You are on page 1of 117

Modelling the Maximum

Capacity of a Pulp Pressure


Screen
by

Hayder Jabber Salem

B.Sc. University of Baghdad, 1993


M.Sc. University of Baghdad, 1997
M.Sc. American University of Sharjah, 2006

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

The Faculty of Graduate and Postdoctoral Studies

(Mechanical Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA


(Vancouver)
September 2013

c Hayder Jabber Salem 2013
Abstract

Pressure screens are used as a means of separating pulp fibres from contaminants.
They are also used to improve pulp quality by fractionating fibres by length. Both
functions are limited by the capacity of the device. Three studies were conducted in
this thesis to understand the factors that affect maximum capacity.
Capacity is determined by the complex hydrodynamics in the region between the
screen rotor and the screen wall. To better understand this flow, the stream-wise
velocity and aperture velocities were measured using particle image velocimetry. The
vortex generated above the aperture and its size is shown to be strongly dependent
on the aperture velocity, wall roughness and, to some extent, on the rotor speed.
The vortex diminishes in size at higher aperture velocities that increase the exit layer
height. The experiments also show that the reversal flow through the slot decreased
with lower rotor speeds and increased slot velocities. This observation challenges
the existing models of apertures being cleared simply by flow reversal driven by a
suction pulse caused by flow acceleration between the foil and cylinder. In its place,
this study identifies elements of a more sophisticated flow model that considers such
factors as the depletion of the zone below the rotor and flow disturbances in the wake
of the foil. The effects of screen cylinder geometry, pulp type, rotor and flow velocities
on capacity were also investigated. Five types of screen cylinders were tested using
different ratios of softwood/hardwood kraft pulp and different reject rates. It was
found that the average fibre length has a significant impact on capacity.

ii
Abstract

A comprehensive understanding of pulp screen capacity remains elusive. The


present research has, however, provided insights which move away from the simplistic
”backflush” models used in the past and supports a more sophisticated model that
also considers: 1) the dynamics of fibre accumulation and removal rates of the slot
entry, 2) the importance of a small-scale perturbations created by turbulent flow
for fibre removal, and 3) the mechanics of fibre trapping on the downstream edge
of the slot. These advancements also provide some direction for future equipment
developments.

iii
Preface

In this work, I was responsible for conducting all parts of the research and all of
the simulations and analysis of the data presented. Dr. Olson and Dr. Martinez
supervised the research and provided feedback and reviewed the manuscript. Dr.
Robert Gooding from AFT also provided valuable feedback and reviewed the work
presented.

iv
Table of Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

List of Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 Flow of Fibres Near a Screen Wall . . . . . . . . . . . . . . . . . . . . 4
2.2 Effect of Wall Geometry, Local Flows and Flocculation on Fibre Screening 5
2.3 Effect of Rotor on Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 7

v
Table of Contents

3 Experimental Study of Flow Fields Near Screen Cylinder Surface 9


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Experimental Details . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.3.1 Steady State Flow Analysis . . . . . . . . . . . . . . . . . . . 14
3.3.2 Time-Varying Flow Analysis . . . . . . . . . . . . . . . . . . . 19
3.3.3 Fibre Concentration Analysis . . . . . . . . . . . . . . . . . . 33
3.4 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 37

4 Maximum Capacity of a Pilot Pressure Screen . . . . . . . . . . . 39


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Experimental Details . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.1 Effect of Consistency on Capacity . . . . . . . . . . . . . . . . 44
4.3.2 Effect of Feed Flow Rate on Capacity . . . . . . . . . . . . . . 48
4.3.3 Effect of Pulp Properties and Contaminants on Capacity . . . 49
4.3.4 Effect of Cylinder Geometry on Capacity . . . . . . . . . . . . 50
4.3.5 Effect of Rotor and Slot Velocity on Reject Thickening . . . . 56
4.4 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 56

5 New Model for Pressure Screen Capacity . . . . . . . . . . . . . . . 59


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.1.1 Visual Observation of Fibre Motion Near the Screen Surface . 60
5.2 Mathematical Model Details . . . . . . . . . . . . . . . . . . . . . . . 60
5.3 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 74

vi
Table of Contents

6 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . 76


6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.1.1 Flow Field Study (Chapter 3) . . . . . . . . . . . . . . . . . . 76
6.1.2 Pilot Pulp Screen Capacity (Chapter 4) . . . . . . . . . . . . . 77
6.1.3 Model of Pulp Screen Capacity (Chapter 5) . . . . . . . . . . 78
6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
A Flow and Wire Geometry Effects on Fibre Fractionation . . . . . . . 87
B PIV Uncertainty Analysis . . . . . . . . . . . . . . . . . . . . . . . . 93
C Characterization of Fibres Used for Capacity Trials . . . . . . . . . . 96

vii
List of Tables

3.1 Coupon geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13


3.2 Experimental test conditions . . . . . . . . . . . . . . . . . . . . . . . 13

4.1 Screen wire geometries . . . . . . . . . . . . . . . . . . . . . . . . . . 43

B.1 Principal dimensions for PIV measurment . . . . . . . . . . . . . . . 94


B.2 Uncertainties for velocity . . . . . . . . . . . . . . . . . . . . . . . . . 95
B.3 Uncertainties for position . . . . . . . . . . . . . . . . . . . . . . . . . 95

viii
List of Figures

1.1 Schematic of the principal flows through a pulp screen. . . . . . . . . 2

3.1 Flow geometry studied experimentally. . . . . . . . . . . . . . . . . . 10


3.2 Schematic diagram of the Cross-Sectional Screen apparatus (CSS). . . 12
3.3 Close-up of the wire cross-section at the entry to the slot showing
characteristic wire dimensions . . . . . . . . . . . . . . . . . . . . . . 13
3.4 Smooth rotor-wall flow geometry. . . . . . . . . . . . . . . . . . . . . 15
3.5 Slot-entry vortex for a high contour at (a) Vt = 10 m/s, Vs = 1 m/s,
(b) Vt = 10 m/s, Vs = 4 m/s, (c) Vt = 20 m/s, Vs = 1 m/s, (d) Vt =
20 m/s, Vs = 4 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.6 Slot-entry vortex for a low contour at (a) Vt = 10 m/s, Vs = 1 m/s,
(b) Vt = 10 m/s, Vs = 4 m/s, (c) Vt = 20 m/s, Vs = 1 m/s, (d) Vt =
20 m/s, Vs = 4 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.7 Vortex stagnation point length. . . . . . . . . . . . . . . . . . . . . . 17
3.8 Exit layer height (normalized by slot width) versus the ratio of slot
velocity and upstream velocity. . . . . . . . . . . . . . . . . . . . . . 17
3.9 High contour circulation within the vortex as a function of flow ratio. 18
3.10 Low contour circulation within the vortex as a function of flow ratio. 18
3.11 Foil rotor-wall geometry. . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.12 ROI’s for u and v components. . . . . . . . . . . . . . . . . . . . . . . 20

ix
List of Figures

3.13 The flow between the foil and cylinders surface are examined for two
alternate frames of reference: (a) a steady flow problem follows from
fixing the frame of reference on the foil, which is moving at u1 . The flow
between the foil and cylinder surface increases by an amount ∆u. (b)
An unsteady flow situation results from fixing the frame of reference to
the screen cylinder. The apparent velocity under the rotor is decreased
by an amount ∆u relative to the general flow velocity. . . . . . . . . . 21
3.14 u component for low contour at Vt = 15 m/s and Vs = 1 m/s. . . . . 25
3.15 u component for high contour at Vt = 20 m/s and Vs = 1 m/s. . . . . 25
3.16 u component for high contour at Vt = 20 m/s and Vs = 2 m/s. . . . . 25
3.17 Rotor position in terms of chord length. . . . . . . . . . . . . . . . . 26
3.18 The spatio-temporal description of the u-component in the flow field
for the high contour (Vt = 20 m/s, Vs = 1 m/s) is shown above (a)
along with the measuring volume (in yellow) used for averaging. The
lower plot (b) is the associated averaged flow for the u-component. . . 27
3.19 Spatio-temporal behaviour of u component for high contour surface at
Vt = 10 m/s, Vs = 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . 28
3.20 Spatio-temporal behaviour of u component for high contour surface at
Vt = 10 m/s, Vs = 3 m/s. . . . . . . . . . . . . . . . . . . . . . . . . 28
3.21 Spatio-temporal behaviour of u component for high contour surface at
Vt = 20 m/s, Vs = 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . 29
3.22 Spatio-temporal behaviour of u component for high contour surface at
Vt = 20 m/s, Vs = 3 m/s. . . . . . . . . . . . . . . . . . . . . . . . . 29
3.23 v component for high contour at Vt = 20 m/s. . . . . . . . . . . . . . 30
3.24 v component for low contour at Vt = 10 m/s. . . . . . . . . . . . . . . 30

x
List of Figures

3.25 Reversal flow boundary. . . . . . . . . . . . . . . . . . . . . . . . . . 31


3.26 Spatio-temporal behaviour of Vd for low contour surface at Vt = 10 m/s,
Vs = 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.27 Spatio-temporal behaviour of Vd for high contour surface at Vt =
20 m/s, Vs = 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.28 Light intensity as a function of fibre concentration . . . . . . . . . . . 34
3.29 High contour concentration ROI’s . . . . . . . . . . . . . . . . . . . . 34
3.30 Concentration for high contour: (a) Vt = 10 m/s, Vs = 1 m/s, (b) Vt =
10 m/s, Vs = 2 m/s, (c) Vt = 10 m/s, Vs = 3 m/s, (d) Vt = 20 m/s,
Vs = 1 m/s, (e) Vt = 20 m/s, Vs = 2 m/s, (f) Vt = 20 m/s, Vs =3 m/s. 36

4.1 The UBC MR8 pilot pressure screen. . . . . . . . . . . . . . . . . . . 42


4.2 Screen open area blinding procedure: (a) wrapping the spool with
rubber strip and (b) securing the strip with a stainless steel clamp to
prevent leakage from the spool edges. . . . . . . . . . . . . . . . . . . 43
4.3 Plugging envelope for Type B cylinder at different consistencies, (50/50
hardwood/softwood mixture). . . . . . . . . . . . . . . . . . . . . . . 46
4.4 The extrapolated Vs -Vt relationship of Figure 4.3. . . . . . . . . . . . 46
4.5 The x-intercepts of the extrapolated Vs -Vt relationship of Figure 4.4
is plotted against consistency suggesting that there would be no offset
when there are no fibres in the suspension (CF = 0%). . . . . . . . . 47
4.6 Pressure difference across Type B screen cylinder prior to plugging at
a range of slot velocities (50/50 hardwood/softwood mixture). . . . . 47
4.7 Power consumption by the rotor as a function of rotor speed at different
slot velocities (CF = 0%). . . . . . . . . . . . . . . . . . . . . . . . . 48

xi
List of Figures

4.8 Effect of reject rate on plugging envelope for the Type B screen (100%
softwood suspension with 1.0% consistency). . . . . . . . . . . . . . . 49
4.9 Plugging envelope for the Type B screen using different softwood/hardwood
ratios, CF =1.0%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.10 Pressure difference across the Type B cylinder for different softwood/hardwood
ratios, CF = 1.0%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.11 Minimum rotor tip speed for the Type B screen at different contam-
inant ratios (50/50 hardwood/softwood mixture, CF = 1.5%) Vt = 2
m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.12 Pressure difference across the Type B screen at different contaminant
ratios, (50/50 hardwood/softwood mixture, CF = 1.5%. . . . . . . . . 52
4.13 Plastic speck diameter distribution. . . . . . . . . . . . . . . . . . . . 53
4.14 Effect of contour height on the plugging envelope. The contour heights
of the Type A, B and C screen cylinders is 1.2 mm, 0.9 mm and 0.6
mm, respectively. (50/50 hardwood/softwood mixture, CF = 1.5%). . 54
4.15 Pressure difference across cylinders, (50/50 hardwood/softwood mix-
ture, CF = 1.5%). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.16 Effect of wire width and slot width on plugging boundary, (50/50 hard-
wood/softwood, CF = 1.0%). Type B: 0.9 mm contour height/3.2 mm
wire width/0.15 mm slot width. Type D: 0.6 mm /2.3 mm/0.15 mm.
Type E: 0.6 mm /3.2 mm/0.10 mm. . . . . . . . . . . . . . . . . . . 55
4.17 Thickening of pulp with rotor and slot velocity changes (Type A screen,
CF = 1.0%, 50:50 SW:HW). . . . . . . . . . . . . . . . . . . . . . . . 57

xii
List of Figures

5.1 Fibre trapping and clearing by rotor, Vt = 5 m/s and Vs = 3 m/s. (a)
xr /chord = -1.0, (b) xr /chord = 0.35, (c) xr /chord = 0.6, (d) xr /chord
= 1, (e) xr /chord = 1.35, (f) xr /chord = 4.0. . . . . . . . . . . . . . 61
5.2 Tension on a fibre trapped on wire edge. . . . . . . . . . . . . . . . . 63
5.3 Tension on the fibre portion above slot at Vt = 5 m/s. . . . . . . . . 64
5.4 Tension on fibre portion above slot at Vt = 20 m/s. . . . . . . . . . . 65
5.5 Tension on fibre segments above the slot as a function of smooth rotor
tip speed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.6 Universal curve-fitted u component above the cylinder surface as a
function of rotor position for 0 ≤ Vs ≤ 4 m/s. . . . . . . . . . . . . . 66
5.7 Corrected slot velocity along a 3 mm fibre within the slot. . . . . . . 67
5.8 Tension changes with different trapped fibre positions are shown schemat-
ically in images (a) and (b) and analytically in a comparison of the fibre
forces. In this example, a fibre with l/Lf in the range of 0.50 to 0.66
is trapped (i.e. differences in drag force are less than the friction force). 68
5.9 Changes in trapping position due to changes in the flow field (Vt = 5
m/s and Vs = 1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.10 Estimates of the accumulation of 3 mm fibres on the low contour wire
during one foil cycle at Vt = 10 m/s. . . . . . . . . . . . . . . . . . . 72
5.11 Accumulation of fibres within screen slots during rotor cycles, (a) ac-
cumulation is less than the cleansing effect, i.e. no fibres left from
the previous cycle (b) accumulation effect equals the cleansing effect
(plugging point), (c) accumulation effect is greater than the cleansing
effect (plugging takes place). . . . . . . . . . . . . . . . . . . . . . . . 73

xiii
List of Figures

5.12 Accumulation number, Nac , for a 3 mm fibre as a function of slot


velocity at different rotor speeds. . . . . . . . . . . . . . . . . . . . . 74

A.1 Effect of slot velocity on fibre length distribution with a low contour
wire and smooth rotor at Vt = 20 m/s. . . . . . . . . . . . . . . . . . 89
A.2 Effect of slot velocity on fibre length distribution with a low contour
wire and smooth rotor at Vt = 25 m/s. . . . . . . . . . . . . . . . . . 89
A.3 Effect of slot velocity on fibre length distribution with a high contour
wire and smooth rotor at Vt = 20 m/s. . . . . . . . . . . . . . . . . . 90
A.4 Effect of slot velocity on fibre length distribution with a high contour
wire and smooth rotor at Vt = 25 m/s. . . . . . . . . . . . . . . . . . 90
A.5 Effect of smooth rotor speed and contour height on fibre fractionation
at various slot velocities. . . . . . . . . . . . . . . . . . . . . . . . . . 91
A.6 Effect of contour height and vortex strength on short and long fibre
fractionation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
A.7 Effect of rotor on fibre fractionation at the same flow field. . . . . . . 92
C.1 Fibre length distributions of 0:100 SW:HW sample. . . . . . . . . . . 96
C.2 Fibre length distributions of 25:75 SW:HW sample. . . . . . . . . . . 97
C.3 Fibre length distributions of 50:50 SW:HW sample. . . . . . . . . . . 97
C.4 Fibre length distributions of 75:25 SW:HW sample. . . . . . . . . . . 98
C.5 Fibre length distributions of 100:0 SW:HW sample. . . . . . . . . . . 98
C.6 Frequency of 3 mm fibres in different SW:HW samples. . . . . . . . . 99

