, TAPPI Journal,
78(7), 241-244 (1995).

Saeed, S., F. Ein-Mozaffari and S. R. Upreti, “Using Computational Fluid Dynamics Modelling
and Ultrasonic Doppler Velocimetry to Study Pulp Suspension Mixing”, Industrial &
Engineering Chemistry Research, 46, 2172-2179 (2007).

Shekarriz, A. and D.M. Sheen, “Slurry Pipe Flow Measurements using Tomographic Ultrasonic
Velocimetry and Densitometry”, Proceedings of FEDSM, June 21-25, 1998, Washington, D.C.,
USA.
118

Silvester, R.S., Mixing of Non-Newtonian Media: A Technical Review, BHRA Fluid


Engineering, England (1985).

Solomon, J., T.P. Elson and A.W. Nienow, “Cavern Sizes in Agitated Fluids with a Yield Stress”,
Chemical Engineering Communication, 11, 143-164 (1981).

Soltanzadeh, A., S. Bhattacharya, G.A. Dumont and C.P.J. Bennington, “Estimation of


Residence Time and Channeling in Agitated Pulp Chests”, Nordic Pulp Pap. Res. J., 24(1), 66-67
(2009).

Stephenson, D.R., M. Cooke, A. Kowalski and T.A. York, “Determining Jet Mixing
Characteristics using Electrical Resistance Tomography”, Flow Measurement and
Instrumentation, 18, 204-210 (2007).

Takeda, Y., “Development of an Ultrasound Velocity Profile Monitor”, Nuclear Engineering and
Design, 126, 277-284 (1991).

Takeda, Y., “Velocity Profile Measurement by Ultrasonic Doppler Method”, Experimental


Thermal and Fluid Science, 10, 444-453 (1995).

Takeda Y., “Ultrasonic Doppler Method for Velocity Profile Measurement in Fluid Dynamics
and Fluid Engineering”, Experiments in Fluids, 26, 177-178 (1999).

TAPPI, Generalized Method for Determining the Pipe Friction Loss of Flowing Pulp
Suspensions, TIS 0410-14 (1998).

Wahren, D., “Fibre Network Structures in Paper Making Operations”, Conference Paper Science
and Technology, The Cutting Edge, Institute of Paper Chemistry, Appleton, WI, 112-132 (1980).

Wang, L., K.L. McCarthy and M.J. McCarthy, “Effect of Temperature Gradient on Ultrasonic
Doppler Velocimetry Measurement During Pipe Flow”, Food Research International, 37, 633-
642 (2004).

Wang, M., A. Dorward, D. Vlaev and R. Mann, “Measurements of Gas-liquid Mixing in a


Stirred Vessel using Electrical Resistance Tomography (ERT)”, Chemical Engineering Journal,
77(1-2), 93-98 (2000).

Wang, M., “Impedance Mapping of Particulate Multiphase Flows”, Flow Measurement and
Instrumentation, 16, 183-189 (2005).

Wang, T., J. Wang, F. Ren and Y. Jin, “Application of Doppler Ultrasonic Velocimetry in
Multiphase Flow”, Chem. Eng. J., 92, 111-122 (2003).

Wichterle, K. and O. Wein, “Threshold of Mixing of non-Newtonian Liquids”, International


Chemical Engineering, 21(1), 116-119 (1981).
119

Wilkens, R.J., C. Henry and L.E. Gates, “How to Scale-Up Mixing Processes in Non-Newtonian
Fluids”, Chem. Eng. Prog., 99(5), 44-52 (2003).

Wilkens, R.J., J.D. Miller, J.R. Plummer, D.C. Dietz and K.J. Myers, “New Techniques for
Measuring and Modeling Cavern Dimensions in a Bingham Plastic Fluid”, Chemical
Engineering Science 60, 5269-5275 (2005).

William, R.A., X. Jia and S.L. McKee, “Development of Slurry Mixing Models using Resistance
Tomography”, Powder Technology, 87, 21-27 (1996).

Wilstrom, T. and Rasmuson A., “The Agitation of Pulp Suspensions with a Jet Nozzle Agitator”,
Nordic Pulp Pap. Res. J., 13(2), 88-94 (1998a).

Wikstrom, T. and Rasmuson, A., “Yield Stress of Pulp Suspensions. The Influence of Fibre
Properties and Processing Conditions”, Nord. Pulp Pap. Res. J., 13(3), 243-246 (1998b).

