+ All Categories
Home > Documents > The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on...

The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on...

Date post: 25-Feb-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
219
The Simulation of Polymer Aggregation/Breakage at High Solid Fraction in Turbulent Flow by Population Balance A dissertation submitted to Murdoch University Department of Engineering In partial fulfilment of the Requirements for the degree of DOCTOR OF PHILOSOPHY Alex Heath 2002 Supervisors Dr Parisa Bahri Dr Phillip Fawell
Transcript
Page 1: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

The Simulation of Polymer Aggregation/Breakage at High Solid

Fraction in Turbulent Flow by Population Balance

A dissertation submitted to

Murdoch University

Department of Engineering

In partial fulfilment of the

Requirements for the degree of

DOCTOR OF PHILOSOPHY

Alex Heath

2002

Supervisors

Dr Parisa Bahri

Dr Phillip Fawell

Page 2: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

1

I declare that this thesis is my own account of my research and contains as its main

content work which has not previously been submitted for a degree at any tertiary

education institution.

………………………………….

Alex Heath

Page 3: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

2

Acknowledgements

I would like to thank the following people for their contributions towards this Ph.D.

project:

Dr John Farrow - Identifying and setting up the project, overall direction

Dr Parisa Bahri - Primary supervisor, help with maths, coding,

engineering, and general support

Dr Phillip Fawell - Co-supervisor, chemistry, references, thesis preparation

Mrs Jean Swift - Laboratory & pilot plant trials, flocculation and settling

Mrs Evalyn Beaumont - Library

Dr John Rumball - General support, reading thesis

Dr Peter Koh - CFD, end user

Page 4: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

3

Abstract A mathematical model has been developed to describe the size and settling rate of

aggregates formed by flocculation in thickeners/clarifiers used in the mineral processing

industry. The aggregation process was simulated with a population balance model,

which calculates the aggregate size distribution through time as a function of the

competing processes of aggregation and breakage. The population balance was written

specifically to describe aggregation by high molecular weight polymer flocculants that

have now largely replaced coagulant salts in mineral processing operations, and the

model includes terms to describe flocculant/slurry mixing and irreversible aggregate

breakage (Table 1). The model was written to form part of a full computational fluid

dynamics (CFD) model of a thickener/clarifier, allowing the eventual simulation and

optimisation of the full-scale unit.

In addition to the effects of fluid shear and residence time normally described by

population balance models, additional process variables have been considered, with the

model correctly accounting for changes in the flocculant dosage, primary particle size

and solid fraction. The model has also been extended to describe the hindered settling

rate under the same range of conditions, considerably increasing its usefulness by

forming a link between the aggregation kinetics in the feedwell and the subsequent

setting rate as the aggregates enter the settling zone of the thickener.

The model was fitted to experimental data from a turbulent pipe reactor. A variety of

pipe sizes, lengths, and flow rates were used to give a range of fluid shear rates and

mean residence times, with the aggregate size distribution measured by an on-line

sizing probe placed at the end of the pipe, immediately before a settling column.

Experimental data was collected under a sparse matrix of experimental conditions, with

the fluid shear rate (G), flocculant dosage (θf), primary particle size (dp) and solid

fraction (φ) varied independently away from a common baseline. Additional data was

collected from conditions in the gaps of the experimental matrix, and was used to check

the predictivity of the model.

The population balance model is based on the discretised balance by Hounslow et al.

(1988) and Spicer and Pratsinis (1996), in this case using 35 channels covering the size

Page 5: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

4

range: 0.2-3500 μm. The aggregation rate is described by Saffman and Turner’s (1956)

turbulent collision kernel, used in conjunction with a capture efficiency term. The

capture efficiency is initially taken to be zero (no successful collisions) before

flocculant addition, but increases to unity (all collisions successful) as described by a

flocculant-suspension mixing term.

The breakage rate is described as a function of the aggregate size, energy dissipation

rate, suspension viscosity and effective flocculant surface coverage. The effective

flocculant coverage is taken to decrease through time, reflecting the loss of flocculant

activity caused by repeated aggregation/breakage. Aggregate porosity is also included,

using fractal geometry, and is used to calculate the aggregate capture radius and

effective suspension solid volume fraction. The solid volume fraction is used to

determine the suspension viscosity, accounting for experimental data showing the

pressure drop in the pipe is a function of the aggregate size.

The model equations were solved numerically as an initial value problem with a

commercial dynamic simulation package (gPROMS), which was also used to estimate

the unknown model parameters. The model was found to be robust and stable, and gave

good predictions of the additional experimental data sets.

The extension of the model to also describe the hindered settling rate allowed a dynamic

optimisation to give the maximum settling flux, and hence unit throughput. Various

optimisation strategies were investigated, in particular the use of recycle to find the

optimal feed solid fraction. The inclusion of the model within a full CFD simulation

will allow further optimisation strategies to be investigated, in particular changes to the

feedwell geometry to give good mixing but without subjecting fully formed aggregates

to regions of destructive high fluid shear.

Page 6: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

5

Table 1: Model Summary Process Equation Comment

D/Lfk1e1M −−=

where: (Equation 5-6)

M = Mixing index [0,1], used to estimate α k1 = Fitted parameter # 1 (0.343) f = Pipe friction factor L = Pipe length (m) D = Pipe diameter (m)

Simplified macro-scale mixing equation for turbulent pipe flow by Etchells & Short (1988), Godfrey & Amirtharajah (1991). Assumes flocculant/particle collision and adsorption are rapid. Likely to be highly scale-up dependent, requires further work.

( )3jiij aa294.1 +ν∈

α=β where: (Equation 2-7)

βij = Collision kernel) (m3s-1) ∈ = Energy dissipation rate (m2s-3) μ = Fluid viscosity (N s m-2) ai = Radius of the ith particle (m) α = Capture efficiency [0,1] (taken as = M)

Saffman & Turner’s (1956) collision kernel. ∈ and ν determined experimentally (Equations 4-2, 4-3, 4-7, Table 4-1) and predicted in the model (Equations 5-8, 5-16). Particle radius incorporates aggregate porosity (Equation 5-23), with the aggregate size given by an on-line, in-pipe sizing probe (FBRM).

dK

Sf

i,aggk

2i

3

θ

μ∈=

where: (Equation 5-19)

Si = Breakage rate of the ith sized particle (s-1) θf = Flocculant surface coverage (kg m-2) k2 = Fitted parameter # 2 := 38.1 k3 = Fitted parameter # 3 := 0.677 M = Mixing index [0,1] mf = Mass of flocculant (kg) dagg,i = Diameter of the ith particle (m)

Breakage kernel, similar to others in the literature (Table 2-6) incorporates flocculant dosage term (Equations 5-18, 5-20), i.e. more flocculant produces stronger aggregates and reduced breakage. Partially irreversible breakage introduced via reduced effective flocculant coverage (Equation 5-20), containing parameter k4.