xiv
List of Symbols

Alphanumeric Symbols
Symbol Meaning Unit
A Screen open area m2
A0 , A2 Parameters used in CD curve fits
Av Vortex area m2
C Consistency (concentration) of fibres
C1 , C2 Capacity characteristic constants
CF Feed consistency
CS Consistency through slot
CU Consistency of the stream above slot
Chord Air foil chord length m
CD Drag coefficient
CD⊥ Drag coefficient of a cylinder normal to flow
CL Lift coefficient
ci Sensitivity coefficient
d Stagnation point length m
D Fibre and particle diameter m
fc Coefficient of friction
h Wire contour height m
H Exit layer height m
Hr Rotor-wire tip gap m
k1 , k2 , k3 Accumulation constants
l Fibre length above slot m
L Wire inclined surface length m
Lf Fibre length m
Lr Distance on the image plane m
lr Distance of the reference point m
Nac Accumulation number
P Pressure kg/(m.s2 )

xv
List of Symbols

Symbol Meaning Unit


Re Reynolds number
Rv Reject rate
s Slot width m
T Thickening factor
Ts Tension on fibre portion within slot kg.m/s2
Tu Tension on fibre portion above slot kg.m/s2
u x-direction velocity component m/s
us Standard uncertainty
v y-direction velocity component m/s
Vd Velocity parallel to wire inclined surface m/s
Vs Slot velocity m/s
Vs∗ Maximum slot velocity m/s
Vt Rotor tip speed m/s
Vt∗ Threshold rotor tip speed m/s
Vu Upstream flow velocity m/s
w Wire width m
X Horizontal coordiante of particle in image plane m
Xe End location of particle pixel
Xs Start location of particle pixel
xr Foil leading edge-origin tangential distance m
xac Accumulation distance per foil cycle m
z Stream line elevation m

Greek Symbols
Symbol Meaning Unit
α Angle between the flow and fibre degrees
α1 Magnification factor m/pixel
µ Dynamic viscosity kg/(m.s)
θ Contact angle degrees
κ1 , κ2 Parameters used in CD,⊥ curve fits
ρ Density kg/m3
Γ Circulation m2 /s

xvi
List of Abbreviations

Abbreviations
Abbreviation Meaning
CSS Cross Sectional Screen
CFD Computational Fluid Dynamics
NACA National Advisory Committee for Aeronautics
HSV High speed videography
HW Hardwood
LDV Laser Doppler Velocimetry
OCC Old corrugated container
PIV Particle Image Velocimetry
ROI Region of interest
SW Softwood
VFD Variable frequency drive

xvii
Acknowledgements

During the last few years that I spent working on this research, I have received
help and support from many people to whom I would like to express my sincere
appreciation.
First of all, I would like to thank my supervisors Prof. James Olson and Prof.
Mark Martinez who patiently introduced me to the challenging field of pulp screening.
I would like to acknowledge their continuous support, advice and patience. Without
their help, this work would have not been accomplished.
I would also like to acknowledge the financial support of NSERC and Aikawa Fibre
Technologies (AFT).
Special thanks to Dr. Robert Gooding for his support, feedback and proofreading
of this thesis and published work.
I would like to thank all my colleagues and friends at the University of British
Columbia for their valuable help especially George Soong, Nici Darychuck, Dr. Shel-
don Green, Ario Madani, Sean Delfel, my brother Zaid and Mahdi Salehi.
Finally, I would like to thank my lovely wife Jennifer for her support and under-
standing.

xviii
Chapter 1

Introduction

The purpose of the present investigation arises from an interest in the screening of
papermaking fibre suspensions. Periodic roughness elements, formed into a screen
surface, are used as a means of separating pulp fibres from contaminants. Figure 1.1
shows the principal features of a typical industrial screen used for pulp screening. A
screen receives a feed stream of pulp and separates it into an accept stream of cleaned
pulp and a reject stream where oversized contaminants are concentrated. Screens are
also used to enhance pulp quality by fractionating one type of fibre length from oth-
ers. A screen cylinder within the pulp screen has apertures that allow pulp fibres to
pass while blocking fibre bundles, plastic specks and other contaminants. The rotor
is thought to work by creating three important effects: (1) generating pressure pulsa-
tions which backflush and clean the screen plate apertures of any incipient blockages,
(2) accelerating the pulp suspension on the feed side of the screen to a high tangen-
tial velocity and (3) inducing turbulence at the surface of the screen to help keep the
apertures clear.
The capacity of the screen is commonly defined as the maximum throughput (i.e.
the mass flow rate) before the apertures become plugged with pulp. For the purpose
of this study, ”capacity” will be expressed as slot velocity, with the screen open
area, slot width and accept consistency taken as constants. Capacity is a function
of fibre length, feed consistency, aperture size and geometry, rotor speed and rotor

1
1.1. Objectives

Figure 1.1: Schematic of the principal flows through a pulp screen.

type. Over the decades, there have been various experimental and computational
studies conducted to understand the performance of pressure screening in terms of
fibre passage and debris removal efficiency. However, there are relatively few studies
on the fundamental aspects of capacity.

1.1 Objectives

The goal of this study is to develop a theoretical and experimental understanding


of the factors that limit the capacity of pulp pressure screens. This requires the
assessment of the flow patterns above and within the apertures of a pressure screen
and the investigation of the maximum capacity in terms of main variables such as fibre
length, feed consistency, aperture size and geometry, rotor type and debris content.
The specific objectives of this thesis are to:

• Use High Speed Videography (HSV) to investigate the flow of pulp fibres in the

2
1.2. Outline

region between the pulp screen rotor and the slotted screen wall.

• Use Particle Image Velocimetry (PIV) to measure the stream-wise and aperture
velocities and understand the complex hydrodynamics in the critical region
between the rotor and the screen wall.

• Experimentally determine the effect of rotor speed on screen capacity.

• Experimentally determine the effect of pulp type on screen capacity.

• Experimentally determine the effect of screen aperture geometry on screen ca-


pacity.

• Analytically determine the factors that affect screen capacity and develop a
capacity model.

1.2 Outline

A review of published work is presented in Chapter 2. Chapter 2 mainly focuses


on the flow of fibres near screen surface and the effect of wall geometry, rotor and
flocculation on fibre screening. Chapter 3 presents experimental assessments of the
flow behaviour, using PIV technique, and observations of the flow of individual fibres
near the screen surface. Chapter 4 assesses the factors that limit capacity using a pilot
pressure screen. By utilising the data presented in the previous chapters, Chapter 5
discusses a new model for screen capacity. Finally, Chapter 6 provides conclusions
and discusses opportunities for future work.

3
Chapter 2

Literature Review

2.1 Flow of Fibres Near a Screen Wall

Gooding and Kerekes [18] used high-speed photography to capture the trajectories of
fibres near a single slot in a cross flow. They showed a low fibre concentration layer
near the wall which was one of the factors limiting fibre passage, and was called the
"wall effect". They also observed that under certain flow conditions, fibres would
become "trapped" on the downstream edge of the slot. Kumar [31] experimentally
determined the effect of fibre length, slot width and flow velocities on the probability
of fibre passage. He found that fibre passage varied with the ratio of fibre length
to slot width. This work was later investigated by Olson and Wherrett [41]. They
showed that fibre passage can be characterized by the single function first proposed
by Kumar.
Yu and DeFoe [52] studied the flow patterns at the feed-side surface of smooth and
contoured screen cylinders under steady flow conditions. They speculated that fibre
mats containing fibres and contaminants can form on the surface of the cylinder and
that these mats remixed and were diluted during screen operation. In a further study,
Yu and DeFoe [53] showed that for contoured cylinders, vortex flow is an effective
means of preventing fibres from residing on slot openings or stapling between two
openings. Stapling is a mechanism in which the two ends of a fibre are drawn into

4
2.2. Effect of Wall Geometry, Local Flows and Flocculation on Fibre Screening

adjacent apertures with the fibre immobilized on the land area between them [17].
Gooding [16] examined the flow patterns and turbulence levels of various smooth
and contoured screen slots in a steady cross-flow to obtain a mechanistic understand-
ing of what determines slot resistance and the conditions that lead to minimum values
of slot resistance. He also observed how fibres build up within the slot and considered
the contribution of the incipient fibre blockages to flow resistance. The concentration
and orientation of fibres in a turbulent flow near a smooth wall was assessed by Olson
[39] and found to be a function of fibre length.
Yong et al. [51] used high-speed video photography to determine the trajectory
of nylon fibres approaching narrow screen apertures in a laboratory pulp screen.
However, the very stiff model fibres and the relatively low rotor speeds differed from
industrial screening conditions, which limited the relevance of the study.

2.2 Effect of Wall Geometry, Local Flows and

Flocculation on Fibre Screening

Screen cylinder contours have an important effect on screening efficiency and capacity.
Heise [23] recommends the use of screen cylinders which minimize turbulence between
the rotor and the screen surface in order to maximize screening efficiency. Niinimaki et
al. [38] found that the efficiency of probability screening is more dependent on contour
roughness than on slot width. Increasing the contour height led to an improvement
in screen capacity and reduction in debris removal efficiency.
Halonen et al. [21] defined fluidization as a process of loosening fibre contacts
inside flocs. Depending on the conditions and properties of the pulp on the contoured
screen surface, the fluidized state may extend the cleansing effect induced by the rotor

5
2.2. Effect of Wall Geometry, Local Flows and Flocculation on Fibre Screening

according to Frejborg [13]. Bliss [4] agreed with the idea of extended fluidization and
stated that turbulence allows the openings to pass more fibres after the cleaning pulse
of the rotor in the case of a contoured screen than for that of a smooth screen.
Mokamati et al. [35] developed a CFD simulation of the flow through a screen
slot and considered the effect of various screen contours. As contour height increased,
turbulence intensity near the wall increased. Moreover, turbulence intensity near the
wall increased with decreasing wire width. It was thus shown that the ratio of contour
height to wire width controls the boundary layer thickness and turbulence intensity
near the wall.
Yu and DeFoe [54] showed that the size of the apertures in a screen cylinder
was the dominant factor affecting a screen’s contaminant removal efficiency. As one
would expect, the larger the aperture size, the lower was the contaminant removal
efficiency and the higher was the volumetric and mass throughput. They also showed
that aperture velocity cannot be used as a general guideline for predicting screen
performance.
Kerekes [30] investigated the behaviour of pulp flocs at the entry to constrictions
using high speed cine photography. The elongational strain imposed by the accel-
erating flow in the entry stretched the floc but did not rupture it. When flocs did
rupture, they did so by tensile stretching rather than by shear. On entering the con-
striction, fibres did not maintain their relative lateral position perpendicular to the
flow direction; rather, they tended to concentrate along the wall of the constriction,
making subsequent floc dispersion more difficult. Blaser [3] observed the behaviour of
preflocculated ferric hydroxide flocs subjected to either a two-dimensional straining
flow or simple shear flow. He found that the simple shear flow led to the rotation
of the flocs. In the extensional flow, no rotation occurred and the flocs were broken

6
2.3. Effect of Rotor on Flow

apart along the straining axis.


Paul et al. [44] showed that the use of higher viscous liquid as the suspending
medium produces increased screen capacity. The maximum mass flow rate, i.e. max-
imum aperture velocity, increases as the suspending medium viscosity increases. The
reduction in floc size in more viscous liquids produces a decrease in the pressure loss
across the screen plate. The exit layer height increases with increasing viscosity at
constant accept and rotor velocities. The exit layer was defined by Gooding [18] and
mathematically analyzed by Olson and Kerekes [40]. Fibres within this layer are
candidates to pass through screen apertures.

2.3 Effect of Rotor on Flow

Rotor design is critical in determining pulp screen performance. The effects of pres-
sure pulse magnitude, pulse width and frequency on capacity are not, however, well
understood.
Karvinen and Halonen [29] assessed pressure pulsations using experimental and
computational techniques for a foil-type rotor. They speculated that the backflushing
action of the pressure pulse arose from a Venturi effect created by the acceleration of
the flow through the gap between the moving rotor tip and stationary screen plate.
This acceleration causes the local pressure on the feed side of the screen plate to
decrease to the point that there is a reversal in the flow through the aperture.
Pinon et al. [45] measured the pressure pulse in a laboratory pulp screen. Their re-
sults showed that increased rotor speed increased pulse strength. Although increased
rotor speed shortened the duration of the pulse, the shape of the pulse was relatively
unchanged.

7
2.3. Effect of Rotor on Flow

Levis [32] found that a wider gap between the foil and screen surface leads to
increased debris removal efficiency, albeit with lower capacity. Niinimaki et al. [37]
found that screen capacity is sensitive to changes in foil "angle of attack" and thus
the foil angle could be varied to optimize screen performance.
Feng et al. [12] simulated the pressure pulse using CFD and compared the results
to experimental measurements over a wide range of foil tip speeds, angles-of-attack,
clearances, and foil cambers. The pressure pulse peak was found to increase linearly
with the square of tip speed for all angles-of-attack studied. The positive pressure
peak near the leading edge of the foil was eliminated for foils operating at a positive
angle-of-attack. The results also showed that the magnitude of the negative pressure
peak increased as clearance decreased.
The effect of varying pulse frequency on rotor performance was studied by Delfel
et al. [9] numerically and experimentally. They used a novel multi-element foil (MEF)
rotor. The results show that a two-foil rotor had greater capacity and reduced power
consumption in comparison to a three-foil rotor with identical foil geometry. The
CFD model they introduced hypothesised a back-flush flow caused by the accelera-
tion of the flow under the foil (Bernoulli’s effect). However, they used a solid wall
boundary condition rather than modelling the slots in the cylinder wall, which may
have influenced the results.
In general, most published work has focused on the effect of flow on screen per-
formance without taking into full account such industrial factors such as the effect of
the fibre network or contour geometry on overall performance. More importantly, the
published work has not led to a detailed and rigorous understanding of the hydraulic
aperture clearing mechanism which is critical to reliable screen operation.

8
Chapter 3

Experimental Study of Flow Fields


Near Screen Cylinder Surface

3.1 Introduction

This work considers the turbulent cross-flow of a Newtonian fluid within a moving
upper boundary, i.e. the screen rotor, and a contoured, slotted lower boundary un-
dergoing suction as shown in Figure 3.1. The two-dimensional roughness of the lower
boundary is formed by placing the roughness elements in a test coupon transverse
to the approaching flow. The test coupon is made from contoured wires banded to-
gether to form a regularly-spaced array of two-dimensional roughness elements which
define narrow slots in between. Turbulent flow over a rough wall has been studied
extensively because of its importance in a range of industrial applications ([24], [43],
[5], [19]).