Xu, H., “Measurement of Fiber Suspension Flow and Forming Jet Velocity Profile by Pulsed
Ultrasonic Doppler Velocimetry”, PhD thesis (2003).

Xu, H. and Aidun, C.K., “Characteristics of Fiber Suspension Flow in a Rectangular Channel”,
International Journal of Multiphase Flow, 31, 318-336 (2005).

Yackel, D.C., Pulp and Paper Agitation: The History, Mechanics, and Process, TAPPI Press,
Atlanta (1990).

York, T.A., “Status of Electrical Tomography in Industrial Applications”, Process Imaging for
Automatic Control, Proceedings of SPIE 4188, 175-190 (2001).

Zlokarnik, M., Scale-up in Chemical Engineering, Wiley-VCH (2006).


120

APPENDIX A: Details of ERT components

A. Sensor system

The size and the material of the electrodes are both important in producing and sensitively

measuring the electrical field distribution (Mann et al., 1997). Wang et al. (1995) found that the

smaller the size of electrodes, the higher the sensitivity of the voltage measurement electrodes

and the higher common voltage at current driven electrodes, when measurement and current

excitation use the same electrode system. However, the higher common voltage will produce a

poor signal-to-noise ratio. Also, current-injecting electrodes should have large surface areas in

order to generate an even current density (Dickin and Wang, 1996). Thus, the electrode size

should not be too large or too small. In addition, the electrodes must be made of a material more

electrically conductive than the process fluid (Ricard, 2005). Metallic electrodes are thus often

used for process applications, e.g., stainless steel, silver, gold or platinum.

B. Data acquisition system (DAS)

In order to track the small changes of resistivity in real-time, the data collection has to be quick

and accurate, enabling the reconstruction algorithm to provide a precise indication of the true

resistivity distribution. The DAS is composed of a signal source, an electrode multiplexer array,

voltmeters, signal demodulators and a system controller (Mann et al., 1997). This complexity is

required because of the low amplitude of measurements at the boundary, the small responses of

dynamic change, the large number of electrode channel operations, the high common voltages

and the large stray capacitance of coaxial cable. The signal source consists of a master oscillator

and a voltage-to-current converter (Plaskowski et al., 1995). The oscillator, serving as a

frequency and amplitude reference for all of the current sources channels and as a switching
121

function for the demodulator stage, generates a harmonically pure sine-shaped waveform signal

in order to “probe” the material under investigation. The sine-wave voltage output from the

oscillator is fed into a voltage-to-current converter. Current is used in preference to voltage as

the electrical “probe” due to the variation of contact impedance between the electrode and the

fluid inside the vessel. The multiplexers are used to “share” the current source and voltage

measurement stages between any numbers of electrodes. They must exhibit a number of

properties: low on-resistance, fast switching speed, low inter-switch cross talk and low power

consumption. The signal demodulators are used to “decode” the voltage signal and to optimize

the signal-to-noise ratio by recovering the amplitude attenuation and phase shift of the sine wave

signal as a result of passing through a resistive medium.

C. Image reconstruction system

The image reconstruction algorithm can be thought of simply as a series of procedures performed

repeatedly on digitized measurement data to determine the distribution of regions of different

resistivities (e.g., component concentrations) within the process vessel. There are two different

reconstruction algorithms: “qualitative” and “quantitative”. The qualitative algorithm produces

images depicting a change in resistivity relative to an initially acquired set of “reference” data

and the quantitative one creates images depicting values of resistivity or conductivity for each

pixel (Dickin and Wang, 1996).

The technique used in the qualitative algorithm is referred as backprojection between

equipotential lines. The potential difference, calculated by the forward solver, between two

equipotential lines on the boundary is back-projected to a resistivity value in the area enclosed by

the two lines for all possible injection/measurement combinations. The main advantage of this
122

algorithm is that it can be performed in a single step using a pre-calculated pixel sensitivity

matrix and the image is simply reconstructed via a matrix/vector multiplication.

The quantitative algorithm is an iterative Newton-Raphson-based algorithm specifically

developed for nonlinear problems. It is intended to quantify the variation of the conductivity in

the region of interest during the process. The reconstruction process is initiated when a set of

resistivities for the region of interest is fed into the forward problem solver. When the least-

squares error between the calculated boundary voltages and the data acquisition voltages is less

than the pre-defined one, the reconstruction process is halted and the final updated set of

resistivities will be the solution.