∑∑

∑∑∞

=

=

=−−−

=−−

+−

Γ+−β−

β−β+β= −

ijjjj,iii

ijjj,ii

1i

1jjj,ii

21i1i,1i

2i

1j21

j1ij,1i1iji

NSNSNN

N2NNNN2dt

dN ij

where: (Equation 2-45)

Ni = Number of ith sized particles (m-3) Γij = Daughter fragment distribution function

Population balance for aggregation and breakage, after Hounslow et al. (1988), Spicer & Pratsinis (1996) (see Appendix). Solved numerically with gPROMS on SUN mainframe.

( ) 65.4D3

p

aggs

3D

p

aggls

2agg

h

f

f

d

d1

18

d

dgd

U⎟⎟⎟

⎜⎜⎜

⎟⎟

⎜⎜

⎛φ−

μ

⎟⎟

⎜⎜

⎛ρ−ρ

=

where: (Equation 6-10)

Uh = Settling velocity (m s-1) d = Diameter (p = primary particle, agg = aggregate) (m) g = Gravity (9.8 m s-2) ρ = Density (s = solid, l = liquid) (kg m-3) μ = Fluid viscosity (N s m-2)

Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law (Equations 6-1, 6-2). Aggregate porosity introduced using fractal geometry (Equations 6-8, 6-9). Equation used to estimate fractal dimension (Df = 2.4) from experimental data (Equation 6-11, Figure 6-6).

Dilute flocculant stream

Particle Suspension

Pipe flow →

Flocculant/particle mixing & adsorption

Particle or aggregate collision & aggregation

+

Aggregate breakage

Initial hindered settling velocity

Simulation/computing

Page 7: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

6

List of terms Aggregate Generic term describing an assemblage of smaller particles; may be

formed by the addition of a coagulant or flocculant. Capture efficiency (α) Ratio of the number of collisions resulting in stable aggregates to

the total number of particle collisions. Coagulant Soluble salt, usually a multivalent cation salt like alum

(Al2(SO4)3.16H2O), which reduces the surface charge resulting in aggregation (specifically: coagulation).

Coalescence Analogous to aggregation, except with liquid droplets. Sometimes

used to approximate aggregation due to the simplifications of the spherical shape and no porosity.

Compression Compaction or de-watering of sediment in thickener bed region.

Occurs when the height of the bed builds up to the point where the overburden weight exceeds the compressive yield stress, resulting in the restructuring of the network and the forcing of displaced fluid back up through the structure.

Fractal Dimension (Df) Exponent (typically 2-2.5) of power law relationship between

aggregate size and mass. Used to describe aggregate porosity by assuming the aggregate volume increases as the cube of the diameter.

Flocculant High molecular weight (∼ 15-20 × 106 g mol-1) polymer with

charged or polar functional groups to make it water-soluble and to attract it to the particle surface. Results in aggregation (specifically: flocculation) through a combination of surface charge neutralisation and polymer bridging.

Kernel Size dependent rate equation describing the kinetics of aggregation

or breakage. Solved within the population balance. Permeability Ability of fluid to pass through porous structure. Body may be

porous, but impermeable, i.e. if the pores do not join up. Porosity Void space (ε) in aggregate or sediment network due to the packing

efficiency of the primary particles. Population balance Dynamic model describing the change in the aggregate size

distribution through time. Usually solved in discrete form, where the continuous particle size distribution is sectioned, resulting in a closed set of ordinary differential equations that can be solved numerically as an initial value problem.

Sedimentation As per settling, but usually describing the network settling in the

hindered or compression regimes. Settling Generic term describing effect of gravity to drag particles down

through fluid of lower density.

Page 8: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

7

Contents

1. INTRODUCTION 9 1.1 Aggregation of suspended particles 9 1.2 Solid/liquid separation 10 1.3 Modelling aggregation kinetics by population balance 11 1.4 Aggregate settling 12 1.5 Project aim 13 1.6 Thesis overview 13

2. LITERATURE REVIEW 15 2.1 Introduction 15 2.2 Aggregation 16

2.2.1 Particle collision mechanisms 16 2.2.2 Turbulent collision, particles smaller than micro-scale 18 2.2.3 Turbulent collision, particles larger than micro-scale 21 2.2.4 Summary - particle collision 23 2.2.5 Effect of short range inter-particle forces/bonding mechanisms 26 2.2.6 Coagulation with salt 28 2.2.7 Flocculation with polymer 28 2.2.8 Loss of flocculant activity by polymer degradation 31

2.3 Aggregate porosity 32 2.4 Fluid behaviour 38

2.4.1 Laminar flow, shear stress (τ), viscosity (μ), and fluid shear (γ), 38 2.4.2 Turbulent flows, G, Kolmogoroff micro-scale, eddy spectrum 38 2.4.3 Deviations from ideal Newtonian behaviour 40 2.4.4 Effect of suspended solid on viscosity 42 2.4.5 Multi-phase mixing 44

2.5 Aggregate breakage 46 2.5.1 Cause of aggregate breakage 46 2.5.2 Effect of aggregate size on breakage 49 2.5.3 Daughter particle distribution 51

2.6 Population balance models 54 2.6.1 Smoluchowski’s equation, aggregate breakage 54 2.6.2 Discrete models, numerical solution 55

2.7 Solid-liquid separation by gravity settling 59 2.7.1 Particulate settling (low solid fraction) 61 2.7.2 Hindered settling (intermediate solid fraction) 68 2.7.3 Compression (high solid fraction) 77

2.8 Conclusions to literature review 81

3. ON-LINE PARTICLE SIZING 83 3.1 Particle sizing techniques 83 3.2 Principle of FBRM 85 3.3 Estimating particle size from FBRM data 86

3.3.1 Theoretical methods 86 3.3.2 Empirical methods 88

3.4 Experimental 90 3.4.1 Equipment 90 3.4.2 Samples 91

3.5 Results and discussion 91 3.5.1 Effect of applying weightings to FBRM chord distributions 91 3.5.2 Comparison of average particle size by various techniques 93 3.5.3 Effect of solid volume fraction on indicated size 94 3.5.4 Comparison between C and F-electronics 97 3.5.5 Effect of particle material 98 3.5.6 Minor effects 101 3.5.7 Effect of solid fraction on particle counts 104