3.2 Experimental Details

The present study measures the flow field and concentration of pulp fibres near a
pulp screen wall. The experiments were conducted with a Phantom
R
high-speed
video camera using a cross-sectional screen (CSS). The CSS is a laboratory-scale

9
3.2. Experimental Details

Figure 3.1: Flow geometry studied experimentally.

screen modelled on the cross-section of a Hooper PSV 2100 pulp screen [[14], [12],
[45], [51]]. The CSS has a depth of 50 mm and the partially-slotted cylinder has an
inside diameter of 290 mm. Figure 3.2 provides a schematic drawing of the associated
flow loop. Feed flow is supplied from a 150 litre reservoir through the feed port to the
CSS. Accept and reject flows are controlled using magnetic flow meters, flow control
TM
valves and a LabView -based control system. Two rotors were used in this study:
The one used for most tests had NACA 0015 foils with a 6 degree angle-of-attack.
This angle minimizes the positive component of the pressure pulse according to Feng
et al. [12], which is desirable since the positive pressure pulse component is not
thought to help clear the aperture, but may instead jam the slot with trapped fibres
and force deformable contaminants through the aperture. Pulp screens can typically
be found with cylinders ranging from 0.3 to 1.2 m in diameter. The chord length of
the foil has an important effect on the shape and magnitude of the pressure pulse
and, in turn, on the performance of a pulp screen [14]. The chord length for this
first rotor was 4 cm which is comparable to the length of some foils used in industrial
screens and gives 11 chord lengths per revolution. The gap between the rotor and

10
3.2. Experimental Details

the wall was set to 2 mm. The second rotor was a non-industrial, solid-core rotor
with no elements on its surface and was used to provide a steady circumferential flow
within the screen (i.e. without pulsations). The gap between this rotor and the wall
was 14.5 mm.
The primary component of the experimental system is the test coupon which
represents the slotted rough wall. All test coupons used in these tests were 50 mm
wide. Two coupons with different contour heights were tested (see Figure 3.3 and
Table 3.1). The geometry of the rough wall used in this study was modelled on
industrial pulp screen cylinders. The coupons were made from commercial, stainless
steel, screen cylinder wires. The design of the flow channel is modular so that the
test coupons can be flush-mounted inside the apparatus and replaced easily.
The high-speed video camera was placed in front of the screen coupon. Exposure
time was set at 10–20 µs. Depth of field was less than 1 mm. This allows the
examination of velocity in the flow away from walls of the CSS. The framing rate was
set at 20000–54000 frames-per-second (f ps). Thus a fibre moving at 10 m/s in the
field of view would move approximately 0.25 mm between frames and have a blur of
approximately 12 µm, which is less than a fibre diameter. A 1000 W halogen lamp
was set up at the back of the apparatus. The front cover and a small window on
the back side of the CSS were constructed of 25 mm Plexiglas
R
plate. The back
window was covered by a diffuser for better lighting uniformity. All recorded movies
were converted into grey scale images for further analysis by MATLAB
R
. A particle
image velocimetry software (PIVLab) was used to investigate the flow field above
the rough wall. PIVlab is a MATLAB
R
time-resolved particle image velocimetry
(PIV) tool. The high-speed camera allows the use of the PIV technique without the
need for specialized lighting. The present approach does, however, require a very

11
3.2. Experimental Details

high framing rate (> 50000 f ps) resulting in a low image resolution (512–380 pixels).
There is a compromise between the camera frame rate and the image resolution. Due
to noise limitations, pulp fines, which are up to 250 µm in length, were used as seeding
particles. Laser Doppler Velocimetry (LDV) was used to compare the traceability of
pulp fines to Dantec PSP 20 seeding particles by measuring the u component 2 mm
above the coupon with the solid-core rotor. The disagreement between the pulp fines
and the seeding particles was less than 1% for rotor speeds up to 20 m/s.
Concentration tests were conducted with a very low concentration of 0.03% (30
g/100 litres of water) of bleached softwood kraft pulp having an average fibre length
of 2.5 mm. At such low consistencies, pulp suspensions have the same flow properties
as water when in a fully turbulent state [20]. The flow velocities in pulp screens are
generally in the order of 5 m/s and consistency is in the range of 1–3% . Under these
conditions the flow of pulp would be in the turbulent flow regime [17]. Slot velocities
in the range of 0 to 4 m/s were used in the experiments. Table 3.2 summarizes the
test conditions.

Figure 3.2: Schematic diagram of the Cross-Sectional Screen apparatus (CSS).

12
3.2. Experimental Details

Figure 3.3: Close-up of the wire cross-section at the entry to the slot showing char-
acteristic wire dimensions

Table 3.1: Coupon geometry

Contour type Height, h Wire Width, w Slot Width, s


(mm) (mm) (mm)
Low 0.6 3.2 0.15
High 1.2 3.2 0.15

Table 3.2: Experimental test conditions

Test Suspension Rotor Speed, Vt Slot Velocity Vs


(concentration) (m/s) (m/s)
Flow field fibre fines (0.02%) 5 - 20 0-4
Concentration bleached kraft pulp (0.03%) 10 - 20 0-3

13
3.3. Results and Discussion

3.3 Results and Discussion

3.3.1 Steady State Flow Analysis

The flow patterns at the slot entry, including the location of the stagnation point
and height of the exit layer are studied in this section as a function of the contour
geometry, rotor velocity and slot velocity. Figure 3.4 shows the wall geometry for
these tests where a smooth rotor was used to provide a steady circumferential flow.
The effect of flow conditions and contour height on vortex size are shown in Figures
3.5 and 3.6. The figures show a reduction in vortex size as slot velocity increases,
consistent with Gooding [16]. Vortex size also increases with increasing roughness
(i.e. increased contour height). This agrees with Mokamati et al. [35]; however, the
present study suggests a stagnation point closer to the slot entry than they found,
which may be a consequence of differences in the channel geometry and the type of
particles used in the studies. The average velocity between the smooth rotor and the
screen surface, i.e. the upstream velocity, Vu , was found to be proportional to rotor
speed (i.e. Vu ∼ 40% Vt ) but independent of the wall roughness and slot velocity. The
presence of the side walls at the front and back of the CSS slows the fluid rotational
velocity with respect to the foil [12].
Figure 3.7 shows the effect of velocity ratio (i.e. Vs /Vu ) on vortex size. Vortex
size is assessed here by the reattachment distance, d, measured from the edge of the
slot entry to the stagnation point on the wall, as shown in Figure 3.4 and normalized
by the wire slope length, L. As expected, the reattachment distance decreases when
slot velocity increases or rotor speed decreases with the two effects being of roughly
equal significance. Figure 3.8 shows the effect of the velocity ratio (Vs /Vu ) on exit
layer height, H, normalized by the slot width. Exit layer height is defined as the

14
3.3. Results and Discussion

vertical distance from the wall to the exit layer streamline above the wire, i.e. the
upper limit of the flow passing from the adjacent wall through the slot. Accordingly,
higher slot velocities will increase the exit layer height proportionally for a constant
upstream velocity. These results agree with Olson’s [39] and Gooding’s [15] work.
While Figure 3.8 shows a fairly linear effect, this relationship is strongly influenced
by the data points at high Vs /Vu , (i.e. 2) while a typical range of industrial values is
approximately between 0.1 and 0.3.
Circulation, Γ, can be calculated by integrating the vorticity within the vortex
above the slot. For a finite area, circulation divided by area gives the average normal
component of vorticity in the region. Circulation increases with higher slot and
upstream velocities as shown in Figures 3.9 and 3.10. The circulation levels for the
high and low contours were roughly the same, with the lower contour having higher
circulation values in some instances. The significance of circulation in pulp screening
applications has not been conclusively determined, but it may be that higher levels
of circulation tend to encourage the passage of both fibres and contaminants as well
as discouraging the accumulation of fibres at the slot entry.
Appendix A shows the effect of stream-wise and slot velocities on fibre fractiona-
tion.

Figure 3.4: Smooth rotor-wall flow geometry.

15
3.3. Results and Discussion

Figure 3.5: Slot-entry vortex for a high contour at (a) Vt = 10 m/s, Vs = 1 m/s, (b)
Vt = 10 m/s, Vs = 4 m/s, (c) Vt = 20 m/s, Vs = 1 m/s, (d) Vt = 20 m/s, Vs = 4 m/s.

Figure 3.6: Slot-entry vortex for a low contour at (a) Vt = 10 m/s, Vs = 1 m/s, (b)
Vt = 10 m/s, Vs = 4 m/s, (c) Vt = 20 m/s, Vs = 1 m/s, (d) Vt = 20 m/s, Vs = 4 m/s.

16
3.3. Results and Discussion

Figure 3.7: Vortex stagnation point length.

Figure 3.8: Exit layer height (normalized by slot width) versus the ratio of slot velocity
and upstream velocity.

17
3.3. Results and Discussion

Eff t f Fl
Effect of Flow on Circulation 
Ci l ti
6000
Vt = 5 m/s
5000 Vt = 10 m/s
Vt = 15 m/s
4000 Vt = 20 m/s

A v (1/s)
3000

2000

1000

0
0 0.25 1
0.4 10.65 0.8
2 1
2.5
Vs/Vut

Figure 3.9: High contour circulation


MG 1232 within the vortex as a function of flow ratio.

Eff t f Fl
Effect of Flow on Circulation 
Ci l ti
6000
Vt = 5 m/s
5000 Vt = 10 m/s
Vt = 15 m/s
4000 Vt = 20 m/s
A v (1/s)

3000

2000

1000

0
0 0.25 0.4
1 0.6
1.5 0.8
2 1
2.5
5 Vs/Vut

MG0632
Figure 3.10: Low contour circulation within the vortex as a function of flow ratio.

18
3.3. Results and Discussion

3.3.2 Time-Varying Flow Analysis

The rotor foil creates a complex flow field near the wall and is thought to impart both
a time-varying circumferential flow and a pressure pulsation that, in turn, creates a
flow reversal in the slot. Figure 3.11 shows the rotor-wall geometry used for this study.
The foil location is used as a reference for all time-varying data. Figure 3.12 shows the
regions of interests (ROI) where the dynamic behaviour of the velocity components in
the x direction, u, and y direction, v, are averaged, as well as the reference plane where
spatio-temporal values are assessed. The ROI for the v component is chosen above the
slot since velocity components within the slot could not be evaluated experimentally
in this study due to noise and shading effects. While the ROI for the v component
may lie over part of the recirculating zone, it is assumed that the integrated flow
across the entry to the contour corresponds to the flow through the slot.

Figure 3.11: Foil rotor-wall geometry.

Figures 3.14–3.16 show that the u component of velocity drops to approximately


0.3 of the rotor tip speed under the foil (i.e. at xr /chord = 0.2) before it accelerates
under the trailing half of the foil. The highest value is seen at the trailing tip of

19
3.3. Results and Discussion

Figure 3.12: ROI’s for u and v components.

the foil (i.e. one chord length) and the u-component continues at higher than the
average stream-wise velocity for up to three chord lengths. The finding that the
flow decelerates during the passage of the foil, as shown in Figures 3.14–3.16, merits
some discussion since the passage of the foil is also associated with a decrease in
pressure (i.e. a suction pulse). The notion of a decrease in pressure accompanying a
decrease in velocity is counter to an intuitive application of the Bernoulli Equation.
An explanation for this paradox comes through an understanding of the steady and
unsteady forms of the Bernoulli Equation.
From a frame of reference set on the foil with the foil moving at a velocity of u1
and with a circumferential flow velocity u2 as shown in Figure 3.13a, the apparent
velocity approaching the foil is −u3 . This apparent velocity accelerates to a velocity
of −u4 as the flow passes through the restriction between the foil and screen cylinder.
Conservation of mass leads to an increase in velocity of ∆u. Following on the steady
form of the Bernoulli Equation [46], and neglecting elevation affects a pressure drop
is predicted according to the following equation,

u24 − u23
(P3 − P4 ) = ρ (3.1)
2

20
3.3. Results and Discussion

Figure 3.13: The flow between the foil and cylinders surface are examined for two
alternate frames of reference: (a) a steady flow problem follows from fixing the frame
of reference on the foil, which is moving at u1 . The flow between the foil and cylinder
surface increases by an amount ∆u. (b) An unsteady flow situation results from fixing
the frame of reference to the screen cylinder. The apparent velocity under the rotor
is decreased by an amount ∆u relative to the general flow velocity.

With the same flow conditions, but with the frame of reference set at a location
on the screen cylinder, the apparent velocity under the rotor can be seen to decrease
from u2 to u5 as shown in Figure 3.13b. While the pressures remain the same (i.e.
P3 = P2 , P5 = P4 ), the differences between the under-foil velocities result from the
presence of the integral term in the unsteady form of the Bernoulli Equation

5
P2 u22 P5 u25
Z
∂u
+ + gZ2 = + + gZ5 + ds (3.2)
ρ 2 ρ 2 2 ∂t

As discussed previously, values of the u component in Figures 3.14–3.16 are based


on the averages made over the ROI shown in Figure 3.12, where the u component
ROI is located at a distance of 0.5 mm from the cylinder surface (i.e. y/Hr = 0.25).
A different perspective on the same dataset is given in Figure 3.17, which shows the
changes in the velocity vector field as a function of rotor position. These vector images

21
3.3. Results and Discussion

reinforce what was shown in Figures 3.14–3.16, with the velocity being relatively low
as the rotor passes (xr /chord = 0.1) but then rising in the foil wake (xr /chord =
1.25). It is useful to understand variations of the u component as a function not
only of xr , but also for a range of distances from the cylinder surface. The power
of the PIV technique used in this study is that it can be used to create databases
that can generate not only vector plots (such as those shown in Figure 3.17) or plots
describing the variation in an averaged flow value (e.g. Figures 3.14–3.16), but the
PIV data can also yield the full spatio-temporal description of the flow field, as shown
in Figure 3.18(a). In these spatio-temporal plots, the magnitude of the u component
is expressed chromatically not only as a function of xr , but for the full range of y =
0 to Hr . In particular, the value of the u component is assessed in Figure 3.18(a) at
a series of y values along the reference plane shown in Figure 3.12. To reduce the
measurement noise, averages are taken in the xr -direction over a relatively narrow (0.2
mm wide) band for each value of y. To illustrate the connection between the spatial-
temporal and line graphs, Figure 3.18(a) and 3.18(b) present two perspectives on the
same dataset. Figure 3.18(a) also includes an illustration of the averaging volume used
for the purpose of generating Figure 3.18(b) consistent with the methodology shown
in Figure 3.12. Figure 3.18(b) is the result of the averaging process, comparable to
Figure 3.15, albeit over a smaller range of xr /chord. The finding remains the same:
illustrating a fairly constant value of the u component of velocity ahead of the foil
(i.e. for negative values of xr /chord), then a drop in velocity with the passage of
foil and acceleration in the wake. Figures 3.19-3.22 show similar spatio-temporal
plots of the u component (normalized by Vt ) for a range of rotor tip and average slot
velocities. In general, the value of u/Vt is seen to drop to below 0.3 in the gap under
the rotor, consistent with Figures 3.14–3.16. These figures also show the acceleration