123

References

Dickin, F. and M. Wang, “Electrical Resistance Tomography for Process Applications”, Meas.
Sci. Technol., 7, 247-260 (1996).

Mann, R., F.J. Dickin, M. Wang, T. Dyakowski, R.A. Williams, R.B. Edwards, A.E. Forrest and
P.J. Holden, “Application of Electrical Resistance Tomography to Interrogate Mixing Processes
at Plant Scale”, Chemical Engineering Science, 52, 2087-2097 (1997).

Plaskowski, A., M.S. Beck, R. Thorn and T. Dyakowski, Imaging Industrial Flows, Applications
of Electrical Process Tomography, Institute of Physics Publishing (1995).

Ricard, F., C. Brechtelsbauer, Y. Xu, C. Lawrence and D. Thompson, “Development of an


Electrical Resistance Tomography Reactor for Pharmaceutical Process”, The Canadian Journal
of chemical Engineering, 83(2), 11-18 (2005).

Wang, M., F.J. Dickin and R.A. Williams, “The Grouped-node Technique as a Means of
Handling Large Electrode Surfaces in Electrical Impedance Tomography”, Physiological
Measurement, 16, 219-226 (1995).
124

APPENDIX B: MATLAB program for dynamic model parameter


estimation
Program “procmodeliden.m”; this program is used to estimate the dynamic model
parameters from the input-output data in M-file format.

clc
clear
for i=1:1
filen=strcat('lab','1')
load(filen);
Ts=ts;
Tres=3;
datac=iddata(cout',cin',Ts)
% datac=resample(datac,1,Tres)
% datac=datac(10:end);
datac.int='foh';
% ze =datac(1:floor(length(datac.u)/2));
% defining data set
ze=datac(1:floor(length(datac.u)/2));
%removing mean
ze=detrend(ze,'constant')
advice(datac)
% identification command
m =
pem(ze,'P1D','kp',{'max',2},'kp',{'min',0.5},'Td',{'max',100},'Td',{'min',2},
'Tz',{'min',0},'dist','arma2')
m = pem(ze,m)
present (m)
% zv=datac(floor(length(datac.u)/2):length(datac.u));
zv=datac(floor(length(datac.u)/2):length(datac.u));
zv=detrend(zv,'constant');
Kp(i)=m.kp.value;
Td(i)=m.td.value%*Tres*Ts;
Tz(i)=m.tz.value%*Tres*Ts;
Tp1(i)=m.tp1.value%*Tres*Ts;
Tp2(i)=m.tp2.value%*Tres*Ts;
tu1(i)=min(Tp1(i),Tp2(i));
tu2(i)=max(Tp1(i),Tp2(i));
f(i)=(Tz(i)-tu1(i))/(tu2(i)-tu1(i));
figure(i)
compare(zv,m);
resid(zv,m);
advice(m);
% figure(i+1)
[yo,fit(i)]=compare(zv,m);
ou=yo{1,1};
subplot(3,1,1)
plot(ou.sa,ou.y,'b',zv.sa,zv.y,'k')
title(['exp no' int2str(i) ': ''Td=' num2str(Td(i)) ', ' 'K='
num2str(Kp(i)) ', ' 'f=' num2str(f(i)) ',' 'Tz=' num2str(Tz(i)) ', ' 'Tp1='
num2str(Tp1(i)) ', ' 'Tp2=' num2str(Tp2(i)) ', ' 'Ts=' num2str(Ts)])
legend('m out',['real ' '%Fit= ' num2str(fit(i))])
125

subplot(3,1,2)
plot(ze.sa,ze.u,'b.',ze.sa,ze.y,'k')
title(['exp no' int2str(i) ': ' 'Estimation Data'])
legend('input data','output data')
subplot(3,1,3)
plot(zv.sa,zv.u,'b.',zv.sa,zv.y,'k')
title(['exp no' int2str(i) ': ' 'Validation Data'])
legend('input data','output data')

end
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'A1','
P2DZ,Arma1');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'A2',f
it');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'B1','
Kp');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'B2',K
p');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'C1','
Td');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'C2',T
d');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'D1','
Tz');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'D2',T
z');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'E1','
Tp1');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'E2',T
p1');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'F1','
Tp2');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'F2',T
p2');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'G1','
tu1');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'G2',t
u1');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'H1','
tu2');
126

%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'H2',t
u2');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'I1','
f');
%
write2excel('C:\MATLAB7\work\Ali\Identifyingonedelaymodel\fitresult',0,'I2',f
');