3.6 Conclusions 106

Page 9: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

8

4. CALCITE FLOCCULATION IN TURBULENT PIPE FLOW 107 4.1 Experimental 109 4.2 Results and discussion 114

4.2.1 Effect of mean fluid shear rate 114 4.2.2 Effect of flocculant dosage 122 4.2.3 Effect of primary particle size 124 4.2.4 Effect of suspension solid fraction 125

4.3 Conclusions 127

5. POPULATION BALANCE MODEL 128 5.1 Model development 128

5.1.1 Aggregation 128 5.1.2 Aggregate breakage 133 5.1.3 Simplification for coding/simulation 137

5.2 Comparison with experimental data 140 5.2.1 Effect of flow regime (shear rate) 140 5.2.2 Effect of flocculant dosage 147 5.2.3 Effect of primary particle size 148 5.2.4 Effect of solid fraction 150 5.2.5 Additional runs to check model predictivity 153 5.2.6 Effect of changing fractal dimension 154

5.3 Conclusions 155

6. RELATING HINDERED SETTLING RATE WITH AGGREGATE SIZE 156 6.1 Experimental 158 6.2 Results and discussion 159

6.2.1 Effect of fluid shear 159 6.2.2 Effect of flocculant dosage 161 6.2.3 Effect of suspension solid fraction 162 6.2.4 Effect of primary particle size 164 6.2.5 Fractal geometry 164

6.3 Conclusions 169

7. DYNAMIC OPTIMISATION AND EXTRAPOLATION TO FULL-SCALE 170 7.1 Predicting settling behaviour –Dynamic optimisation 170 7.2 Extrapolation to Industrial scale 174 7.3 Conclusions 176

8. CONCLUSIONS AND FUTURE WORK 177 8.1 Conclusions 177 8.2 Suggested future research 181

9. NOTATION 182

10. REFERENCES 185

11. APPENDIX 209 11.1 Hounslow et al.’s (1988) population balance 209 11.2 Fractal geometry, calculation of dagg and ρagg from dm, dp & Df 216 11.3 Fractal geometry, calculation of effective mean aggregate density 217

Page 10: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

9

1. Introduction

1.1 Aggregation of suspended particles Solid-liquid separation in mineral processing and wastewater treatment is frequently

performed by gravity sedimentation (Amirtharajah 1991, Williams 1992, Shamlou

1993, Rushton et al. 1996, Bustos et al. 1999, Svarovsky 2000). Fine particles settle

slowly in viscous flow, and are usually aggregated to increase their settling rate. The

particles typically carry a surface charge that causes electrostatic repulsion between the

particles, preventing aggregation. However, this repulsion can be overcome by the

addition of soluble ions (coagulants), which are attracted to the particle surface where

they neutralise and compress the electrical double layer. The particles are then able to

approach each other closely enough that van der Waals attractive forces can hold the

particles together as stable aggregates (Kohler 1993, Hughes 2000).

However, in mineral processing circuits coagulants have been largely superseded by

high molecular weight (∼ 20 × 106 g mol-1) polymer flocculants (Bagster 1993, Farinato

& Dubin 1999). Flocculants have polar or charged functional groups attached to a

hydrocarbon backbone, making the polymer water-soluble and attracting it to the

mineral surface. In addition to the effect of surface charge neutralisation, the length of

high molecular weight flocculants allows them to span the gap between fine particles,

and polymer bridging (Figure 1-1) is usually assumed to be the dominant bonding

mechanism (Healy 1961, Gregory 1989).

Figure 1-1: Stages (somewhat concurrent) of flocculation

Aggregate breakage

Particle collision & aggregation

+

→ +

Flocculant/particle mixing & adsorption

Page 11: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

10

1.2 Solid/liquid separation For solid/liquid separation on an industrial scale, the concept of a simple settling-pond

is extended to continuous (under) flow by adding a conical base and a rotating rake

assembly (Figure 1-2) (Dahlstrom & Fitch 1985, Bustos et al. 1999). The rake aids

sediment bed compaction as it encourages the solid down towards the underflow pump

at the base (Pearse 1977, Svarovsky 2000). The unit is termed a thickener or clarifier,

depending whether the primary process objective is to increase or decrease the solid

fraction of the feed stream (Williams & Simons 1992, Perry & Green 1997). Several

units can be connected in series, forming a washer train or counter current decantation

(CCD) circuit (Dahlstrom & Emmet 1985).

A clarifier typically receives a feed with a low solid concentration (typically < 1 %

w/v), and may benefit from underflow solid recycle to increase fines capture during

flocculation. Conversely, a thickener is likely to have a relatively high feed solid

concentration (typically 2-50 % w/v) (Pearse 1977, Perry & Green 1997, Dahlstrom &

Fitch 1985) and may benefit from dilution by overflow recycle, either via the overflow

weir or within the feedwell, according to the propriety design (Amirtharajah 1991,

Svarovsky 2000).

Figure 1-2: Cross section of thickener/clarifier.

Dilute flocculant solution (< 1 % w/v) is mixed with the feed suspension, usually in the

feedwell or feed pipe (Dahlstrom & Fitch 1985, Perry & Green 1997). The primary

purpose of the feedwell is to dissipate the momentum of the feed stream, which might

otherwise disrupt the settled sediment. However, the agitation intensity generated in the

feedwell provides good mixing, making it a suitable flocculant addition point.

Flocculant

Overflow weir

Hindered settling region

Free settling (clarification)

Underflow

Feedwell Feed suspension

Sediment compaction Rotating rake

Page 12: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

11

Adequate mixing is required, initially to disperse the flocculant evenly through the feed

stream, preventing localised overdosing (Farinato & Dubin 1999, p. 18). Further

mixing is then required to cause flocculant/particle collision and adsorption, followed

by particle/particle collision to form aggregates. However, if the agitation is too intense

the aggregates will break, irreversibly degrading the polymer flocculant by chain

scission or rearrangement (Healy 1961, Williams & Simons 1992, Bagster 1993,

Gregory 1993).

The complex flow behaviour in industrial vessels can now be described by

computational fluid dynamics (CFD) models (Lainé et al. 1999, Ducoste & Clark 1999,

Farrow et al. 2000), which have been developed since the introduction of powerful

computers and efficient numerical routines. However, although many fluid flow

characteristics (kinetic and dissipative energy, fluid shear, multi-phase mixing, recycle,

etc) are readily calculated, CFD cannot currently predict the effect of these parameters

on the kinetics of aggregation or breakage.