22
3.3. Results and Discussion

(i.e. increasing u) that takes place under the trailing half of the rotor foil (xr /chord
between 0.5 and 1). The degree of acceleration near the wires (i.e. y/Hr < 0.1) is
smaller in comparison to the flow behind the rotor in the mainstream where suction
has less effect on the stream-wise flow.
Changes in the v component of velocity for different rotor and slot velocities
and high and low contours are shown in Figures 3.23–3.24. These figures show that
reversal flows (i.e. periods of positive v component values) are supported with high
rotor speeds and low average slot velocities. For the high contour and Vt = 20 m/s
(Figure 3.23), the reversal flow starts at xr /chord ∼ 0.5 for Vs = 1 m/s and lasts
for ∼ 1.5 chord lengths. When the slot velocity is increased to 4 m/s, the reversal
flow is relatively unchanged in absolute terms, but is naturally much smaller when
instantaneous v component is normalized by the average slot velocity, Vs . When rotor
speed is reduced (Figure 3.24), the flow reversal is seen to disappear when slot velocity
is increased from 1 to 4 m/s. The implication for an industrial screen application is
that plugging of the slots may become more problematic when rotor speeds are low
and slot velocities are high, as is commonly reported. Note that the v component is
applied through a ROI that is 2.75 mm wide versus the 0.15 mm slot width where Vs
is calculated. Thus a Vs of 4 m/s corresponds to an average value of v = - 0.22 m/s,
which is consistent with the values seen in Figures 3.23 and 3.24.
The existence of a pressure (suction) pulse associated with the passage of a rotor
foil has been shown in many studies ([45], [14], [27], [7], [37]). Theoretical studies,
using computational fluid dynamics ([11], [12], [50], [10]) have also shown that the
acceleration of flow under the foil gives rise to a suction pulse, but these studies
were done either with a solid or highly-simplified model of the porous cylinder wall.
In general, most of these studies have associated pressure pulse signature with an

23
3.3. Results and Discussion

immediate flow reversal without taking into consideration the conservation of a con-
fined mass passing through the slot between the rotor and the screen surface, which
could explain the discrepancy with the results shown in Figures 3.23–3.24, where the
backflush flow is delayed slightly.
Based on the above results, Figure 3.25 shows the cases where reversal flow occurs.
It was experimentally shown by Gonzalez [14] that the magnitude of pressure pulses
decreases as the consistency increases. Moreover, the presence of the side walls at the
front and back of the CSS slows the rotational velocity of the fluid relative to the foil.
The increased relative velocity would account for higher pressure pulses [12]. Given
the low consistency used in these tests and the lower fluid velocity relative to the foil,
this figure represents the largest possible backflushing region for this particular rotor.
An investigation of the velocity changes along the contour surface is also of interest.
A velocity, Vd , along the contour surface was defined in Figure 3.11 and measured as
a function of rotor position for both contours at different flow conditions as shown in
Figures 3.26 and 3.27. Of particular interest is the location of the stagnation point
(Vd = 0) which is relatively close to the slot entry (xd /L ∼ 0.2) for both contour
heights and does not move significantly with the passage of the rotor – even when
there is flow reversal.
The uncertainty of the PIV measurements is mainly caused by the technique
random error and the setup systematic error. The highest relative uncertainties of
the velocity above the wires and within the vortex with 95% confidence intervals
were ±6.6% and ±16.4%, respectively. The higher uncertainty within the vortex
was expected as rotation and shear can be a large source of uncertainty in PIV
measurements [47]. Details of PIV uncertainty analysis is shown in Appendix B.

24
3.3. Results and Discussion

Figure 3.14: u component for low contour at Vt = 15 m/s and Vs = 1 m/s.

Figure 3.15: u component for high contour at Vt = 20 m/s and Vs = 1 m/s.

Figure 3.16: u component for high contour at Vt = 20 m/s and Vs = 2 m/s.

25
3.3. Results and Discussion

Figure 3.17: Rotor position in terms of chord length.

26
3.3. Results and Discussion

Figure 3.18: The spatio-temporal description of the u-component in the flow field
for the high contour (Vt = 20 m/s, Vs = 1 m/s) is shown above (a) along with the
measuring volume (in yellow) used for averaging. The lower plot (b) is the associated
averaged flow for the u-component.

27
3.3. Results and Discussion

1 1

0.8 0.8

0.6 0.6
y / Hr

u/Vt
0.4 0.4

0.2 0.2

0 0
−1 0 1 2 3 4 5
xr / chord

Figure 3.19: Spatio-temporal behaviour of u component for high contour surface at


Vt = 10 m/s, Vs = 1 m/s.

1 1

0.8 0.8

0.6 0.6
y / Hr

u/Vt
0.4 0.4

0.2 0.2

0 0
−1 0 1 2 3 4 5
xr / chord

Figure 3.20: Spatio-temporal behaviour of u component for high contour surface at


Vt = 10 m/s, Vs = 3 m/s.

28
3.3. Results and Discussion

1 1

0.8 0.8

0.6 0.6
y / Hr

u/Vt
0.4 0.4

0.2 0.2

0 0
−1 0 1 2 3 4 5
xr / chord

Figure 3.21: Spatio-temporal behaviour of u component for high contour surface at


Vt = 20 m/s, Vs = 1 m/s.

1 1

0.8 0.8

0.6 0.6
y / Hr

u/Vt
0.4 0.4

0.2 0.2

0 0
−1 0 1 2 3 4 5
xr / chord

Figure 3.22: Spatio-temporal behaviour of u component for high contour surface at


Vt = 20 m/s, Vs = 3 m/s.

29
3.3. Results and Discussion

0.8
0.6 Vt= 20 m/s, Vs = 1 m/s
0.4 Vt= 20 m/s, Vs = 4 m/s
0.2
v (m/s)

0
-0.2
-0.4
06
-0.6
-0.8
-1 0 1 2 3 4 5 6 7 8 9 10
xr/chord
Figure 3.23: v component for high contour at Vt = 20 m/s.

0.88
0
0.6
06 Vt= 10 m/s
V m/s, V = 1 m/s
/ , Vs /
0.4
04 Vt= m/s,
V 10 m/s V = 4 m/s
/ , Vs /
02
0.2
0
02
-0.2
-0
04
0.4
06
-0
-0.6
-0.8
08
0.8
-11 0 1 2 3 4 5 6 7 8 9 10
xr/chord
/ h d
Figure 3.24: v component for low contour at Vt = 10 m/s.

30
3.3. Results and Discussion

Figure 3.25: Reversal flow boundary.

31
3.3. Results and Discussion

1 0.6

0.8
0.4

Vd / Vu
0.6
xd / L

0.2
0.4
0
0.2

0 −0.2
−1 0 1 2 3 4 5
xr / chord

Figure 3.26: Spatio-temporal behaviour of Vd for low contour surface at Vt = 10 m/s,


Vs = 1 m/s.

1 0.6

0.8
0.4

Vd / Vu
0.6
xd / L

0.2
0.4
0
0.2

0 −0.2
−1 0 1 2 3 4 5
x / chord
r

Figure 3.27: Spatio-temporal behaviour of Vd for high contour surface at Vt = 20 m/s,


Vs = 1 m/s.

32
3.3. Results and Discussion

3.3.3 Fibre Concentration Analysis

The dynamic change of pulp concentration near the screen surface was also studied
for dilute (0.03%) pulp suspensions. In contrast to the PIV studies, kraft pulp fibres
rather than pulp fines were used for the fibre concentration tests. As a first step,
the relationship between fibre concentration and light intensity was determined by
measuring the average light intensity over 1000 frames for an area away from the slot
entrance. At this remote location, the local fibre concentration could be assumed to
be equal to the feed concentration. For calibration purposes, a smooth rotor was used
and the tip speed was set at 10 m/s with no flow through the slot. Figure 3.28 shows
the associated calibration curve, with a linear relationship between light intensity
and concentration. A foil rotor was then introduced to study the variations in fibre
concentration. Two regions of interest were chosen for concentration analysis: The
first region was between the rotor and the screen surface (”above screen” region), and
the second region is in the discharge flow from the slot (”slot discharge” region) as
shown in Figure 3.29.
Concentration as a function of flow for the high contour screen coupon is shown in
Figure 3.30. Local concentration values were normalized by the average concentration
of the region, Co , above the screen surface at Vt = 10 m/s and Vs = 0 m/s. A number
of interesting observations can be made: First, one can see that the ”above screen”
concentration is slightly (10%) above the mainstream concentration and this effect is
seen fairly consistently for various Vt and Vs . This may be because of the withdrawal
of fluid through the slotted surface with a lower concentration of fibres. The fibres
that do not pass through the slots enrich the flow adjacent the screen surface slightly.
The concentration in the slot exhaust region shows more complex influences. At
the lower tip speed (10 m/s), the slot exhaust flow is slightly depleted of fibres at

33
3.3. Results and Discussion

R² = 0.994
Normaalized Light Intensity 0.8

0.6

0.4

0.2

0
0.05 0.1 0.15 0.2 0.25
Fiber concentration (g/litre)

Figure 3.28: Light intensity as a function of fibre concentration

Figure 3.29: High contour concentration ROI’s

34
3.3. Results and Discussion

the lowest slot velocity (1 m/s), as one would expect from the aforementioned ”wall
effect”. Likewise, it follows from various fundamental screening studies [[18], [31],
[41]] that concentration should increase with slot velocity, which is seen here. At the
highest slot velocity (3 m/s; Figure 3.30(c) the concentration was found to be higher
than the ”above screen” concentration, which suggests that the higher slot velocity
has eliminated the influence of the wall effect and fibre trapping may have boosted
the consistency within the slot entrance.
At the higher, and more typical, rotor speed (20 m/s) the effects are similar,
as shown in Figure 3.30(d)–3.30(f), but a low-consistency zone of that is relatively
deplete of fibres is now apparent one to two chord lengths behind the rotor foil, which
may be an important and beneficial factor in ensuring reliable screen capacity. This
depletion may occur because of larger flow structures in the wake of the foil combined
with some degree of backflushing through the slot.

35
3.3. Results and Discussion

1.25 1.25
(a) (d)
1 1

0.75 0.75
−1 0 1 2 3 4 5 6 7 8 9 −1 0 1 2 3 4 5 6 7 8 9
1.25 1.25

1 1
(b) (e)
0.75 0.75
−1 0 1 2 3 4 5 6 7 8 9 −1 0 1 2 3 4 5 6 7 8 9
1.25 1.25
o
C/C

1 1
--- Above screen ROI
(c) --- Slot exhaust ROI (f)
0.75 0.75
−1 0 1 2 3 4 5 6 7 8 9 −1 0 1 2 3 4 5 6 7 8 9
xr/chord
Figure 3.30: Concentration for high contour: (a) Vt = 10 m/s, Vs = 1 m/s, (b) Vt =
10 m/s, Vs = 2 m/s, (c) Vt = 10 m/s, Vs = 3 m/s, (d) Vt = 20 m/s, Vs = 1 m/s, (e)
Vt = 20 m/s, Vs = 2 m/s, (f) Vt = 20 m/s, Vs =3 m/s.

36
3.4. Summary and Conclusions

3.4 Summary and Conclusions

The flow field near a rough slotted surface in steady and time-varying cross-flow was
studied experimentally for different degrees of roughness and different upstream and
aperture velocities. While this problem is of general interest, the present study was
made in the context of the flow of a fibrous suspension in a pulp screen. Particle-
image velocimetry was used in this study to measure the flow patterns near the
slotted surface. The size of the vortices within the roughness contours was found
to be dependent on the geometry of the contours and the flow velocities. Vortices
increase in size with lower aperture velocities and higher contour heights. This study
also considered the flow through slots in the slotted surface. The height of the exit
layer that passes from the main flow through the slotted apertures was found to
increase with slot velocity.
For the dynamic studies, the stream-wise flow decelerates under the leading edge
of the rotor before it increases under the trailing half of the rotor then gradually drops
to an average value that is independent of slot velocity. A flow reversal was observed
in some, but not all, of the flow configurations.
Studies of fibre concentration showed that the concentration within the contours
was generally (but not always) lower than in the flow above the screen surface. This
is consistent with published mechanisms of fibre motion near a bifurcating flow. A
zone with reduced fibre concentration was observed in the wake of the rotor foil. This
may be industrially significant in avoiding the blockage of the apertures with fibres,
though the reversal flow through the slot and turbulent flows in the wake of the foil
are also believed to be important in avoiding aperture blockages.
This study provides a number of insights into this complex and important flow
problem of unsteady flow adjacent to a rough slotted wall. It builds on previous

37
3.4. Summary and Conclusions

studies to elucidate mechanisms related to the flow resistance of the flow through
the slotted wall and the potential for the apertures in the wall to be obstructed with
fibres carried in the flow.

38
Chapter 4

Maximum Capacity of a Pilot


Pressure Screen

4.1 Introduction

The objective of the work presented in this chapter is to investigate the factors af-
fecting the capacity of fibre suspension pressure screens. The capacity of the screen
is defined in this work as the maximum throughput before the apertures permanently
plug with pulp. Throughput, in turn, is represented by slot velocity, Vs , which is
the key process variable driving throughput. Maximum slot velocity is a function
of feed fibre length, feed consistency, screen plate aperture size and geometry, rotor
speed and rotor type. Although many studies have been conducted to understand the
factors affecting the performance of pulp screens, only a few were focused on factors
affecting the capacity of screening,([34], [17], [33], [25], [22]).
Niinimaki [36] categorized the matting of fibres on the inner surface of the screen
and pulsations by the foils as macro phenomena, and the hydrodynamic forces ex-
erted on the fibres and fluidization due to turbulence on the screen surface as micro
phenomena. Halonen et al. [21] defined fluidization as a process of loosening fibre
contacts inside flocs. Fibre suspension can display fluid-like behavior that may justifi-
ably be called fluidized under certain circumstances. This occurs in turbulent regime

39
4.1. Introduction

and, in this state, the suspension can be assigned a viscosity and otherwise follows
the laws of fluid mechanics [2]. By using contoured screen cylinders, the turbulence
conditions on the screen surface are improved and fluidization can extend the effect
induced by the rotor, [13]. It is widely believed that rotor pulses disrupt the fibre
mat and lift trapped particles away from the screen apertures, [36].
Martinez et al. [33] introduced a screen capacity model by assuming that screen
blinding limits volumetric capacity and that blinding occurs when a fibre floc becomes
immobilized in a screen slot. The essential analysis was a force balance on a single
floc in a slot with the forces arising from the flow through the slot, friction of the floc
against the slot wall, and the rotor pulsation. However, this model does not embrace
many of the complexities of screen operation such as fibre trapping and rotor wake
effects.
Hamelin et al. [22] conducted a series of pilot screening trials for a range of foil
configurations. For each rotor configuration tested, the maximum slot velocity-power
curve was determined by setting the rotor speed to a constant value and increasing
the slot velocity until the onset of plugging. The point of maximum slot velocity
versus rotor power, creates a failure envelope (or maximum slot velocity envelope) for
each rotor and is used to compare the performance of various configurations.
The mechanism of plugging is still not clear. This study examines the factors
affecting maximum capacity from a macroscopic level and theorises a simple plugging
model that can be used in future work to understand screen capacity.