Program “write2excel.m”

function write2excel(fileloc,promptforsave,varargin)
% write2excel(fileloc,promtforsave,range1,data1,range2,data2,...)
%
%
% Uses ActiveX commands to write data_n into range_n in an
existing Excel
% spreadsheet. Inputs (excluding fileloc and promptforsave) must
be paired.
% As of 10/04 update, you may provide the target range (upper
left
% cell to lower right cell) OR just the upper right cell. If
% the range is specified, the function will verify that the
% corresponding data block is the correct size, and give an
% error if not. (This may be useful for error checking, for
instance.)
% If only the upper left cell is provided, write2excel will
% compute the target range.
%
% (Please use caution, as you can now overwrite data pretty
easily.)
%
% Additionally, you may now specify cells by address (eg., 'H3')
OR row, column
% (eg, '[3,8]').
%
% FILELOC: Enter a string representing the location of an Excel
file.
% Example: 'c:\brett\my archives\test1.xls'
% PROMPTFORSAVE: binary variable. 1 (DEFAULT) = Prompt before saving
% 0 = No prompt required
% RANGE SPECIFIER(S): Enter the range(s) to read. You may use Excel
% cell-references, as in:
% 'B1:P5'
% 'B1:B1' (or simply 'B1')
% OR, alternatively, you may specify the row, column values, as
% in:
% '[8,3]:[12,4]' (to write from row 8, column
% 3 to row 12, column 4);
% '[8,3]' to write from row 8, column 3
% TO WHATEVER RANGE IS REQUIRED FOR THE
DATA BLOCK.
%
% DATA: NOTE: To enter multiple strings, use cell arrays. Size
compatibility is verified
127

% if a cell range is given, and is not if only the starting


% position is specified.
%
% EXAMPLES: write2excel('c:\brett\my archives\test1.xls', 1,
'C1:E3',magic(3));
% write2excel('c:\brett\my archives\test1.xls', 0,
'C1:E3',{'string1','string 2', 'string3'});
% write2excel('c:\brett\my archives\test1.xls', 0,
'[3,8]',magic(4));
%
% Written by Brett Shoelson, Ph.D.
% Last update: 1/04.
% 10/04: Allow "dynamic specification" of cell ranges, and
% allow specification of cells by row, column format.
%
% SEE ALSO: readfromexcel

if nargin < 4
msgstr = sprintf('At a minimum, you must specify three input
arguments.\nThe first is a string indicating the location of the excel
file,\nthe second is a range to be written, and the third contains the data
to write.');
error(msgstr);
elseif ~iseven(nargin-2)
msgstr = sprintf('Please enter input variables in pairs...\n''write
range'',data,''write range'',data')
error(msgstr)
end
tmp = varargin;
sheetchanges = [];counter = 1;
for ii = 1:length(tmp)
if ischar(tmp{ii}) & (strcmp(tmp{ii},'sheet') |
strcmp(tmp{ii},'sheetname'))
sheetchanges(counter) = ii;
counter = counter + 1;
end
end
if ~isempty(sheetchanges)
[sheetnames{1:length(sheetchanges)}] = deal(varargin{sheetchanges+1});
end

[pathstr,name,ext] = fileparts(fileloc);
if isempty(ext)
fileloc = [fileloc,'.xls'];
end
if isempty(pathstr)
fileloc = which(fileloc,'-all');
if size(fileloc,1) ~= 1
error('File was either not located, or multiple locations were found.
Please reissue readfromexcel command, providing absolute path to the file of
interest.');
end
end