1.3 Modelling aggregation kinetics by population balance Aggregation kinetics have been extensively studied on a laboratory scale, and widely

modelled by population balance. The population balance approach was first described

by Smoluchowski (1916, 1917):

∑∑∞

=

=+=

β−β=1i

kiik

1k

kji,1ijiij2

1k NNNNdt

dN 1-1

However, current population balance models for flocculation typically also include a

aggregation capture efficiency (α) term, and terms to describing aggregate breakage,

giving (Argaman & Kaufman 1968, Randolph & Larson 1988, Lick & Lick 1988,

Spicer & Pratsinis 1996, Serra & Casamitjana 1998b, Kramer & Clark 1999):

∑∑∑∞

+=

=

=+=

Γ+−αβ−αβ=1kl

lllkkk1i

kiik

1k

kji,1ijiij2

1k NSNSNNNNdt

dN 1-2

Page 13: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

12

where: Ni = Number of ith sized particles (m-3) t = Time (s) α = Capture efficiency [0,1] βij = Rate of collision between i and j sized particles (aggregation kernel) (m3s-1) Sk = Breakage rate (kernel) of kth sized particles (s-1) Γlk = Breakage distribution function (number of k size particles produced from the breakage of a l sized particle)

Despite the body of work already published in this area, few current population balance

models simultaneously account for changes to the process variables encountered in

practice (fluid shear rate, flocculant dosage, primary particle size, solid fraction). There

is also a need to link the predicted aggregate size back to the primary process objective:

solid-liquid separation.

1.4 Aggregate settling The velocity of an individual spherical particle settling under gravity was described by

Stokes (1851), who assumed creeping flow (Re → 0) and solved the remainder of the

Navier-Stokes equation, giving (Happel & Brenner 1973, Seville et al. 1997):

( )μρ−ρ

=18

gdU ls2

1-3

where: U = Settling velocity (m s-1) d = Particle diameter (m) g = Gravity (9.8 m s-2) ρ = Density (s = solid, l = liquid) (kg m-3)

μ = Fluid viscosity (N s m-2) However, aggregates are a loose assemblage of primary particles, with an irregular

shape and porous structure. These reduce the settling velocity by increasing the

hydrodynamic drag force, and reducing the effective particle density (Clift et al. 1978).

The settling velocity is also reduced as a function of the suspension solid volume

fraction, initially due to inter-particle hindrance, but eventually also from mechanical

support propagating up through the compressing sediment (Richardson & Zaki 1955,

Tory 1996, Bustos et al. 1999).

Page 14: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

13

1.5 Project aim The aim of this work was to develop a mathematical description of the kinetics of

aggregation/breakage, providing a suitable sub-model for a CFD model of a thickener

feedwell. This will ultimately allow CFD to explore various feedwell design options

(e.g. flocculant dosing point(s), feed entry points and/or velocity, baffle size/position,

recycle, etc) towards an improved design.

The model needed to describe the various stages of aggregation, from the initial

flocculant/suspension mixing and adsorption, the subsequent rate of aggregation, and

finally the (partially irreversible) aggregate breakage. In addition, the model had to

adequately account for the effects of the process variables found in operating plant, for

example: the fluid shear rate (G), the flocculant dosage (θ), the primary particle size

(dp) and solid fraction (φ).

The model was also extended to describe the hindered settling rate of the aggregates

formed under the same process conditions, considerably increasing the usefulness of the

model by allowing the calculation and optimisation of the unit throughput.

1.6 Thesis overview Aggregation and settling are mature areas of research. Chapter 2 reviews this material,

with separate sections for aggregation, aggregate structure, fluid behaviour, aggregate

breakage, population balance models, and finally gravity settling. While these topics

are individually well described, there is a surprising lack of synthesis between the

topics, in particular the effect of aggregation on the settling rate. Workers studying

aggregation have usually implicitly assumed that a larger aggregate will lead to a faster

settling rate, and although this is typically the case, it represents a limited view of

reasonably complex settling behaviour. Conversely, workers studying particle settling

generally continue to assume that the settling rate is a unique function of the solid

fraction, ignoring the effects of aggregation. This situation has resulted partially from

the previous lack of instrumentation available to give on-line, in-pulp aggregate size

measurement.

Chapter 3 considers the difficulty of in-pulp aggregate sizing, and describes the

Lasentec FBRM (Focussed Beam Reflectance Measurement) instrument. FBRM gives

Page 15: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

14

a particle chord distribution, as opposed to a conventional particle diameter distribution,

and the major thrust of this section was to develop a method to convert chord size data

to diameter data suitable for the population balance (Chapter 5) and settling (Chapter 6)

models.

Chapter 4 describes the experimental methods used to collect aggregation kinetics data.

Aggregation was performed in a turbulent pipe reactor, with a range of pipe lengths,

diameters, and suspension flow rates used to give a range of fluid shear rates and

residence times. The FBRM aggregate sizing probe was placed at the end of the pipe

reactor, immediately before the isolatable column used to take hindered settling rate

measurements. The effect of flocculant dosage (θ), primary particle size (dp) and solid

fraction (φ) were also studied under conditions approximating those found in full-scale

plant.

Chapter 5 describes the population balance model used to simulate the aggregation

kinetics measured in the experimental campaign. The model describes flocculant

mixing, aggregation, and (partially irreversible) breakage as a function of the process

variables covered by the experimental runs. Wherever possible the model equations

were kept physically realistic, describing the dominant aspects of the process, but

without including so many terms and effects that the model became unwieldy and

difficult to incorporate into CFD.

Chapter 6 describes the relationship between the aggregate size and the hindered

settling rate as a function of the process variables considered above. A relationship was

developed from established descriptions of the hindered settling rates of non-aggregated

suspensions (Stokes 1851, Richardson and Zaki 1955), with additional terms describing

the effect of aggregate porosity.

Chapter 7 brings the various aspects together, and explores ways that the model may be

used to optimise the performance of a full-scale unit. Various possibilities were

considered, in particular the use of recycle to alter the feed solid concentration and

maximise the solid settling flux (or unit throughput).

Page 16: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

15

2. Literature Review 2.1 Introduction This review covers a range of topics directly relating to the aggregation and settling of

particles, as occur in thickeners/clarifiers in mineral processing circuits. In addition

there is a considerable quantity of relevant material relating to similar processes, in

particular aggregation and settling behaviour in water treatment plant, or occurring

naturally in river estuary systems. These systems have been actively studied for over a

century, and intensively since the 1950s, resulting in a daunting body of literature.