40
4.2. Experimental Details

4.2 Experimental Details

This work studies the effect of screen cylinder geometry and pulp properties on ca-
pacity. The experiments were conducted at the Pulp and Paper Centre, UBC with
a Beloit MR8 laboratory pressure screen, shown in Figure 4.1. The MR8 is 212 mm
in diameter and is equipped with a variable frequency drive (VFD) to control the
rotor speed up to 20 m/s. An AFT EP foil rotor was used for all trials and the rotor-
cylinder gap was kept at 4 mm. The pulp is supplied from a 1500 L tank through a
feed port. The reject and accept streams are returned to the feed tank. Pulp tem-
perature was maintained constant at a nominal value of 40 o C by the virtue of a 5
kW electric heater attached to the feed tank. The pulp feed consistency, CF , was also
monitored and kept constant throughout every test. To prevent the reject line from
plugging, the reject volumetric rate, Rv , for most tests was fixed at 20% .
Five screen cylinder designs which differed in slot width and wire type were tested.
Table 4.1 shows the cylinders and their geometric configurations. The cylinder open
area varies greatly with the slot and wire widths, which then calls for large changes in
pumping rates to maintain the specific slot velocity. Due to limited pumping capacity
of the loop, the cylinders’ open area was modified by blinding part of the cylinder’s
open area. Also, it was found in preliminary tests that the first segment of the cylinder
was plugged even when all other segments were clear. This is thought to be because
of insufficient acceleration of the tangential velocity of the pulp adjacent to the screen
cylinder and that the pulp is not fluidized, [1]. Blinding of the initial segment of the
cylinder is also thought to reduce the influence of this artefact. Figure 4.2 shows
the blinding procedure. To minimize the effect of feed flow on screen performance,
the difference in open area for different cylinders was minimized by blinding the first
and last spools for most cylinders (Type A-D) and the first spool only for the Type

41
4.2. Experimental Details

Figure 4.1: The UBC MR8 pilot pressure screen.

E cylinder. The number of blinded spools was limited to two in order not to affect
the thickening behaviour over the cylinder’s axial length. The effects of consistency,
reject rate and contaminant content were also studied in these tests. The pulp used
for most tests was a mix of 50/50 hardwood/softwood kraft pulp to simulate a de-
ink, kraft pulp [26]. The ratio of hardwood/softwood was also changed to study the
impact of fibre length on maximum capacity.
Trials were conducted by reducing the rotor speed, Vt , at constant slot velocity,
Vs , until the screen plugged. The slot velocity was then increased and the procedure
was repeated. A plugging envelope for the minimum rotor speed was generated for
different cylinders, pulp types and consistencies. It was found that the minimum
rotor speed needed to prevent plugging was repeatable within ± 0.2 m/s and the
error associated with the measured values was small enough to be neglected.

42
4.2. Experimental Details

Table 4.1: Screen wire geometries

Cylinder Designation Contour Height Wire Width Slot Width Open Area
(mm) (mm) (mm) (m2 )
A 1.2 3.2 0.15 0.00453
B 0.9 3.2 0.15 0.00453
C 0.6 3.2 0.15 0.00453
D 0.6 2.3 0.15 0.00619
E 0.6 3.2 0.10 0.00482

Figure 4.2: Screen open area blinding procedure: (a) wrapping the spool with rubber
strip and (b) securing the strip with a stainless steel clamp to prevent leakage from
the spool edges.

43
4.3. Results and Discussion

4.3 Results and Discussion

4.3.1 Effect of Consistency on Capacity

The effect of pulp consistency is discussed in this section. Figure 4.3 shows the plug-
ging envelope for the Type B cylinder. This figure merits special attention inasmuch
as it illustrates the essential topic of this thesis study. The zone to the right of the
line represents the normal operating zone of a pulp screen, while plugging occurs to
the left. The dividing line shown in the figure thus represents screen capacity for a
particular furnish and screen configuration. As pulp consistency increases, the rotor
speed needed to prevent plugging at the same slot velocity also increases. At 2.0%
consistency, tests were conducted only at slot velocities greater than 1.5 m/s due to
reject line thickening issues. This graph also reinforces some valuable findings:

• Linearity: The relationship between Vs and Vt is remarkably linear, which leads


to the proposal of a simple capacity equation of

Vs∗ = C1 + C2 Vt (4.1)

This Equation suggests that plugging occurs when Vs > Vs∗ . C1 and C2 are
characteristic capacity constants.

• Parallelism: Another significant feature of Figure 4.3 is that the slopes of the
lines, C2 , are relatively similar and difference occurs mainly on the offset of
these lines, C1 , as shown in Figure 4.4. Indeed if one plots the x-intercepts of
the lines, (−C1 /C2 , which is equal to Vt∗ ), as a function of consistency (Figure
4.5), one is reassured by the suggestion that at 0% consistency, there is no offset.

44
4.3. Results and Discussion

• Threshold Rotor Speed: It follows from Figure 4.5 that a minimum rotor speed,
Vt∗ , must be attained to obtain any significant capacity. Moreover, a more
challenging application, (i.e. a high consistency, increased presence of long
fibre, narrow slots) would require a stronger / more frequent rotor action and
higher threshold rotor tip speed.

By comparing the reversal flow envelope, Figure 3.25 with the plugging envelope
for the Type B screen cylinder, one can see that the screen is running at a region
where no reversal flow is present, suggesting other mechanisms that maintain the
screen operation. For example, at CF = 1% and Vs∗ = 2 m/s, Vt ∼ 8.5 m/s in Figure
4.3 while Figure 3.25 suggests reversal occurs for Vt between 10 and 15 m/s. While
the screens are not the same, the reversal flow pressure pulses generated by the CSS
rotor are in fact greater than those generated by the MR8 rotor running at the same
speed for the following reasons:

• Smaller foil-screen surface gap for the CSS tests.

• Higher angle of attack for the CSS foil (0o for the MR8 rotor).

• Water was used for the reversal flow test but the plugging envelope zone was
produced with 1.0% pulp.

• The presence of front and back walls in the CSS tests reduced the relative
velocity of the fluid with respect to rotor speed, creating higher pressure pulses
for its rotor.

It is also found that the pressure difference across the screen, i.e. the pressure
measured in the feed line to the screen minus the pressure in the accept line, increases
with slot velocity and pulp consistency, as shown in Figure 4.6.

45
4.3. Results and Discussion

4.5
C = 0.5% CF= 1.0%
F
4 CF = 1.5% CF = 2.0%
3.5
3 Plugging Zone
Vs (m/s) 2.5
2
1.5
1 Operational Zone
0.5
0
4 6 8 10 12 14 16
Vt (m/s)
Figure 4.3: Plugging envelope for Type B cylinder at different consistencies, (50/50
hardwood/softwood mixture).

4.5
C = 0.5%
4 F
CF = 1.0%
3.5 CF = 1.5%
3 CF = 2.0%
Vs (m/s)

2.5
2
1.5
1
0.5
0
0 2 4 6 8 10 12 14 16
Vt (m/s)
Figure 4.4: The extrapolated Vs -Vt relationship of Figure 4.3.

46
4.3. Results and Discussion

9
8 y = 4.37x - 0.075
R² = 0.9987
7
6

Vt* (m/s)
5
4
3
2
1
0
0 0.5 1 1.5 2
CF (%)
Figure 4.5: The x-intercepts of the extrapolated Vs -Vt relationship of Figure 4.4 is
plotted against consistency suggesting that there would be no offset when there are
no fibres in the suspension (CF = 0%).

100
water
CF = 0.5%
80
CF = 1.0%
CF = 1.5%
60 CF = 2.0%
DP (kPa)

40

20

0
0 1 2 3 4 5
Vs (m/s)
Figure 4.6: Pressure difference across Type B screen cylinder prior to plugging at a
range of slot velocities (50/50 hardwood/softwood mixture).

47
4.3. Results and Discussion

5
4.5 Vs = 1 m/s
4 Vs = 4 m/s

3.5

Power (kW)
3
2.5
2
P

1.5
1
0.5
0
0 5 10 15 20 25
Vt (m/s)
Figure 4.7: Power consumption by the rotor as a function of rotor speed at different
slot velocities (CF = 0%).

4.3.2 Effect of Feed Flow Rate on Capacity

The first trial to evaluate the effect of flow rate on power consumption was conducted
with water. Figure 4.7 shows the power consumed by the rotor for different slot
velocities. Higher slot velocities are naturally associated with higher feed rates at
the same reject rate. Higher feed rates increase the tangential velocity within the
screen and reduces the power consumed by the rotor. However, a higher tangential
flow, which is created here with higher Rv , also reduces the relative flow velocity with
respect to rotor speed causing reduced pulse strength and plugging of the cylinder at
a higher Vt , as shown in Figure 4.8. This behaviour was previously observed by Delfel
et al. [9]. They showed that a three-foil rotor reduces the capacity of a cylinder with
respect to two-foil rotor as it increases the circumferential (swirl) flow velocity with
respect to rotor speed.

48
4.3. Results and Discussion

4
Rv = 0.15
3.5
Rv = 0.25
3 Rv = 0.35

2.5 Rv = 0.45
Vs (m/s)
2

1.5

0.5

0
4 6 8 10 12 14
Vt (m/s)
Figure 4.8: Effect of reject rate on plugging envelope for the Type B screen (100%
softwood suspension with 1.0% consistency).

4.3.3 Effect of Pulp Properties and Contaminants on

Capacity

The influence of pulp properties was assessed by mixing softwood and hardwood
kraft pulp in different ratios (see Appendix C for more details). Figure 4.9 shows the
plugging envelope for the Type B cylinder. The figure shows the linearity between
rotor speed and slot velocity. Average fibre length is seen to have a large impact on
the plugging envelope. As the average fibre length increases, (i.e. as the softwood
ratio increases) the rotor speed needed to prevent plugging becomes higher. Figure
4.10 shows the pressure difference across the cylinder immediately prior to plugging.
It is significant that for the same slot velocity, the pressure difference across the
screen does not change with pulp type; nevertheless, the rotor speed needed to prevent
plugging does increase with increasing softwood ratio. This indicates that an increased

49
4.3. Results and Discussion

negative pressure pulse is required for increasing softwood concentrations which may
be because the softwood creates stronger incipient blockages (softwood has a higher
yield stress [6]). For example, at 3 m/s slot velocity and 100% softwood pulp, the
negative pressure pulse associated with the rotor minimum speed is 45% greater
than the negative pulse needed at the same slot velocity but with 100% hardwood
suspension.
The effect of contaminants on capacity was also studied given that contaminants
levels can be high, for example, in recycled pulp. Mill debris, i.e. heavily contam-
inated material gathered from the rejects of third stage of OCC screen, was added
to a 1.5% 50/50 hardwood/softwood pulp blend and the minimum rotor tip speed to
maintain reliable screen operation was assessed. The mass of the debris with respect
to the mass of the pulp was gradually increased, but no significant effect was found
on capacity. The minimum rotor tip speed was, however, sensitive to the addition
of cubical polyethylene specks but even a very high (2%) loading level led only to a
modest increase in rotor speed (Figure 4.11). One explanation for the increased Vt
is the reduction in cylinder open area which, in turn, increases the slot velocity and
causes early plugging. Figure 4.12 shows an increase in pressure difference with an
increased specks concentration, which results from the higher slot velocity and smaller
open area. The other explanation is a faster accumulation of fibres on the plugged
specks that leads to faster plugging. The specks ranged in size, but were mostly in
the range of 0.05 to 0.3 mm in nominal diameter, as shown in Figure 4.13.

4.3.4 Effect of Cylinder Geometry on Capacity

This section focuses on the effect of screen cylinder geometry on maximum capacity.
Figure 4.14 shows the effect of contour height on the plugging envelope for three

50
4.3. Results and Discussion

5
4.5 (0:100) SW:HW
(25:75)
4
(50:50)
3.5 (75:25)
3
Vs (m/s)
(100:0)
2.5
2
1.5
1
0.5
0
4 6 8 10 12 14 16
Vt (m/s)
Figure 4.9: Plugging envelope for the Type B screen using different soft-
wood/hardwood ratios, CF =1.0%.

100
(0:100) SW:HW

80 (25:75)
(50:50)
(75:25)
60
P (kPa)

(100:0)

40

20

0
0 1 2 3 4 5
Vs (m/s)
Figure 4.10: Pressure difference across the Type B cylinder for different soft-
wood/hardwood ratios, CF = 1.0%.

51
4.3. Results and Discussion

16

14

12

10
Vt (m/s)
8

4 Mill Debris
Plastic specks
2

0
0 2 4 6 8 10 12 14
mass of debris/total solid mass (%)

Figure 4.11: Minimum rotor tip speed for the Type B screen at different contaminant
ratios (50/50 hardwood/softwood mixture, CF = 1.5%) Vt = 2 m/s.

40

35

30

25
P (kPa)

20

15

10
Mill debris
5 Plastic specks

0
0 2 4 6 8 10 12 14
mass of debris/total solid mass (%)
Figure 4.12: Pressure difference across the Type B screen at different contaminant
ratios, (50/50 hardwood/softwood mixture, CF = 1.5%.

52
4.3. Results and Discussion

Figure 4.13: Plastic speck diameter distribution.

different cylinder types. As reported in industry, increased contour height increases


capacity at lower slot velocities, but the benefit of increased contour height appears
to disappear, or at least to be substantially diminished, in the range of industrial
interest (i.e. above 12 m/s). This lack of benefit is somewhat surprising, but it may
be that the contour height is of significant benefit at low rotor speeds (as shown
in Figure 4.14) and higher consistencies due to increased turbulence that results in
the flow field above the cylinder surface that the relative importance of this effect
diminishes with higher slot velocities. The overall pressure difference was not found
to change with contour height as shown in Figure 4.15.
Tests also were conducted with different cylinder slot and wire widths. The plug-
ging envelopes for Types D and E cylinders are compared with Type B are shown
in Figure 4.16. The open area for Type D cylinder is larger than other cylinders
which requires higher feed flow rate for the same slot velocities. This figure shows

53
4.3. Results and Discussion

4.5
4 Type A
Type B
3.5
Type C
3

Vs (m/s)
2.5
2
1.5
1
0.5
0
4 6 8 10 12 14 16
Vt (m/s)
Figure 4.14: Effect of contour height on the plugging envelope. The contour heights
of the Type A, B and C screen cylinders is 1.2 mm, 0.9 mm and 0.6 mm, respectively.
(50/50 hardwood/softwood mixture, CF = 1.5%).

90
Type A
80
Type B
70 Type C
P (kPa)

60

50

40

30

20
0 1 2 3 4 5
Vs (m/s)
Figure 4.15: Pressure difference across cylinders, (50/50 hardwood/softwood mixture,
CF = 1.5%).

54
4.3. Results and Discussion

4.5
4
3.5
3

Vs (m/s)
2.5
2
1.5
Type E
1 Type D
0.5 Type B

0
4 6 8 10 12 14 16
Vt (m/s)
Figure 4.16: Effect of wire width and slot width on plugging boundary, (50/50
hardwood/softwood, CF = 1.0%). Type B: 0.9 mm contour height/3.2 mm wire
width/0.15 mm slot width. Type D: 0.6 mm /2.3 mm/0.15 mm. Type E: 0.6 mm
/3.2 mm/0.10 mm.

that relative to the Type B cylinder, the shallower contour and narrower wire of the
Type D cylinder leads to reduced capacity. Looking back at Figure 4.14 one sees that
the contour height has limited influence on capacity. The impact of wire width may
come from a narrower slot ”pitch” (i.e. the distance between slots). This is likely due
to the presence of the softwood pulp and potential for slot-to-slot stapling given that
over 60% of the fibre mass are candidates for stapling, which is far above the levels
at which stapling is expected to occur [17].
The effect of the narrow slot is seen for the Type E cylinder. Capacity is substan-
tially reduced, which is likely a consequence of fewer fibres being required to fill the
slot to create a blockage.