% Ensure that range sizes and data are size-matched


for ii = 1:2:nargin-2
128

if ismember(ii,sheetchanges) | ismember(ii,sheetchanges + 1)
continue
end
% How are cells specified?
if any(ismember(double(varargin{ii}),[65:90,97:122]))
addrtype = 'letternumber';
else
addrtype = 'rowcol';
end
% Is range provided, or should it be auto-calculated?
autorange = isempty(findstr(varargin{ii},':'));
switch addrtype
case 'letternumber'
if autorange
r1{ii} = varargin{ii};
[rx1,cx1] = an2nn(r1{ii});
rx2 = rx1 + size(varargin{ii+1},1)-1;
cx2 = cx1 + size(varargin{ii+1},2)-1;
r2{ii} = nn2an(rx2,cx2);
else
tmp = findstr(varargin{ii},':');
r1{ii} = varargin{ii}(1:tmp-1);
r2{ii} = varargin{ii}(tmp+1:end);
[rx1,cx1] = an2nn(r1{ii});
[rx2,cx2] = an2nn(r2{ii});
end
case 'rowcol'
if autorange
r1{ii} = varargin{ii};
[t,r]=strtok(r1{ii},',');
rx1 = str2num(t(2:end));
cx1 = str2num(r(2:end-1));
r1{ii} = nn2an(rx1,cx1);
rx2 = rx1 + size(varargin{ii+1},1)-1;
cx2 = cx1 + size(varargin{ii+1},2)-1;
r2{ii} = nn2an(rx2,cx2);
else
tmp = findstr(varargin{ii},':');
r1{ii} = varargin{ii}(1:tmp-1);
[t,r]=strtok(r1{ii},',');
rx1 = str2num(t(2:end));
cx1 = str2num(r(2:end-1));
r2{ii} = varargin{ii}(tmp+1:end);
[t,r]=strtok(r2{ii},',');
rx2 = str2num(t(2:end));
cx2 = str2num(r(2:end-1));
r1{ii} = nn2an(rx1,cx1);
r2{ii} = nn2an(rx2,cx2);
end
end
if ~autorange % Validate size match for target range, data block
sz = [rx2 - rx1 + 1, cx2 - cx1 + 1];
switch class(varargin{ii+1})
case {'double','cell'}
sz2 = size(varargin{ii+1});
case 'char'
sz2 = [size(varargin{ii+1},1),1];
129

end
if ~isequal(sz,sz2)
error(sprintf('Mismatched range/data size for input pair %d.
Specified range is %d x %d, data block is %d x
%d.',(ii+1)/2,sz(1),sz(2),sz2(1),sz2(2)));
end
end
end

Excel = actxserver('Excel.Application');
Excel.Visible = 0;
w = Excel.Workbooks;

try
excelarchive = invoke(w, 'open', fileloc);
catch
invoke(Excel, 'quit');
release(w);
delete(Excel);
error(sprintf('Sorry...unable to open file %s',fileloc));
end
Sheets = Excel.ActiveWorkBook.Sheets;

archive = Excel.Activesheet;
initval = get(archive,'Index');
archive.Unprotect;

% Read appropriate ranges into output variables


chgcount = 1;
for ii = 1:2:nargin-2
if ismember(ii,sheetchanges)
try
sheet = get(Sheets,'Item',sheetnames{chgcount});
invoke(sheet,'Activate');
archive = Excel.Activesheet;
chgcount = chgcount + 1;
continue
catch
invoke(Excel, 'quit');
release(w);
delete(Excel);
error(sprintf('\nUnable to find/open sheet
%s.',sheetnames{chgcount}));
end
elseif ismember(ii,sheetchanges + 1)
continue
end

archiverange = get(archive, 'Range', r1{ii}, r2{ii});


set(archiverange, 'value', varargin{ii+1});
release(archiverange);
end

sheet = get(Sheets,'Item',initval);
130

invoke(sheet,'Activate');

if ~promptforsave
invoke(excelarchive,'save');
end
invoke(Excel, 'quit');
release(excelarchive);
release(w);
delete(Excel);

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%
%SUBFUNCTIONS
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%

function k=iseven(x)
k = x/2==floor(x/2);
return

function [r, c] = an2nn(cr)


% convert alpha, number format to number, number format
t = find(isletter(cr));
t2 = abs(upper(cr(t))) - 64;
if(length(t2) == 2), t2(1) = t2(1) * 26; end
c = sum(t2); r = str2num(cr(max(t) + 1:length(cr)));
return

function cr = nn2an(r, c)
% convert number, number format to alpha, number format
%t = [floor(c/27) + 64 floor((c - 1)/26) - 2 + rem(c - 1, 26) + 65];
t = [floor((c - 1)/26) + 64 rem(c - 1, 26) + 65];
if(t(1)<65), t(1) = []; end
cr = [char(t) num2str(r)];

Program “testexcel.m”; this program converts the raw data in excel file to M-file format
for processing

%This file attains mat File of Farhad's Excel data.