Despite this, recent advances in computing and instrumentation make this an active area

of research.

This review begins with a description of different collision mechanisms that may cause

aggregation, and considers how the addition of flocculants or coagulants leads to the

formation of stable aggregates. The second section considers the porosity of the

aggregates and how it may be described using fractal geometry. Aggregates are usually

highly porous, which significantly impacts on several key areas. Porous aggregates

have an increased collision radius, increasing the aggregation rate compared to solid

particles of the same mass. However, highly porous aggregates may have a reduced

number of primary particle contact points, reducing aggregate strength, leading to an

increased breakage rate. In addition, porous aggregates occupy an increased volume,

increasing the suspension viscosity, while also decreasing the subsequent hindered

settling rate.

Fluid dynamics is of central importance to both aggregation and settling and is

discussed at various points throughout this thesis. However, a separate section is

included in the review, describing fundamental relationships like viscosity and shear

rate, and then considering some of the deviations from idealised behaviour that are

likely to impact on the aggregation and settling rates.

Aggregate breakage is reviewed next, with separate sections describing the causes of

breakage, the effect of aggregate size on the settling rate, and the number and size of the

daughter fragments produced. Once the aggregation and breakage rates have been

described, the overall process can be simulated to describe the change in aggregate size

distribution through time. This is usually done with a population balance, a

Page 17: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

16

mathematical model that accounts for the number of particles in each size range while

also conserving mass. A population balance is a dynamic model, written as a closed set

of differential equations. These equations are usually intractable analytically, and are

normally solved numerically as an initial value problem.

The final section describes the gravity settling of the solid particles. This section is

divided into the usual categories depending on the solid fraction. At a low solid fraction

the particles are taken to settle independently of each other (free-settling), which is

normally taken to be the case in the clarification zone at the top of a thickener/clarifier.

At an intermediate solid fraction the particles are constrained to settle as a loosely

connected network, reducing the settling velocity. This is referred to as

hindered-settling, and is frequently taken to limit the thickener throughput in the

thickener settling-zone beneath the feedwell. At a high solid fraction, as found in the

thickener bed region, the solid particles are in intimate contact, resulting in a network

structure that can transmit force up through the sediment network, further reducing the

settling rate.

2.2 Aggregation 2.2.1 Particle collision mechanisms

Aggregation results from successful particle or aggregate collision. Particle collision

was first described by Smoluchowski (1916, 1917) as a combination of three possible

mechanisms: differential settling, Brownian motion and fluid shear (Figure 2-1). The

settling velocity of a particle in viscous fluids increases as a function of its diameter

(Equation 1-3), with larger particles settling more rapidly, sweeping smaller particles

from the suspension. Larger particles also have a larger collision radii, giving

(Smoluchowski 1916, 1917, as referenced in: Stumm 1992, Gregory 1989):

( )( ) ji3

jilsij aaaa9

g2−+ρ−ρ

μπ

=β 2-1

where: βij = Aggregation kernel (m3s-1) ai = Radius of the ith particle (m) g = Gravity (9.8 m s-2) ρ = Density (s = solid, l = liquid) (kg m-3) μ = Fluid viscosity (N s m-2)

Page 18: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

17

Figure 2-1: Mechanisms of particle collision.

Smoluchowski also described particle collision due to Brownian motion (perikinetic

aggregation) (Stumm 1992, Gregory 1989):

ji

ij aa3KT2μ

=β 2-2

where: K = Boltzmann’s constant (1.38 ×10-23 J K-1) T = Absolute temperature (K) and for laminar fluid shear (orthokinetic aggregation) (Stumm 1992, Gregory 1989): ( )3ji3

4ij aa +γ=β 2-3

where:

γ = Laminar shear rate (s-1) The overall rate of particle collision will be a combination of the above three kernels

(Smoluchowski 1917, Swift & Friedlander 1964, Feke & Showalter 1983, Han and

Lawler 1992, Adachi et al. 1994), however, in practice one mechanism is likely to

dominate and to reduce mathematical complexity usually only one kernel is used.

The Peclet number (Pe) is the ratio of the viscous and thermal forces and is commonly

used to demonstrate that shear aggregation dominates Brownian aggregation in

experimental studies (Oles 1992):

Differential settling

Brownian motion (perikinetic)

Fluid shear (orthokinetic)

Page 19: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

18

KT

Ga6Pe3πη

=

2-4

where: η = Dynamic viscosity (N s m-2) G = Average shear rate (s-1) a = Particle radius (m) K = Boltzmann’s constant (1.38 × 10-23 J K-1) T = Temperature (K)

A Peclet number greater than unity indicates that shear is dominant (Hansen & Ottino

1996), with most workers reporting much higher numbers: >120 (Serra & Casamitjana

1998), >160 (Oles 1992), 1300-21000 (Krutzer et al. 1995), >1 Wistrom and Farrell

(1998) (however they acknowledge some Brownian aggregation).

Differential settling only becomes significant in quiescent systems and when the

particles are widely differing in size (Lawler 1993), and it is generally accepted (Tambo

& Watanabe 1979, Wagberg & Lindstroem 1987), that fluid shear is the dominant

aggregation mechanism in industrial-scale turbulent flows (de Boer et al. 1989a, Brunk

et al. 1998). However, Smoluchowski’s description (Equation 2-3) of collision due to

fluid shear assumes laminar flow, and is not appropriate for turbulent systems.

2.2.2 Turbulent collision, particles smaller than micro-scale

Camp and Stein (1943) substituted a term, G, the spatially averaged shear rate

(see section 2.4.2) and proposed:

( )3ji3

4ij aaG +=β 2-5

where:

μ=

ΦG 2-6

and: Φ = Energy dissipation rate per unit volume (kg m-1s-3)

μ = Viscosity (kg m-1s-1, or N s m-2)

[units G = ( (kg m-1s-3)/(kg m-1s-1) )0.5 = (s-2)0.5 = s-1 ]

Camp and Stein’s kernel has been critiqued repeatedly. Glasgow and Kim (1986) point

out that the equation only strictly applies to isotropic turbulence (where the shear is the

Page 20: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

19

same in all directions) and may not be applicable to systems like stirred tanks. The

shear rate in vessels like stirred tanks may vary considerably, both in direction and

magnitude (Koh et al. 1984). For example, the shear rate at the impeller tips is typically

5-10 times the amount in the rest of the tank (Perry and Green 1997, 18-6). Koh et al.

developed a multi-component model for the effective shear rate through a stirred tank

using the first moment of the shear rate distribution. They suggest that the type of shear

(laminar or turbulent) is not important, but rather it is that the extent of mixing is

critical.