55
4.4. Summary and Conclusions

4.3.5 Effect of Rotor and Slot Velocity on Reject

Thickening

Rotor and slot velocities have a significant impact on accept and reject consistencies.
To study this effect, reject and feed samples were collected for the Type A screen
cylinder at different slot velocities to evaluate thickening behaviour. Thickening fac-
tor, T , is defined as the reject consistency divided by the feed consistency. Sample
collection started at a higher rotor speed and stopped just before screen plugging.
Figure 4.17 shows the thickening behaviour of 1.0% feed consistency pulp as rotor
speed was reduced for two slot velocities. Thickening factor is fairly constant at higher
rotor speeds but below ∼ 12 m/s, T starts to increase linearly with reduced rotor
speed. It may be that at higher rotor speeds, the reversal flow dilutes the pulp above
the screen cylinder. When the rotor speed decreases, water reversal decreases and
higher consistency results. It may also be that at lower rotor speeds, the incipient
blockages are not completely cleared by the rotor, resulting in an effectively narrower
slot and reduced fibre passage. Higher slot velocity also means more fibre passage
through the screen slots as reported in the fundamental studies of Gooding [15] and
Kumar [31]. Due to reject line thickening, this test could not be easily conducted
with higher consistencies or lower slot velocities.

4.4 Summary and Conclusions

Pilot screen experiments with different screen cylinders and pulp combinations were
conducted to show the effect of common industrial variables on the maximum capacity
of a pressure screen.
Feed consistency and fibre length were found to have very significant influences

56
4.4. Summary and Conclusions

2.4

2.2 Vs = 1 m/s
Vs = 2 m/s
2

1.8
T
1.6

1.4

1.2

1
6 8 10 12 14 16
Vt (m/s)
Figure 4.17: Thickening of pulp with rotor and slot velocity changes (Type A screen,
CF = 1.0%, 50:50 SW:HW).

on capacity. Contaminants had a very limited effect. A higher contour was found to
delay plugging but this effect was limited to low slot velocities. Slot width and wire
width had significantly more influence on capacity.
The results leads to an appreciation of two regimes of screen operation: (1) a
plugging zone, below the threshold rotor speed, Vt∗ and (2) an operational regime,
where there are appreciable slot flows with appreciable fibre concentrations. The
boundary between these two regimes is relatively linear and is described by threshold
values of Vs∗ and Vt∗ (Equation 4.1).
A particular cylinder and pulp combination led to some estimates of the constants
leading to the equation:
Vs∗ = 0.5Vt − 2.2CF (4.2)

Valid for 0.5 ≤ C ≤ 2%, 0 ≤ Vs ≤ 4 m/s and 0 ≤ Vt ≤ 16 m/s.

57
4.4. Summary and Conclusions

Future studies may be directed to a more comprehensive examination of the con-


stants of Equation 4.1 as well as the influence of different rotor types. The linear form
of Equation 4.1 suggests a relatively simple fundamental model may drive capacity.

58
Chapter 5

New Model for Pressure Screen


Capacity

5.1 Introduction

The capacity of a pulp screen is defined as the maximum throughput before the aper-
tures become plugged with pulp. Capacity is a function of fibre length, consistency,
aperture size and geometry, rotor speed and rotor type ([34], [17], [33], [25], [8]).
Various experimental and computational studies have been conducted to understand
the performance of pressure screening in terms of fibre passage and debris removal
efficiency. Relatively little is known, however, about the fundamental aspects of ca-
pacity.
A new approach is introduced here to understand the maximum capacity of a pulp
pressure screen. The model is based on the concept of trapped fibres accumulating
on the cylinder slot edge which in turn leads to plugging. The lodged fibres on
the slot edge continue to build up when there are no changes in the tangential flow
field above the cylinder surface. These changes delay the accumulation for a certain
period of time of the rotor cycle. If the accumulation continues beyond a certain
threshold value, cylinders plug immediately. The model introduced complements the
effect of the reversal flow on fibre accumulation. Indeed while this work defines and

59
5.2. Mathematical Model Details

analyses some fundamental mechanisms that govern capacity, a comprehensive model


and rigorous validation remain beyond the scope of this study.

5.1.1 Visual Observation of Fibre Motion Near the Screen

Surface

The motion of individual fibres near the low contour coupon using CSS was examined.
Figure 5.1 shows the accumulation of a dilute suspension fibres on the down stream
edge of a slot. When the rotor approaches the slot (Figure 5.1(b)), trapped fibres
remain on the edge and only start dislodging when the flow at the trailing edge of the
rotor accelerates, not because of the rotor negative pulse (Figure 5.1(c/d)). The slot
edge remains clear of any accumulation of fibres (Figure 5.1(e)) until the tangential
flow velocity drops to a certain value when accumulation starts again (Figure 5.1(f)).
It is believed that the same behaviour takes place in real applications, even though
the consistency is much higher. The linearity and similarity in plugging envelopes
(shown in Chapter 4) supports the application of this concept to include a wide range
of fibre concentrations.

5.2 Mathematical Model Details

The effect of the flow field on the forces applied on pulp fibres and consequently their
effect on capacity is studied herein. The experimental data used for this model are
presented in detail in Chapters 3 and 4. The model considers a force balance on a
lodged fibre on the downstream edge of the slot that prevents this fibre from moving.
The fibre continues to be lodged as long as this force difference is lower than the
frictional force generated between the fibre and the wire surface. In order to assess

60
5.2. Mathematical Model Details

Figure 5.1: Fibre trapping and clearing by rotor, Vt = 5 m/s and Vs = 3 m/s. (a)
xr /chord = -1.0, (b) xr /chord = 0.35, (c) xr /chord = 0.6, (d) xr /chord = 1, (e)
xr /chord = 1.35, (f) xr /chord = 4.0.

these forces, an accurate drag coefficient estimate at moderate Reynolds number is


needed. Vakil and Green [49] studied the flow around two-dimensional cylinders at
moderate Reynolds numbers, 1 ≤ Re ≤ 40. They presented the drag, CD and lift, CL ,
coefficients as best curve fits to computational data. They also showed that the aspect
ratio Lf /D of the fibre-like cylinders has no significant effect on the drag coefficient
for Lf /D > 10. Their fitted data were used in this study to measure the total forces
on fibre portions above and within the slot. The drag coefficient is calculated based

61
5.2. Mathematical Model Details

on the following equation

CD = CD,⊥ (A2 cos(2α) + A0 ) (5.1)

in which
CD,⊥ (Re , Lf /D) = κ1 Reκ2 (5.2)

and
ρVd D
Re = (5.3)
µ

where α is the angle between the flow and the fibre (and assumed to be 0o in this
study). Details of the parameters A0 , A2 , κ1 , and κ2 can be found in Vakil and Green
[49]. Analysis of the screen slot flow field confirms that the Re values are generally
within the range specified by Vakil and Green, with Re equal to 15 at a slot velocity
of 1 m/s and 45 for a slot velocity of 3 m/s.
In this study, the possibilities of a 3 mm fibre remaining immobilized (or ”trapped”)
on the wire surface were studied. Assuming a fibre diameter, D, of 0.015 mm, the
fibre was divided into 20 segments to keep the Lf /D ratio = 10. These forces then
interpolated to examine the possibility of 300 landing positions of a fibre on the wire
edge.
To evaluate fibre friction on a curved surface, Vakil and Green [48] compared
the fibre model with the Capstan Equation either in the classical or modified form
(see Jung et al. [28]). The simple form of the Capstan Equation provides the ratio
of tension on the tauter side to the slacker side of a rope wrapping a cylinder as
a function of the coefficient of friction and the wrap angle. Though the equation
is derived for a continuous rope, they model such a system with a finite number of
segments in their discrete approach. Generally speaking, they were able to match the

62
5.2. Mathematical Model Details

Figure 5.2: Tension on a fibre trapped on wire edge.

fibre model simulations with the Capstan Equation for low values of the coefficient of
friction and wrap angle. The same approach was adopted here; the average flow field
data were measured previously over different contour heights in Chapter 3. Figure
5.2 shows the tension forces and the contact angle between the fibre and the wire
edge.
A Matlab code was developed to estimate the tension on a fibre portion above and
within the slot by using the steady-state flow field data generated by a smooth rotor,
(Section 3.3) with the low contour. Velocity vectors parallel to the trapped fibre were
used to calculate drag forces on each segment. It was found that the tension on the
fibre portion above the slot, Tu , is not strongly affected by the slot velocity, Vs , as
shown in Figures 5.3 and 5.4. It was possible, therefore, to formulate the tension
above the slot solely as a function of the stream main velocity component, u, above
the screen surface. Figure 5.5 shows the tension forces on fibre portions above the
slot as a function of the smooth rotor speed. By normalizing the u component with
respect to rotor tip speed, Vt , the same profile was observed at different slot velocities,

63
5.2. Mathematical Model Details

80
Vs = 1 m/s
70
Vs = 2 m/s
60 Vs = 3 m/s

50 Average

Tu (10-6 N)
40

30

20

10

-10
0 0.2 0.4 0.6 0.8 1
l/Lf

Figure 5.3: Tension on the fibre portion above slot at Vt = 5 m/s.

as shown in Figure 5.6.


Since the tension on fibre portion above the slot depends on the stream main
velocity, a curve fitted formula (Equation 5.4), was developed to estimates Tu (i) at
any rotor position for any fibre segment above the slot. The tension data presented
in Figure 5.5 were used in this formula.

Tu (i) = [0.71 (u/Vt )Vt + 0.3 l(i)2 + 0.92 (u/Vt )Vt − 0.02 l(i) − 0.75] · 10−6 (5.4)

The total tension on the fibre portion above the slot can then be estimated as:

n
X
Tu = Tu (i) (5.5)
i=1

The tension within the slot was calculated using a corrected slot velocity that is

64
5.2. Mathematical Model Details

80
Vs = 1 m/s
70 Vs = 2 m/s
Vs = 3 m/s
60
Average
50
Tu (10-6 N)
40

30

20

10

-10
0 0.2 0.4 0.6 0.8 1
l/Lf

Figure 5.4: Tension on fibre portion above slot at Vt = 20 m/s.

80
Vt = 05 (m/s)
70
Vt = 10 (m/s)
60
Vt = 15 (m/s)
50 Vt = 20 (m/s)
Tu (10-6 N)

40

30

20

10

-10
0 0.2 0.4 0.6 0.8 1
l/Lf

Figure 5.5: Tension on fibre segments above the slot as a function of smooth rotor
tip speed.

65
5.2. Mathematical Model Details

0.8

0.6
u / Vt
0.4

0.2

0
-2 -1 0 1 2 3 4 5 6 7 8 9 10 11
xr / chord

Figure 5.6: Universal curve-fitted u component above the cylinder surface as a func-
tion of rotor position for 0 ≤ Vs ≤ 4 m/s.

a function of the local slot width. Figure 5.7 shows the changes in the v component
as a function of the slot distance along the slot axis and associated change in slot
width. This correction does not take into consideration the circulation that takes
place within the slot itself (shown by Mokamati et al. [35]) and was only a function
of the slot width
n
X
Ts = Ts (j) (5.6)
j=1

where
1
Ts (j) = ρ v(j)2 CD (i) D l(j) (5.7)
2

By examining the tension difference, |Tu − Ts |, as a function of the rotor position,


the portion of a fibre on the slot leading edge that can support a trapped state can be
estimated. This state is illustrated in Figures 5.8(a) and (b) for a 3 mm fibre trapped
on a low contour wire edge. For example, when the tension on the fibre portion above
the slot is greater than the tension on the fibre within the slot, the fibre remains
trapped as long as the tension difference is lower than or equal to the friction force
excreted on the fibre (Figure 5.8(b)). Trapped fibres will only be released when the

66
5.2. Mathematical Model Details

0.5

y (mm)
1.5

2.5

3
0.2 0.4 0.6 0.8 1
v/Vs

Figure 5.7: Corrected slot velocity along a 3 mm fibre within the slot.

tension difference is higher than the friction force on the trapped fibre. Note that
fibre repositioning on the slot edge is not possible in the present model.
By examining the probability of fibre trapping on a slot edge, the accumulation
of fibres can then be used as an indication of screen plugging as suggested in Section
5.1.1. Figure 5.9 shows trapping possibilities at three different rotor positions. As
the flow velocity above the cylinder surface changes with the rotor position, the
accumulation of fibres will be disrupted due to changes in the force balance applied
to hold fibres with a certain trapped lengths in position. Following on this model,
when the flow above a wire stabilizes at the characteristic quasi-steady u/Vt value,
as shown in Figure 5.6 for ∼ xr /chord > 3, accumulation for a certain fibre position
starts and continues until it gets disturbed by the next foil passage.
Figure 5.10 shows the proportion of a 3 mm fibre trapped on slot edge according
to the aforementioned model. The l/Lf values shown for a particular Vs represent

67
5.2. Mathematical Model Details

1.4

1.2
(b) Tu > Ts region
1
Forcce / Tu at (l=L)

0.8
(a) Tu < Ts region
0.6

0.4

0.2

0
Tu Ts |Tu - Ts| Friction force
-0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
l / Lf

(a) (b)

Figure 5.8: Tension changes with different trapped fibre positions are shown schemat-
ically in images (a) and (b) and analytically in a comparison of the fibre forces. In
this example, a fibre with l/Lf in the range of 0.50 to 0.66 is trapped (i.e. differences
in drag force are less than the friction force).
68
5.2. Mathematical Model Details

35
Tension difference (Xr/chord = +0.85; maximum u)
30 (Xr/chord = -2.0; steady state)
(Xr/chord = +0.4; minimum u)

25 Friction force (Xr/chord = +0.85)

|Tu - Ts | . 10-6 (N)


20

15

10

0
0 0.2 0.4 0.6 0.8 1
l/Lf

Figure 5.9: Changes in trapping position due to changes in the flow field (Vt = 5 m/s
and Vs = 1).

the 3 mm fibres landing with a portion length = l above the slot to be trapped
on the slot edge. These fibres remain trapped and other fibres that land with the
same l/Lf value will build up over time. When the flow field changes, as it does
with rotor position, and to an extent that the l/Lf value is no longer within the
range where trapping is supported for the instantaneous value of u, the fibres will
be cleared. Figure 5.10 also shows that accumulation cannot happen when there is
a reversal flow. For example at Vs = 1 m/s, no accumulation takes place during
reversal flow phase. This reversal time is still shorter than the disturbance time
created by the rotor which is defined here as when u/Vt is substantially different
from its average value, i.e. roughly 0 < xr /chord < 2 in Figure 5.10. Once the flow
stabilizes, accumulation starts again. The figure also suggests that accumulation is
roughly affected by the xr /chord ratio. It is important to note that rotor cycles have
the same overall form but they naturally take shorter times for higher rotor speeds.