%It saves input consistency, output consistency, input signal, and
%sampling time accordingly
%There exist 41 excel files so that j goes from 1 to 41
clc
clear
close all
k=1;
[timee,cine,coute,setp] =
readfromexcel('leo_4018pma','A:A','B:B','C:C','D:D');%Finding input
% %Modifying cell to vector
for i=1:length(cine)
cin(i)=cine{i};
cout(i)=coute{i};
time(i)=timee{i};
end
131

%%%removing last component of cin


cin=cin(1:end-1);
cout=cout(1:end-1);
time=time(1:end-1);
cinnan=cin;
coutnan=cout;
%ploting cin and interploaed
plot(time,cin,'c')
hold on
plot(time,cout)
%saving results
ts=time(2)-time(1);
save('lab1','cin','cout','ts')

Program “readfromexcel.m”

function varargout = readfromexcel(fileloc,varargin)


% VARARGOUT = READFROMEXCEL(FILELOC,VARARGIN)
%
% Uses ActiveX commands to read range(s) from an existing Excel
% spreadsheet.
%
% FILELOC: Enter a string representing the (absolute or relative)
% location of an Excel file. (Extension may be
% omitted, and will be assumed to be .xls.)
% Examples: 'c:\brett\my archives\test1.xls'
% 'test1.xls'
% 'myarchive'
%
% SHEETNAME: (Optional): Any occurrence in the variable argument
list of
% the strings 'sheetname' or 'sheet' prompts the function
to change
% active sheets to the value in the following variable.
That specifier
% must be a string argument matching exactly the name of
an existing
% sheet in the opened file. If this argument is omitted,
the function
% defaults to reading from the first sheet in the file.
%
% RANGE SPECIFIER(S): Enter the range(s) to read. The values stored in these
% ranges will be returned in consecutive output arguments.
% Example: 'B1:B5'
% 'B1:B1' (or simply 'B1')
% 'B1:P4'
% 'B:B' or 'B' (Entire second column)
% '2:2' or '2' (Entire second row)
% 'ALL' (Entire sheet)
% (Additional ranges: Comma separated ranges in the same form as above;
% contents of archive will be returned in output arguments
% 2...n)
%
% NOTE: Specifying range as 'ALL' returns entire used portion
of sheet;
132

% Specifying range as 'B:B' or '2:2' returns


% appropriate row of UsedRange. (Data are selected in
% block form as for 'ALL', then the selected row/column
% is returned.
%
% OUTPUT: If specified range is 1 cell, variable returned is of the same
% class as cell contents. If the range spans more than 1 cell, the
variables will be cell arrays.
%
% EXAMPLES: a = readfromexcel('c:\brett\my archives\test1.xls','C1:C5');
% reads from the currently active sheet
% [a,b] = readfromexcel('c:\brett\my
archives\test1.xls','sheet','sheet2','C1:C5','C1:P3');
% reads from sheet2
% [a,b,c] =
readfromexcel('myarchive','C3:D5','sheet','mysheet','E4','sheet','sheet2','B3
');
% reads a from currently active sheet, switches to sheet
% 'mysheet' to read b, then to sheet 'sheet2' to read c.
%
% Written by Brett Shoelson, Ph.D.
% shoelson@helix.nih.gov
% Update History: 1/04. Version 1.
% 2/2/04. Now allows multiple specifications of sheet name
% (at the suggestion of R. Venkat), and support of
% relative paths (thanks to Urs Schwarz). Also, inclusion
of the
% extension '.xls' is now superfluous.
% 7/21/04. Implements try/catch structure for reading of
% ranges to avoid errors that leave open activex
% connections. (Response to Chris Paterson's CSSM
% query).
% 7/29/04. Accomodates reading of entire sheet, or of
% entire row/column. (Response to email queries by Kinan
Rai
% and CSSM query by Xiong.)
%
% SEE ALSO: write2excel

if nargin < 2
msgstr = sprintf('\nAt a minimum, you must specify three input
arguments.\nThe first is a string indicating the location of the excel
file,\nand the second is a range to be read.');
error(msgstr);
end

sheetchanges =
[strmatch('sheet',varargin,'exact');strmatch('sheetname',varargin,'exact')];
if ~isempty(sheetchanges)
[sheetnames{1:length(sheetchanges)}] = deal(varargin{sheetchanges+1});
end

[pathstr,name,ext] = fileparts(fileloc);
if isempty(ext)
fileloc = [fileloc,'.xls'];
end
133

if isempty(pathstr)
fileloc = which(fileloc,'-all');
if size(fileloc,1) ~= 1
error('File was either not located, or multiple locations were found.
Please reissue readfromexcel command, providing absolute path to the file of
interest.');
end
end