Cleasby (1984) critiqued Camp and Stein’s work on the basis that the absolute viscosity

(μ) is not an appropriate parameter for use in turbulent systems and goes on to suggest

that the use of G be restricted to when the particle size is below the Kolmogoroff micro-

scale. Above the micro-scale, Cleasby suggests βij ∝ ∈2/3, where ∈ is the local energy

dissipation rate per unit mass, as suggested by Saffman and Turner (Equation 2-7).

Camp and Stein’s work has also received criticism from Clark (1985), who performed

tensorial analysis and concluded that Camp and Stein’s conceptualisation is

“fundamentally incorrect, since they essentially require that a three-dimensional flow in

general be represented by a single two-dimensional flow.” (Their italics).

Despite these criticisms, a similar but more rigorously derived formulation by Saffman

and Turner (1956)(Table 2-1):

( )3jiij aaG294.1 +=β 2-7 and:

ν∈

=G 2-8

where: ∈ = Local energy dissipation rate per unit mass (m2s-3) ν = Kinematic viscosity (m2s-1) [units G = (m2s-3 / m2s-1)0.5 = s-1] has reached general acceptance (Clark 1985, Cleasby 1984, Gregory 1989), provided

(Clark 1985):

Page 21: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

20

1. The particles are small relative to the Kolmogoroff micro-scale. 2. The particles are neutrally buoyant and spherical. 3. The shear is high Reynolds number isotropic turbulence. 4. There are no large spatial differences in the local energy dissipation rate. 5. The coagulation is slow. 6. There are no hydrodynamic or colloidal interactions between particles.

Particles smaller than the micro-scale are contained within even the smallest eddies, and

particle collision is primarily caused by localised fluid shear (De Boer 1989a). For

particles in this range there is general agreement (Clark 1985, Cleasby 1984, Spicer

1997, Kusters 1991, Krutzer et al. 1995) that Saffman and Turner’s Kernel

(Equation 2-7) is appropriate, except if there is a significant hydrodynamic interaction

(Section 2.1.2). Saffman and Turner’s kernel appears to be the most widely accepted

(Tables 2-1 & 2-2).

The condition of neutral buoyancy is typically largely ignored, even by Saffman and

Turner, who described the motion of water droplets in clouds. Experimental studies are

typically performed with neutrally buoyant latex particles (Oles 1992, Flesch et al.

1999), although aggregates are also typically highly porous (see Sections 2.2 & 2.6.1),

reducing their effective density considerably.

Various other collision kernels have also been proposed, with different constants. For

example: Chin et al. (1998) calculated the constant as 2.3, apparently in agreement with

Pearson et al. (1984). Delichatsios and Probstein (1975) calculated from mean free

paths a constant of 0.4 (de Boer 1989b). Levich (1962) calculated a much larger

constant of 13.8 (Delichatsios & Probstein 1975) or 9.24 (de Boer 1989a) using a

turbulent diffusion model. Delichatsios and Probstein (1975) suggest that the

discrepancy is likely to be caused by: “the inability of a steady diffusion model to

correctly describe the relative motion of particles in turbulent flows due to the non-

Markovian nature of the turbulent dispersion process”.

de Boer (1989a) compared various aggregation kernels with experimental results from

stirred tank experiments where 0.74 μm polystyrene particles were coagulated with

sodium chloride. In all cases the theoretically predicted aggregation rates exceeded the

Page 22: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

21

experimental result. In contrast, Spicer and Pratsinis (1996) used Saffman and Turner’s

aggregation kernel in conjunction with a breakage kernel to successfully model

coagulation of 0.87 μm polystyrene particles with alum (Figure 2-11) without requiring

a capture efficiency term.

Although Saffman and Turner’s kernel is only strictly correct for particles smaller than

the Kolmogoroff micro-scale, it has been used successfully for somewhat larger

particles. Delichatsios (1980) suggests that the particles may be as much as ten times

the micro-scale, compared to Kusters (1991) who suggests six times larger.

2.2.3 Turbulent collision, particles larger than micro-scale

Particles significantly larger than the Kolmogoroff micro-scale are too large to be

completely contained in the small scale eddies, and inertial effects become important.

For particles larger than the micro-scale Cleasby suggests:

32

ij ∝∈β 2-9 may be more appropriate than:

21

ij ⎟⎠⎞

⎜⎝⎛ν∈

∝β 2-10

and goes on to suggest that the energy used to mix water treatment flocculators could be

reduced as a cost saving. Cleasby’s arguments were both theoretical and based on the

correlation with experimental data from Argaman and Kaufman (1968, 1970). Cleasby

also suggested that turbulent eddies larger than the micro-scale were responsible for

flocculation in industrial systems. Conversely, Casson and Lawler (1990) performed

experiments with an oscillating grid and concluded that flocculation was caused by

eddies of a similar size to the particles, and that larger eddies had little effect, except to

prevent the settling of aggregates in the flocculator.

Cleasby’s suggestion of ∈2/3 is at odds with various other calculations that generally

suggest ∈1/3 in the inertial sub-range (Delichatsios & Probstein 1976, Kusters 1991).

Kusters also comments that no analytical expression for inertial effects has been

Page 23: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

22

formulated for the inertial sub-range for solid/liquid systems, but quotes Kuboi et al.

(1972) that the collision frequency can be calculated from:

( )3

731

jiij aa87.6 +∈=β 2-11 where: ai = radius of the ith particle which is similar to the kernel proposed by Delichatsios and Probstein (1975):

37

31

31

31

jiij vv41.1 ⎟⎠⎞⎜

⎝⎛ +∈=β 2-12

where:

vi = Volume of ith particle Kruis and Kusters (1997) recently proposed the general case:

( ) 2shear

2.accel

2ji)shear.accel(ij wwaa

38 2

1

++⎟⎠⎞

⎜⎝⎛ π

=β + 2-13

where: waccel. = Relative particle velocity of particles due to acceleration (inertial effect) wshear = Relative particle velocity of particles due to fluid shear

In general it appears that there is far less agreement about the aggregation rate in the

inertial sub-range than in the viscous sub-range. A full description becomes even more

complex if the particles are significantly heavier than water and particle inertia also

affects the collision rate. The problem is exacerbated by a general lack of experimental

evidence. Delichatsios and Probstein (1975) did not get good experimental agreement

with their proposed equations and concluded that the error was due to aggregate

breakage.