69
5.2. Mathematical Model Details

This means accumulation is directly proportional to slot velocity and inversely to


rotor speed.
Figure 5.11 describes this model of fibre mass build-up on screen slots. This model
represents a process of accumulation equalling deposition followed by removal. Fibre
deposition occurs over time between rotor passings and removal is caused by rotor
passage over the apertures. The removal allows fibres to pass through the slots but
continued accumulation of fibres would lead to complete plugging of the cylinder,
Figure 5.11(c).
The accumulation of 3 mm fibres within a slot is shown in Figure 5.10 for three
different slot velocities. For Vs = 1 m/s, there is a narrow band of l/Lf possibilities
(∼ 2% of total) that will become trapped. As the rotor foil passes, the band moves
significantly and there is then a reverse flow that is assumed to release all trapped
fibres. At xr /chord ∼ 3, the mainstream flow has stabilized so that fibres can accu-
mulate until the return of the foil passage, i.e. the l/Lf band does not shift more than
the width of the band. This period of accumulation is designated as xac /chord. Note
that the same data for [10 < xr /chord < 11] and [-1 < xr /chord < 0] are repeated
in Figure 5.6. A similar sequence of events is seen for Vs = 2 m/s and Vs = 4 m/s
except that there is no reverse flow in these other two cases. The width of the l/Lf
band and the initial point of the xac /chord interval are approximately the same for
all three cases. This suggests that the initiation of plugging will increase linearly with
lower rotor speeds. This occurs because the tangential flow spends a longer time over
the aperture, giving Vs a longer time to draw fibres into the aperture and thereby
increase deposition.
Screen capacity can, in turn, be modelled as the critical case where fibres accu-
mulate and fill the slot before the release phase occurs. As discussed previously, in

70
5.2. Mathematical Model Details

reference to Figure 3.25, the backflushing action which has traditionally be consid-
ered essential to the release phase, may be supplemented by the action of a turbulent
wake flow or, as discussed in reference to Figure 5.10, flow instabilities which move
the instantaneous value of l/Lf so that previously trapped fibres are shed. What is
most critical to this trapping-capacity model is the accumulation of fibres. It follows
from the above force balance that given l/Lf band that leads to trapping is relatively
constant in width and that the xac /chord interval is relatively constant in length,
one can develop a fibre Accumulation Number, Nac , as described in Equation 5.8. In
particular, and as a first-order approximation, the equation assumes that the delivery
of fibres to the trapping location is proportional to the slot velocity, Vs , and upstream
consistency, CU . The percentage of delivered fibres that impact the downstream slot
edge with a value of l/Lf within the band that produces trapping was found above
to be relatively constant and is embraced within the constant, k1 . The time for the
build-up of fibres is proportional to the length accumulation zone, xac /chord as seen
in Figure 5.10 (with the simple conversion of chord length to distance also being em-
bedded within k1 ) and inversely proportional to upstream velocity, which is related
by a constant to rotor tip speed, Vt . Figure 5.12 is an illustration of Equation 5.8.

xac 1
Nac = [CU Vs ] k1 (5.8)
chord Vt
0
If plugging is assumed to occur at the same critical value, Nac , then Equation 5.8
0
can be rearranged as follows, with k2 = [Nac /chord k1 ] and Vs0 is the minimum slot
velocity for screen operation

1 1
Vs0 = k2 Vt (5.9)
CU xac

71
5.2. Mathematical Model Details

Figure 5.10: Estimates of the accumulation of 3 mm fibres on the low contour wire
during one foil cycle at Vt = 10 m/s.

Equation 5.9 captures the linear nature of the the relationship between Vt and Vs
observed in Figure 4.3, reproduced below as Figure 5.12, and the first term of the
screen capacity, Equation 4.1, reproduced below as Equation 5.10

Vt∗ = k1 Vs + k2 (5.10)

Equation 5.9 does not account for the intercepts with the axis. It only accounts
for fibre deposition, not fibre accumulation which depends on removal as well as
deposition. This suggests that the physical significance of second term in Equation
5.10 is linked to fibre removal. This is supported by the observation that extrapolation
to Vs = 0 gives a threshold velocity Vt∗ = k2 required to attain a non-zero Vs . Values of
k2 from the axis intercepts in Figure 4.4 are nearly linearly proportional to consistency,
i.e. Vt∗ = k3 CF This link between consistency and Vt suggests that removal is governed
by fibre network strength hydrodynamic force.

72
5.2. Mathematical Model Details

(a)
cleaning time
deposition  time by rotor wake
Masss of trapped fibrees 

Rotor foil 1 Rotor foil 2
(b)

Rotor foil 1 Rotor foil 2

(c) Mass  at which 
plugging occurs 

Time
Rotor foil 1 Rotor foil 2

Figure 5.11: Accumulation of fibres within screen slots during rotor cycles, (a) accu-
mulation is less than the cleansing effect, i.e. no fibres left from the previous cycle
(b) accumulation effect equals the cleansing effect (plugging point), (c) accumulation
effect is greater than the cleansing effect (plugging takes place).

73
5.3. Summary and Conclusions

8
Vt = 5 m/s
7
Vt = 10 m/s
6 Vt = 20 m/s

5
Nac
4

0
0 1 2 3 4
Vs (m/s)
Figure 5.12: Accumulation number, Nac , for a 3 mm fibre as a function of slot velocity
at different rotor speeds.

The above observations offer a basis for linking the empirical findings of Chapter 4
to a mechanistic model of screen plugging based on mass balances and force balances.
Doing so is beyond the scope of this thesis but remains a promising topic for future
work.

5.3 Summary and Conclusions

It is proposed that the accumulation of fibres within screen cylinder slots is a key
mechanism leading to cylinder plugging. Fibres need to staple between two screen
apertures to be trapped. Fibres start to accumulate on the downstream edge of the
slot when the drag force difference applied on fibre portions above and within the
slot is smaller than the fibre-wall friction force. Fibres continue to accumulate on the
slot edge until a disturbance in the flow field is introduced. This mechanism likely

74
5.3. Summary and Conclusions

acts in concert with other effects: The negative pulse generated by the rotor can also
prevent accumulation, as does the rotor wake afterwards.
The model suggests that the plugging of screen cylinder is strongly dependent
on fibre length distribution, thickening along the screen cylinder and the fluctuation
of the flow generated by the rotor foils. More variables, such as slot width, fibre
stiffness, fibre-wall friction coefficient, turbulence effects and debris content should
also be included in the development of a comprehensive screen capacity model.

75
Chapter 6

Summary and Conclusions

6.1 Conclusions

In this chapter, the conclusions of the main three chapters of this thesis are first
explained and the final conclusion is then stated.

6.1.1 Flow Field Study (Chapter 3)

The flow field near a slotted porous surface in steady and time-varying crossflow
was studied experimentally in this chapter for different degrees of roughness and
different upstream and aperture velocities. While this problem is of general interest,
the present study was made in the context of the flow of a fibrous suspension in a
pulp screen. Particle-image velocimetry was used in this study to measure the flow
patterns near the slotted surface. The size of the vortex within the roughness contours
was found to be dependent on the geometry of the contour and the flow velocities.
Vortices increase in size with lower aperture velocities and higher contour heights.
This study also considered the flow through slots in the slotted surface. The height
of the exit layer that passes from the main flow through the slotted apertures was
found to increase with slot velocity.
The dynamic studies revealed that the stream-wise flow decelerates under the
leading edge of the screen rotor, for approximately the first half of the rotor element.

76
6.1. Conclusions

The flow increases under the trailing half of the rotor and then gradually drops to an
average value that is independent of slot velocity.
A flow reversal was observed in some, but not all, of the flow configurations. Lower
slot velocities and higher rotor speeds supported the occurrence of a flow reversal.
Indeed a particular rates of these speeds defined the boundary between reversal and
non-reversal flow regimes.
Studies of particle concentration showed that the concentration within the contour
and leading into the slot, was generally, but not always, lower than in the flow above
the screen surface. This is consistent from published mechanisms of fibre motion at
a bifurcating flow. A depleted zone with reduced fibre concentration was observed
in the wake of the rotor foil, which may be industrially significant in avoiding the
blockage of the apertures with fibres.
In conclusion, reversal flow through the slot and turbulent flows (i.e. the rotor
pulse) in the wake of the foil, however, remain the principal mechanisms in avoiding
aperture blockages. This study provides a number of insights into the complex and
industrially important flow problem of unsteady flow adjacent a rough slotted wall
and the potential for the apertures in the wall to be obstructed with fibrous particles
being carried in the flow.

6.1.2 Pilot Pulp Screen Capacity (Chapter 4)

Pilot screen experiments with different screen cylinder aperture geometries and pulp
combinations were conducted in this chapter to determine the factors that affect the
maximum capacity of a pilot pressure screen. The screen plugging envelopes were
shown to be linear, suggesting a simple relationship between rotor and slot velocities.
A characteristic equation was proposed to describe screen capacity.

77
6.1. Conclusions

A higher contour height was found to delay plugging but this effect was limited
to low slot velocities. Slot width and wire width were found to be significantly more
important than contour height. Pulp consistency and character (i.e. fibre length
distribution) both had significant effects on screen capacity.
A comparison of the flow conditions where backflushing occurred (from Chapter
3) with the operating envelope of the pilot screen suggests that the screen can still
operate at lower rotor speeds where a negative pulse is not expected to be present.
In conclusion, this suggests that an alternate fibre removal mechanism exists in
addition to backflushing and wake turbulence.

6.1.3 Model of Pulp Screen Capacity (Chapter 5)

It is theorised in this chapter that the progressive accumulation of fibres on screen


cylinder slots is a precursor of plugging. Fibres need not to staple across adjacent
screen apertures to be trapped. Fibres accumulate on the slot’s downstream leading
edge when the drag force difference applied on fibre portions above and within the
slot is smaller than the friction force between the fibre and slot edge. Fibres thus
accumulate on the slot edge, and continue to build up with time until a disturbance
in the flow field is introduced.
In conclusion, a flow filed disturbance need to be a reversal flow to prevent accu-
mulation is a screen slot. The negative pulse generated by the rotor can clear the slot
(unplug) immediately while the rotor wake continues to prevent fibre deposition after
passage of the foil. A computational model was created based on the aforementioned
force balance mechanism. It suggests that the plugging of pressure screen according
to this mechanism is strongly dependent on fibre average length and the fluctuation
of the flow generated by the rotor foils.

78
6.2. Future Work

6.2 Future Work

Further work on this subject is clearly desirable. Following are specific suggestions.
There is interest in investigating fibre-wall interaction and the effect of fibre stiff-
ness on the total forces applied on trapped fibres, as this is key to the force balance
mechanism of trapping. Additionally, the effect of different rotor designs on stream-
wise flow fields and turbulence levels near screen surface should also be investigated.
Dynamic behaviour of the flow near screen apertures and within slots should be
studied in more details to associate the instantaneous effect on flow direction within
the slots. A detailed understanding of the small-scale geometry of the slot entrance
is required to understand and design high performance screen cylinders. this could
be based on high speed videos to characterize the accumulation of fibres, along with
CFD simulation of the time-varying flow validated by PIV experiments.
Finally, improved screen wire and rotor designs capable of delaying fibre accumu-
lations are also suggested to improve pressure screen capacity. These improvements
should also take in consideration the overall performance of the screen on contaminant
removal efficiency, fractionation, and thickening as well as capacity.

79
References

[1] A. Ammala, O. Dahl, H. Kuopanportti, and J. Niinimaki. The effect of back


flow in an axially fed pressure screen. Papier Ja Puu, 81(4):210–215, 1999.

[2] C. P. J. Bennington and R. J. Kerekes. Power requirments for pulp suspension


fluidization. Tappi J., 29(2):253–258, 1996.

[3] S. Blaser. Flocs in shear and strain flows. J. Colloid and Interface Sci.,
225(2):273–284, 2000.

[4] L. Bliss. Screening in the stock preparation system. In Proc. Tappi Stock Prepa-
ration Short Course, pages 59–75, Atlanta, GA, 1990.

[5] P. Bradshaw and F. Y. H. Wong. The reattachment and relaxation of a turbulent


shear layer. J. Fluid Mechanics, 52:113–135, 1972.

[6] B. Dalpke and R. J. Kerekes. The influence of fibre properties on the apparent
yield stress of flocculated pulp suspensions. J. Pulp and Paper Science, 31(1):39
– 43, 2005.

[7] S. Delfel. A numerical and experimental investigation into pressure screen foil
rotor dynamics. PhD thesis, The University of British Columbia, Canada, 2009.

[8] S. Delfel, C. Ollivier-Gooch, J. Olson, and P. Wallace. Experimental measure-

80
References

ment of pressure pulses from a pulp screen rotor. volume 1, pages 763 – 772,
Montreal, Canada, 2010.

[9] S. Delfel, J. Olson, C. Ollivier-Gooch, and R. Gooding. Effect of pulse frequency


and cylinder diameter on pressure screen rotor performance. In 65th Appita
Annual Conf., pages 89–96, Rotorua, New Zealand, 2011.

[10] S. Delfel, J. A. Olson, D. M. Martinez, A. Regairaz, C. F. Ollivier-Gooch, and


A. Huovinen. Influence of cylinder design and other factors on capacity and
power consumption in a pressure screen. Appita J., 64(1):55–61, 2011.

[11] S. Dong, M. Salcudean, and I. Gartshore. The effect of slot shape on the perfor-
mance of a pressure screen. Tappi J., 3(5):3–7, 2004.

[12] M. Feng, J. Gonzalez, J. A. Olson, C. Ollivier-Gooch, and R. W. Gooding.


Numerical simulation and experimental measurement of pressure pulses produced
by a pulp screen foil rotor. J. Fluids Eng., 127(2):347–357, 2005.

[13] F. Frejborg. Improved operation of TMP plant trough optimization of screening.


Pulp Paper Canada, 89(1):107–112, 1989.

[14] J. Gonzalez. Characterization of design parameters for a free foil rotor in a


pressure screen. MASc thesis, The University of British Columbia, Canada,
2002.

[15] R. W. Gooding. The passage of fibres through slots in pulp screening. MASc
thesis, The University of British Columbia, Canada, 1986.

[16] R. W. Gooding. Flow resistance of screen plate apertures. PhD thesis, The
University of British Columbia, Canada, 1996.

81
References

[17] R. W. Gooding and D. F. Craig. The effect of slot spacing on pulp screen
capacity. Tappi J., 75(2):71–75, 1992.

[18] R. W. Gooding and R. J. Kerekes. Motion of fibres near a screen slot. J. Pulp
Paper Sci., 15(2):59–62, 1989.

[19] G. Gregoire, M. Faver-Marinet, and F. Julien Saint Amand. Modeling of tur-


bulent fluid flow over a rough wall with or without suction. Trans. ASME,
125:636–642, 2003.

[20] J. Gullichsen and E. Harkonen. Medium consistency technology i: Fundamental


data. Tappi J., 64(6):69–71, 1981.

[21] L. Halonen, R. Ljokkoi, and K. Peltonen. Improved screening concepts. In Proc.


Tappi Pulping Conf., pages 61–66, Seattle, WA, 1989.

[22] M. Hamelin, S. Delfel, J. Olson, and C. Ollivier-Gooch. High performance multi-


element foil (MEF) pulp screen rotor - pilot plant and mill trials. J. of Pulp and
Paper Science, 36:3–4, 2011.

[23] O. Heise. Screening foreign material and stickies. Tappi J., 75(2):78–81, 1992.

[24] J. Jimenez. Turbulent flows over rough wall. Ann. Rev. Fluid Mech., 36:173–196,
2004.

[25] H. Jokinen, A. Ammala, J. A. Virtanen, K. Lindroos, and J. Niinimaki. Pressure


screen capacity-current findings on the role of wire width and height. Tappi J.,
6(1):3–10, 2007.

82
References

[26] F. Julien Saint Amand and B. Perrin. Fundamentals of screening: Experimen-


tal approach and modelling. In Proc. Tappi Pulping Conf., pages 1019–1031,
Montreal, 1998.

[27] F. Julien Saint Amand and B. Perrin. Fundamentals of screening: Effect of rotor
design and fibre properties. In Proc. of Tappi Pulping Conf., pages 941–955,
Orlando, FL, 1999.

[28] J. H. Jung, N. Pan, and T. J. Kang. Capstan equation including bending rigidity
and non-linear frictional behavior. Mechanism and Machine Theory, 43(6):661–
675, 2008.