Excel = actxserver('Excel.Application');
Excel.Visible = 0;
w = Excel.Workbooks;
try
excelarchive = invoke(w, 'open', fileloc);
catch
invoke(Excel, 'quit');
release(w);
delete(Excel);
error(sprintf('Sorry...unable to open file %s',fileloc));
end

Sheets = Excel.ActiveWorkBook.Sheets;

archive = Excel.Activesheet;
initval = get(archive,'Index');

% Read appropriate ranges into output variables


chgcount = 1; argcount = 1;
for ii = 1:nargin-1
readinfo = [];
if ismember(ii,sheetchanges)
try
sheet = get(Sheets,'Item',sheetnames{chgcount});
invoke(sheet,'Activate');
archive = Excel.Activesheet;
chgcount = chgcount + 1;
continue
catch
invoke(Excel, 'quit');
release(w);
delete(Excel);
error(sprintf('\nUnable to find/open sheet
%s.',sheetnames{chgcount}));
end
elseif ismember(ii,sheetchanges + 1)
continue
end

%Parse range
rangespec = 0;
if strcmp(lower(varargin{ii}),'all') %Range of the form 'ALL'
rangespec = 1;
else
tmp = findstr(varargin{ii},':');
if isempty(tmp) %Range of the form 'A' or '2' or 'A2'
r1 = varargin{ii};
134

r2 = r1;
if ~any(ismember(r1,num2str([1:9]))) %Range of the form 'A'
rangespec = 2;
elseif all(ismember(r1,num2str([1:9]))) %Range of the form '2'
rangespec = 3;
else %Range of the form 'A2'
rangespec = 4;
end
else % Range of the form 'A2:B3', '2:2', or 'A:A'
r1 = varargin{ii}(1:tmp-1);
if all(ismember(r1,num2str([1:9]))) %Range of the form '2:2'
r2 = r1;
rangespec = 5;
elseif ~any(ismember(r1,num2str([1:9]))) %Range of the form 'A:A'
r2 = r1;
rangespec = 6;
else
%Range of the form 'A1:B2'
r2 = varargin{ii}(tmp+1:end);
rangespec = 7;
end
end
end

try
switch rangespec
case 1
readinfo = get(archive,'UsedRange');
case {2,6}
readinfo = get(archive,'UsedRange');
[r,c] = an2nn(r1);
r1 = nn2an(readinfo.row,c);
[m,n] = size(readinfo.value);
r2 = nn2an(readinfo.row+m,c);
readinfo = get(archive, 'Range', r1, r2);
case {3,5}
readinfo = get(archive,'UsedRange');
[m,n] = size(readinfo.value);
r2 = nn2an(r1,readinfo.row+n);
r1 = nn2an(r1,readinfo.column);
readinfo = get(archive, 'Range', r1, r2);
case {4,7}
readinfo = get(archive, 'Range', r1, r2);
otherwise
readinfo.value = {};
fprintf('Error parsing input argument %d.',ii+1);
end
catch
fprintf('Error reading range specified by input argument %d.',ii+1);
invoke(Excel, 'quit');
release(excelarchive);
release(w);
delete(Excel);
return
end
varargout{argcount} = readinfo.value;
argcount = argcount + 1;
135

end
% Reset to initial active sheet
sheet = get(Sheets,'Item',initval);
invoke(sheet,'Activate');

try
release(readinfo);
end
invoke(excelarchive,'close'); %This closes without saving, so changing the
active sheet is temporary
%invoke(excelarchive,'save'); %Note: Instead of invoke(excelarchive,'close'),
I use the save option after
% switching back to the initially active sheet.
This stops Excel from showing
% "previously saved versions" when the file is
% next opened.
invoke(Excel, 'quit');
release(excelarchive);
release(w);
delete(Excel);
return

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%
%SUBFUNCTIONS
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%
function cr = nn2an(r, c)
% convert number, number format to alpha, number format
%t = [floor(c/27) + 64 floor((c - 1)/26) - 2 + rem(c - 1, 26) + 65];
t = [floor((c - 1)/26) + 64 rem(c - 1, 26) + 65];
if(t(1)<65), t(1) = []; end
cr = [char(t) num2str(r)];

function [r, c] = an2nn(cr)


% convert alpha, number format to number, number format
t = find(isletter(cr));
t2 = abs(upper(cr(t))) - 64;
if(length(t2) == 2), t2(1) = t2(1) * 26; end
c = sum(t2); r = str2num(cr(max(t) + 1:length(cr)));
136