Abrahamson (1975) considered the turbulent macro-scale and suggested:

( ) 2

j2i

2jiij VVaa5 ++=β 2-14

for large particles in industrial level turbulent flow, i.e. in the macro sub-range. Vi is

the particle’s mean square velocity:

Page 24: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

23

2i

2

i V/5.11VV∈τ+

= 2-15

and: V = mean square velocity of the fluid (m s-1) τi = particle relaxation time (s)

The relaxation time is the time it takes a particle to accelerate to the same velocity as

the fluid after a change in the fluid velocity (Brunk et al. 1998).

2.2.4 Summary - particle collision

Particle collision can be caused by a combination of Brownian motion, differential

settling and fluid shear. However, it is generally accepted that in turbulent industrial

feed streams fluid shear is the dominant collision mechanism. While the collision rate

in laminar flow is relatively easy to describe (Equation 2-3), collision in turbulent flow

is far more complex and is influenced by the size of the particle in relation to the size of

the turbulent eddies. Despite this uncertainty, the aggregation kernel by Saffman and

Turner (1956) (Equation 2-7) is widely accepted in the literature (Table 2-2). However,

Saffman and Turner’s kernel typically overestimates the aggregation rate compared to

experimental data, and a collision capture efficiency term is normally introduced to

account for the discrepancy.

Page 25: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

24

Table 2-1: Aggregation kernels

Reference Kernel Comment/References Smoluchowski 1917

( )3ji3

4ij rr +γ=β #1, Table 2-2

Laminar, widely referenced

Camp & Stein 1943

( )3ji34

ij rr21

+⎟⎟⎠

⎞⎜⎜⎝

⎛μΦ

=β #2, Table 2-2 Widely referenced

Saffman & Turner 1956

( )3jiij rr158 2

1

+⎟⎠⎞

⎜⎝⎛ν∈π

=β #3, Table 2-2

Widely referenced

Levich 1962

( )3jiij rr24.921

+⎟⎠⎞

⎜⎝⎛ν∈

=β Shamalou & Titchener-Hooker 1993

Kuboi et al. 1972

( )37

31

jiij aa87.6 +∈=β Inertial, Kusters 1991

Delichatsios & Probstein 1975

( )3jiij rr77.021

+⎟⎠⎞

⎜⎝⎛ν∈

37

31

31

31

jiij vv41.1 ⎟⎟⎠

⎞⎜⎜⎝

⎛+∈=β

Viscal Shamalou & Titchener- Hooker 1993 Inertial

Abrahamson 1975

( ) 2j

2i

2jiij rr0.5 ω+ω+=β

ω = mean sq. velocity of particle

De Boer 1989 a, b

( )3jiij aa4.021

+⎟⎠⎞

⎜⎝⎛ν∈

( )37

31

jiij aa15.2 +∈=β

Viscal Kuboi et al. 1972 is similar except 2.89 rather than 2.15. Inertial Delichatsios & Probstein 1975

Shamalou & Titchener-Hooker 1993

( )37

31

ji2ij rrA +∈=β

Inertial.

Kruis & Kusters 1997

( ) 2shear

2acc

2jiij rr

38

ω+ω+π

=β Viscal & inertial

Chin et al. 1998

3

jiij31

312

1

vv3.2 ⎟⎠⎞⎜

⎝⎛ +⎟

⎠⎞

⎜⎝⎛ν∈

α=β Pearson 1984

Flesch et al. 1999

3D3

3oij

fD1

j

fD1

i

21

f rrkR294.1 ⎟⎠⎞

⎜⎝⎛ +⎟

⎠⎞

⎜⎝⎛ν∈

=β−

Jiang & Logan 1991

Vanni 2000

( ) ( ) 2j

2i

2ji

3fp

ij rrrr27.1 4

1

−+⎟⎟⎠

⎞⎜⎜⎝

⎛ν∈

μ

ρ−ρ=β

Inertial, attributed to Saffman & Turner 1956

Page 26: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

25

Table 2-2: References to simple shear kernels (Table 2-1, #1, #2 & #3) Reference Comment Abrahamson 1975 Effect of larger aggregates Adachi et al. 1994 Akers et al. 1987 Argaman & Kaufman 1968, 1970 Atteia 1998 α = 1 × 10-4 - 0.1 Brunk et al. 1998 φexp = 0.002-0.12, α = 0.05-0.6 Burban et al. 1989 α = 0.15-0.3 Casson & Lawler 1990 Cleasby 1984: ∈2/3 rather than G, φexp = 1.05 × 10-4 Chin et al. 1998 α = 0.05-0.65, φexp = 1 × 10-4 Cleasby 1984 Suggests ∈2/3 rather than G De 1989a, b α ∼ 0.25, range 0.14-0.51, φexp = 2 × 10-5 – 1 × 10-4 Delichatsios & Probstein 1975 φexp = 1 × 10-3 – 6 × 10-3 Ducoste and Clark 1998 α = 0.006 Flesch et al. 1999 φexp = 1.4 × 10-5 Gardner & Theis 1996 Table, α = 0.0003-1 Glasgow & Lueche 1980 Gonzales & Hill 1998 α = 0.015-0.021 Gregory 1989 α = f(λ)CA

0.18 = 0.1-0.5 Gruy & Saint Raymond 1997 α = 0.05-1.4 Han & Lawler 1992 α = f(hydrodynamic effect) = 1 × 10-5 – 1 Higashitani et al. 1983 Kramer & Clark 1999 α = 0.005-0.025 Kruis & Kusters 1997 Krutzer et al. 1995 φexp = 1 × 10-4 Koh et al. 1984, 1987 α = 0.5 × 10-4 – 1 × 10-3 Kusters et al. 1987 α = 0.05-1, φexp = 3.1 × 10-5 – 2.1 × 10-4 Lawler 1993 α = 1E-4 – 0.5, φexp = 2 × 10-4 – 6 × 10-4 Lick & Lick 1988 Lu et al. 1998 α = 0 – 0.5, φexp = 0.5 % ∅degaard 1979 Serra & Casamitjana 1998 φexp = 2.5 × 10-5 Shamlou & Titchener-Hooker 1993 Spicer & Pratsinis 1996 α = 1, φexp = 8.3 × 10-5 Swift & Friedlander 1964 φexp = 3.45 × 10-5 – 8.6 × 10-4 Tambo & Watnabe 1979 α = 0 – 0.02 Thomas et al. 1998 Tomi & Bagster 1978 Whittington & George 1992 φexp = 1.18 g L-1

Wiesner 1992 φexp = 2.1 × 10-6 Wigsten & Stratton 1984 α = 0.2 – 0.33 Wistrom & Farrell 1998 α = 3 × 10-7 – 1 × 10-5, φexp = 50-100 mg L-1

Zeichner & Schowalter1979 φexp = 5 × 10-6 – 2.5 × 10-5 Vanni 2000

Page 27: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

26

2.2.5 Effect of short range inter-particle forces/bonding mechanisms

All the above kernels predict the rate of collision by assuming that there are no inter-

particle forces; the so-called rectilinear model (Lawler 1993). This may not be correct,

and various inter-particle forces may increase or diminish the effective collision radius

of the particles by forcing the particles to deviate from their previous trajectories.