[29] R. Karvien and L. Halonen. The effect of various factors on pressure pulsation
of a screen. Paperi ja Puu, 66(7):80–83, 1984.

[30] R. Kerekes. Pulp floc behavior in entry flow to constriction. Tappi J., 66(1):88–
91, 1983.

[31] A. Kumar. The passage of fibres though screen apertures. PhD thesis, The
Univeristy of British Columbia, Canada, 1991.

[32] S. Levis. Screening of secondary fibers. Progress in Paper Recycling, 1(1):31–45,


1991.

[33] D. M. Martinez, R. W. Gooding, and N. Roberts. A force balance model of pulp


screen capacity. Tappi J., 82(4):181–187, 1999.

[34] C. McCarthy. Various factors affect pressure screen operation and capacity. Pulp
and Paper, 62(9):233–237, 1988.

83
References

[35] S. Mokamati, J. A. Olson, and R. W. Gooding. Numerical study of separated


cross-flow near a two-dimensional rough wall with narrow apertures and suction.
Canadian J. Chem. Eng., 88(1):33–47, 2010.

[36] J. Niinimaki. Phenomena affecting the efficiency of a pressure screen. In Proc.


Tappi Pulping Conf., pages 957–966, 1999.

[37] J. Niinimaki, A. Ammala, H. Kuopanportti, and S. Nissila. The settings of


hydrofoils in a pressure screen. In Proc. Int. Symposium on Filtration, pages
71–78, Las Palmas, Canary Islands, 1998.

[38] J. Niinimaki, O. Dahl, H. Kuopanportti, and A. Ammala. Compa.son of pressure


screen baskets with different slot widths and profile heights - selection of the right
surface for a groundwood application. Paperi ja Puu, 80(8):601–605, 1998a.

[39] J. A. Olson. The effect of fibre length on passage througth a single screen aperture.
PhD thesis, The University of British Columbia, Canada, 1996.

[40] J. A. Olson and R. Kerekes. Motion of fibres in turbulent flow. J. Fluid Mechan-
ics, 377:47–64, 1998.

[41] J. A. Olson and G. Wherrett. A model of fibre fractionation by slotted screen


apertures. J. Pulp Paper Sci., 24(12):398–402, 1998.

[42] Specialist Committee on Uncertainty Analysis. Uncertainty analysis particle


image velocimetry. In Proc. of the 25th ITTC, volume 2, pages 453–455, Fukuoka,
Japan, 2008.

[43] V. C. Patel. Flow at high Reynolds number and over rough surfaces-achilles heel
of cfd. ASME J. Fluid Eng., 120:434–444, 1998.

84
References

[44] T. Paul, G. Duffy, and D. Chen. Viscosity control as a new way to improve
pressure screen performance. Tappi J., 83(9):61–100, 2000.

[45] V. Pinon, R. W. Gooding, and J. A. Olson. Measurements of pressure pulses


from a solid core screen rotor. Tappi J., 2(10):9–12, 2003.

[46] P. J. Pritchard, editor. Fox and McDonald’s Introduction to Fluid Mechanics.


Wiley, 2011.

[47] B. Timmins. Automatic Particle Image Velocimetry Uncertainty Quantification.


MASc thesis, Utah State University, USA, 2011.

[48] A. Vakil and S. Green. Drag and lift coefficients of inclined finite circular cylin-
ders at moderate Reynolds numbers. Computers and Fluids, 38(9):1771–1781,
2009.

[49] A. Vakil and S. Green. Flexible fiber motion in the flow field of a cylinder. Int.
J. Multiphase Flow, 37(2):173–186, 2011.

[50] T. Wikstrom and T. Rasmuson. Transition modelling of pulp suspensions ap-


pliled to a pressure screen. J. Pulp Paper Sci., 28:374–378, 2002.

[51] A. Yong, S. Mokamati, Daniel Ouellet, R. W. Gooding, and J. A. Olson. Exper-


imental measurement of fibre motion at the feed surface of a pulp screen. Appita
J., 61(6):485–489, 2008.

[52] C. J. Yu and R. J. DeFoe. Fundamental study of screening hydraulics, Part 1:


Flow patterns at the feed-side surface of screen baskets; mechanism of fiber-mat
formation and remixing. Tappi J., 77(8):219–226, 1994a.

85
References

[53] C. J. Yu and R. J. DeFoe. Fundamental study of screening hydraulics, Part 2:


Fiber orientation in the feed side of a screen basket. Tappi J., 77(9):119–124,
1994b.

[54] C. J. Yu, R. J. DeFoe, and B. R. Crossley. Fundamental study of screening


hydraulics, Part 3: Model for calculating effective open area. Tappi J., 77(9):125–
131, 1994c.

86
Appendices

A Flow and Wire Geometry Effects on Fibre

Fractionation

The effect of stream-wise flow and slot velocity is discussed briefly in this appendix.
By inspecting Figures A.1-A.4, one can see that the fractionation of longer fibres can
be easily achieved with low contour wires. It is shown in Figures 3.9 and 3.10 that
higher level of circulations, i.e. higher level of energy, can be achieved with higher
slot velocities. Higher circulation levels tend to curl fibres (CSS high speed videos
physically prove it) forcing them to follow the vortex flow and increasing their chances
to enter the slot.
Shorter fibres tend to follow the flow within the vortex and it is easier for them to
enter the slot. Longer and stiffer fibres cannot bend easily to follow the vortex flow
and the chances of them being rejected with lower contour screen wires are higher.
When slot velocity increases, the tendency for fractionation drops as fibres are forced
to enter the slot (Figure A.5). Figure A.6 explains this concept where higher contours
create larger vortices allowing longer fibre to be more easily accepted in comparison
to low contour screen wires.
The foil rotor has a stronger effect on stream-wise flow, creating higher levels of
circulation when it passes the slots than in comparison to a smooth rotor at the same

87
A. Flow and Wire Geometry Effects on Fibre Fractionation

average stream-wise flow velocities, allowing longer fibres to be accepted as well, as


shown in Figure A.7.

88
A. Flow and Wire Geometry Effects on Fibre Fractionation

9
8 Feed Sample
7 Vt = 20, Vs = 1
Vt = 20, Vs = 2
6
Frrequency (%)
Vt = 20, Vs = 3
5
4
3
2
1
0
0 1 2 3 4 5
Fibre Length

Figure A.1: Effect of slot velocity on fibre length distribution with a low contour wire
and smooth rotor at Vt = 20 m/s.

9
8 Feed Sample
7 Vt = 25, Vs = 1
Vt = 25, Vs = 2
6
Frrequency (%)

Vt = 25, Vs = 3
5
4
3
2
1
0
0 1 2 3 4 5
Fibre Length

Figure A.2: Effect of slot velocity on fibre length distribution with a low contour wire
and smooth rotor at Vt = 25 m/s.

89
A. Flow and Wire Geometry Effects on Fibre Fractionation

3
Feed Sample
2.5 Vt = 20 m/s, Vs = 1 m/s
Vt = 20 m/s, Vs = 2 m/s
2
Frequency(%)
1.5
F

0.5

0
0 1 2 3 4 5
Fibre Length (mm)

Figure A.3: Effect of slot velocity on fibre length distribution with a high contour
wire and smooth rotor at Vt = 20 m/s.

3
Feed Sample
2.5 Vt = 25 m/s, Vs = 1m/s
Vt = 25 m/s, Vs = 2 m/s
2
Frequency (%)

1.5
F

0.5

0
0 1 2 3 4 5
Fibre Length (mm)

Figure A.4: Effect of slot velocity on fibre length distribution with a high contour
wire and smooth rotor at Vt = 25 m/s.

90
Smooth rotor fibre fractionation
Smooth rotor fibre fractionation
A. Flow and Wire Geometry Effects on Fibre Fractionation

size fractioon accept / size ffraction feed


0.9

0.8

0.7
Type
Low1,contour,
Ut = 20Vt
contour m/s
= 20 m/s
Low1,contour,
Type Ut = 25Vt = 25 m/s
m/s
0.6 Type
High3,contour,
Ut = 20 Vt
m/s= 20 m/s
High3,contour,
Type Ut = 25 Vt
m/s= 25 m/s
0.5
0 1 2 3 4 5
Vs (m/s)

Figure A.5: Effect of smooth rotor speed and contour height on fibre fractionation at
various slot velocities.

Figure A.6: Effect of contour height and vortex strength on short and long fibre
fractionation.

91
A. Flow and Wire Geometry Effects on Fibre Fractionation

5
Feed
Vt = 20, Vs = 1, EP Rotor
4
Vt = 25, Vs = 1, Smooth Rotor
Frrequency (%)

0
0 1 2 3 4 5
Fibre Length (mm)
Figure A.7: Effect of rotor on fibre fractionation at the same flow field.

92
B. PIV Uncertainty Analysis

B PIV Uncertainty Analysis

Particle Image Velocimetry (PIV) measurement error depends on the PIV algorithm
used, a wide range of user inputs, flow characteristics, and the experimental setup.
Since these factors vary in time and space, they lead to non-uniform error throughout
the flow field [47].
The uncertainty analysis presented in this study followed the procedure recom-
mended by [42]. The principle of the PIV measurement on flow speed u can be
described by the following equation

4X
u = α1 + δu (B.1)
4t

The flow speed is detected by means of the displacement of particle images 4X, and
the time interval of successive images 4t. The magnification factor, α1 , must be be
identified through the calibration. Table B.1 shows the PIV measurement dimensions.
Since no laser was used, the virtual test plane was assumed to be perpendicular to the
view field with a maximum deviation of 5o . Calibration was conducted by measuring
the screen wire width. The magnification factor was determined by the distance of
the reference point, lr and the distance on the image plane, Lr

lr
α1 = (B.2)
Lr

The PIV measurement based on the visualized flow image, and the information
of the image differs from the flow field due to the velocity lag of the tracer particle
from acceleration and the projection procedure from the 3-D physical space to the
2-D image plane. These uncertainty factors of flow visualization are consolidated in

93
B. PIV Uncertainty Analysis

Table B.1: Principal dimensions for PIV measurment

Flow maximum speed 8 m/s


Distance of reference points lr 3.2 mm
Distance of reference image Lr 195 pixel
Magnification factor α1 0.016 mm/pixel
Time interval 4t 1.852 × 10−5 s
Spatial resolution 320 × 240 pixels
Distance from the target lt 60 mm
Correlation area size 64 × 64 pixels
Search area size 32 × 32 pixels

a parameter δu. In general, the δu is hard to detect systematically, and it is usually


categorized as an uncertainty factor rather than a measurement parameter [42]. The
error from the particle velocity lag was experimentally found to be less than 0.5%
and for this case, δu = 8 × 1000 × 0.005 = 40 mm/s. The 3-D out-of-plane velocity
component was assumed to be 1.0% of the maxium flow velocity and the error was
estimated as 8 × 1000 × .01 × tan(1/2 × 5.25/60) = 3.5 mm/s.
TablesB.2 and B.3 show all the uncertainties, propogation and accumulation of
uncertainties of measurment parameters for velocity and position, respectively. The
uncertainty of 4t is small enough to be neglected. Full details of all standard uncer-
tainties, us (xi ), and sensitivity factors, ci , evaluation procedure can be found in [42].
The total combined uncertainties were calculated based on the following summation

p
uc = u2u + (ux ∂u/∂x)2 (B.3)

The expanded uncertainty in velocity was found to be ±3.5% for the 8 m/s flow.
The uncertainties of model test were not included in this analysis.

94
B. PIV Uncertainty Analysis

Table B.2: Uncertainties for velocity


Prmt. Category Error sources us (xi ) ci ci us (xi ) uc
α1 Calibration Reference image 0.5 pixel 8.42 × 10−5 mm/pixel2 4.21 × 10−5
Physical distance 0.005 mm 5.13 × 10−3 1/pixel 2.56 × 10−5
Image distortion 0.95 pixel 8.42 × 10−5 mm/pixel2 8.00 × 10−5
by lens
View field position 0.5 mm 2.74 × 10−4 1/pixel 1.37 × 10−4
Normal view angle 0.087 rad 1.43 × 10−3 mm/pixel 1.24 × 10−4 2.07 × 10−4
4X Acquisition Normal view angle 0.087 rad 1.43 × 10−3 mm/pixel 1.24 × 10−4
Reduction Mismatching error 0.1 pixel 1.0 0.1
Sub pixel analysis 0.03 pixel 1.0 0.03 0.104
δu Experiment Particle trajectory 40 mm/s 1.0 40 mm/s
3-D effects 3.5 mm/s 1.0 3.5 mm/s 40.15

α1 Magnification factor 2.07 × 10−4 mm/pixel 4.88 × 105 pixel/s 101 mm/s
4X Image displacement 0.104 pixel 888.6 mm/pixel/s 92.41
δu Experiment 40.15 mm/s 1.0 40.15 mm/s
Combined uncertainty uu 142.66 mm/s

Table B.3: Uncertainties for position

Parameter Category Error sources us (xi ) ci ci us (xi )


Xs , Xe Acquisition Digital error 0.5 pixel 0.0164 mm/pixel 8.2 × 10−3
Non-uniformity of 16 pixel 0.0164 mm/pixel 0.263
distribution
Xo Calibration Origion correlation 2.0 pixel 0.0164 mm/pixel 0.033
α1 Magnification factor 2.07 × 10−4 mm/pixel 160 pixel 0.033
Combined uncertainty ux 0.267 mm

95
C. Characterization of Fibres Used for Capacity Trials

C Characterization of Fibres Used for Capacity

Trials

Different pulp combinations were used to evaluate the effect of average fibre length
on maximum screen capacity as seen in Section 4.3. The softwood to hardwood ratio
was varied from 0 to 100 during the course of experiments. Figures C.1-C.5 show the
fibre length distributions for these combinations.
The model introduced in Chapter 5 was based on a fibre length of 3 mm. Although
the impact of the average fibre length on the model was not explicitly taken into
consideration, it was found that the frequency of 3 mm fibre, Figure C.6, increased
linearly with the increase of softwood ratio. This might be used in future work to
include the effect of fibre length or the average fibre length in the developed model.

8
0:100 HW:SW
7

5
Freqquency (%)

0
0 1 2 3 4 5
Fibre Length (mm)
Figure C.1: Fibre length distributions of 0:100 SW:HW sample.

96
C. Characterization of Fibres Used for Capacity Trials

8
25:75 HW:SW
7

quency (%) 5

3
Freq

0
0 1 2 3 4 5
Fibre Length (mm)
Figure C.2: Fibre length distributions of 25:75 SW:HW sample.

7 50:50 HW:SW

5
quency (%)

3
Freq

0
0 1 2 3 4 5
Fibre Length (mm)
Figure C.3: Fibre length distributions of 50:50 SW:HW sample.

97
C. Characterization of Fibres Used for Capacity Trials

8
75:25 HW:SW
7

Freqquency (%) 5

0
0 1 2 3 4 5
Fibre Length (mm)
Figure C.4: Fibre length distributions of 75:25 SW:HW sample.

8
100:0 HW:SW
7

5
Freqquency (%)

0
0 1 2 3 4 5
Fibre Length (mm)
Figure C.5: Fibre length distributions of 100:0 SW:HW sample.

98
C. Characterization of Fibres Used for Capacity Trials

2
R² = 0.9964
1.6
Frrequency (%)

1.2

0.8

0.4

0
0 25 50 75 100
SW:HW (%)
Figure C.6: Frequency of 3 mm fibres in different SW:HW samples.

99

You might also like