APPENDIX C: Program for cavern volume determination in CFD


#include "udf.h"

DEFINE_ON_DEMAND(cavern_calc_s)

{
Domain *d;
real strain, volume, vol_tot;
Thread *t;
cell_t c;
d = Get_Domain(1);

thread_loop_c(t,d)
{
begin_c_loop(c,t)
{
volume = C_VOLUME(c,t);
strain = C_STRAIN_RATE_MAG(c,t);

if (strain >= 0.984)


vol_tot += volume;
}
end_c_loop(c,t)
}

printf("The cavern volume is %10.8f", vol_tot);


}
137

APPENDIX D: Program for tracer analysis in CFD


#include <udf.h>

DEFINE_PROFILE(hwtracer_mf_profile, t, i)
{
real flow_time = CURRENT_TIME;
face_t f;

begin_f_loop(f,t)
{
if (flow_time > 3411)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 3284)
F_PROFILE(f, t, i) = 0.1096;
else if (flow_time > 3192)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 3070)
F_PROFILE(f, t, i) = 0.1096;
else if (flow_time > 2986)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 2531)
F_PROFILE(f, t, i) = 0.1096;
else if (flow_time > 2453)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 2208)
F_PROFILE(f, t, i) = 0.1096;
else if (flow_time > 2130)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 1992)
F_PROFILE(f, t, i) = 0.1096;
else if (flow_time > 1803)
F_PROFILE(f, t, i) = 0.0847;
else if (flow_time > 1566)
F_PROFILE(f, t, i) = 0.1096;
else
F_PROFILE(f, t, i) = 0.0847;
}
end_f_loop(f,t)
}
138

APPENDIX E: ERT data


Pulp type: hardwood; Mass concentration: 2%; E/D = 0.6; Z/T = 1.0

Impeller speed ERT images


(N) (rpm)
P1 P2 P3
125

P4 P5

P1 P2 P3
150

P4 P5

P1 P2 P3
200

P4 P5

Conductivity low high


scale
139

Pulp type: hardwood; Mass concentration: 2%; E/D = 0.6; Z/T = 1.0

Impeller speed ERT images


(N) (rpm)
P1 P2 P3
225

P4 P5

P1 P2 P3
250

P4 P5

Conductivity
low high
scale
140

Pulp type: hardwood; Mass concentration: 2%; E/D = 0.4; Z/T = 1.0

Impeller speed ERT images


(N) (rpm)
P1 P2 P3
125

P4 P5

P1 P2 P3
150

P4 P5

P1 P2 P3
200

P4 P5

Conductivity low high


scale
141

Pulp type: hardwood; Mass concentration: 2%, E/D = 0.4; Z/T = 1.0

Impeller speed ERT images


(N) (rpm)
P1 P2 P3
225

P4 P5

P1 P2 P3
250

P4 P5

Conductivity
low high
scale
142

Pulp type: hardwood; Mass concentration: 2%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

125 P1 P2

P3 P4

150 P1 P2

P3 P4

200 P1 P2

P3 P4

Conductivity low high


scale
143

Pulp type: hardwood; Mass concentration: 2%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

225 P1 P2

P3 P4

250 P1 P2

P3 P4

Conductivity
low high
scale
144

Pulp type: hardwood; Mass concentration: 3%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

425 P1 P2

P3 P4

450 P1 P2

P3 P4

475 P1 P2

P3 P4

Conductivity low high


scale
145

Pulp type: hardwood; Mass concentration: 3%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

500 P1 P2

P3 P4

525 P1 P2

P3 P4

Conductivity
low high
scale
146

Pulp type: hardwood; Mass concentration: 4%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

550 P1 P2

P3 P4

575 P1 P2

P3 P4

600 P1 P2

P3 P4

Conductivity low high


scale
147

Pulp type: hardwood; Mass concentration: 4%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

650 P1 P2

P3 P4

675 P1 P2

P3 P4

Conductivity
low high
scale
148

Pulp type: softwood; Mass concentration: 3%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

425 P1 P2

P3 P4

450 P1 P2

P3 P4

475 P1 P2

P3 P4

Conductivity low high


scale
149

Pulp type: softwood; Mass concentration: 3%; E/D = 0.4; Z/T = 0.8

Impeller speed ERT images


(N) (rpm)

500 P1 P2

P3 P4

525 P1 P2

P3 P4

Conductivity
low high
scale

You might also like