These inter-particle forces are often used to calculate a capture efficiency term (α) for

insertion into one of the aggregation kernels above. However, inter-particle forces have

the greatest impact on solid spherical particles being coagulated with salt where the

particles must come very close together before adhesion. Inter-particle forces are less

important to bridging aggregation with polymers (Gregory 1989).

van der Waals and electrostatic forces

Prior to the addition of a flocculant, a suspension may already be aggregated to some

extent due to coagulation. Several forces can act between the particles, with the extent

of aggregation dependent on which forces are dominant. Deryagin and Landau (1941)

and Verwey and Overbeek (1948) independently combined the effects of van der Waals

and electrostatic forces between particles, leading to the so-called DLVO (Deryagin,

Landau, Verwey and Overbeek) theory. The electrostatic repulsion initially repels the

particles (first term of Equation 2-16), but on close approach van der Waals force

dominates (second term)(Figure 2-2).

d12aA

zeKTa32V 111d

2

1t −γ⎟⎠⎞

⎜⎝⎛∈π= κ− 2-16

where: Vt = Interaction energy a1 = Radius of particle

∈ = Permittivity of the medium z = Valency (symmetrical z-z electrolyte assumed) γ1 = Dimensionless function of surface charge κ = Deybe-Huckel parameter d = Separation distance

Page 28: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

27

VA

VT

Energybarrier

Secondaryminimum

Primaryminimum

Inte

ract

ion

Ener

gyd

VE

Figure 2-2: Energy of interaction between two particles, VA is the van der Waals attraction, VE is the electrical repulsion, and VT the total interaction energy. (adapted from Gregory 1989).

Hydrodynamic force

Another force that may cause repulsion between aggregating particles is the so-called

hydrodynamic (or viscous) interaction (Lawler 1997, Gregory 1989). When the

particles come close together the final approach is resisted by a thin film of fluid

separating them. The hydrodynamic force has been used to calculate a capture

efficiency term (van de Ven & Mason 1977, Gregory 1989, Potanin & Uriev 1991,

Adachi et al. 1994):

18.0

AC)(f λ=α 2-17 where:

λ = Characteristic wavelength (m) CA = A/36πμGa3 A = Hamaker constant μ = Viscosity (N s m-2) G = Mean shear (s-1) a = Particle radius (m)

The hydrodynamic force is largely overcome by the van der Waals force of attraction at

very close approach (van de Ven & Mason 1977a, Gregory 1989, Spielman 1978,

Delichatsios 1980). Also, in practice the particles are irregularly shaped and rapidly

aggregate into porous aggregates that allow the fluid to flow through the structure,

reducing the hydrodynamic effect considerably (Adachi 1994, Wolynes & McCammon

1977, Adler 1981b, Torres et al. 1991a, Chellam and Wiesner 1993).

Page 29: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

28

2.2.6 Coagulation with salt

The electrostatic repulsion can be dramatically reduced by the addition of an electrolyte

to the solution, a phenomena that has been used widely to aggregate particles in water

treatment plants. The electrostatic repulsion is reduced because the additional

electrolyte increases the ion concentration in the double layer, compressing it (Stumm

1992). In addition, specific ion adsorption may also occur, directly altering the surface

charge. The critical minimum concentration of salt required for aggregation can be

calculated approximately from (Gregory 1989):

62

4

f ZAKc γ

= 2-18

where: K = Constant related to the fluid γ = Dimensionless function of surface charge A = Hamaker's constant Z = Valency

The suspension is completely destabilised when enough salt is added that the

electrostatic repulsion becomes negligible and all collisions result in aggregation

(Gregory 1989). The fact that the critical concentration is inversely proportional to the

valency raised to the sixth power (it drops to Z2 with lower surface potential - Stumm

1992, p. 266) is exploited by using multivalent ions like Ca2+, Al3+ or Fe3+. The cation’s

charge is more important than the anion’s charge for negatively charged particles, i.e.

CaCl2 is a better coagulant than Na2SO4 (Gregory 1989).

2.2.7 Flocculation with polymer

Polymer flocculants give better performance than salt coagulants and have largely

replaced them in mineral processing applications (Gregory 1989, Adachi et al. 1994,

Kim & Glasgow 1987). Two models of flocculation by polymers have been proposed.

In 1952 Ruehrwein and Ward proposed the bridging model (Figure 2-3) where high

molecular weight polymers can bridge the gap between two particles and cause

aggregation (Chen 1998). This is supported by the evidence that the polymer can be

uncharged, or even carry the same charge as the particles, and still cause aggregation

(Chen 1998, Stumm 1992). However, cationic polymers can also be used so that charge

neutralisation also aids the aggregation. The particular mechanism may be pH

Page 30: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

29

dependent (Lindquist & Stratton 1976, Ditter et al. 1982) with bridging dominating

above pH 9 due to the low cationic charge of the polymer at higher pHs (Chen 1998).

(a) (b) Figure 2-3: Bridging model of polymer flocculation (a), and re-stabilisation from excess polymer (b). (adapted from Gregory 1989)

The second model for polymer flocculation was proposed independently by Kasper

(1971) and Gregory (1973) and is referred to as the electrostatic patch model

(Figure 2-4). In this model a polymer of opposite charge to the particle is thought to

adsorb onto the particle, resulting in “patches” of opposite charge (Gregory 1989).

+--

--

--

-

--

--

-

---+++++++++++ - -

-- --

-

-+++

+++

+++++-

--

- ++

+-

++++++++

++++-- -

---

--

-

-

+++++++++++

- ----- -

---

- ---

-

--

-

- - --

-

-

-- -

+++++++++++

+++++++++

+++++++++

++--- -

-----

- - --

--

-- ---

---- -- - ----------- -

Figure 2-4: Electrostatic patch model (adapted from Gregory 1989)

Various relationships between surface coverage (θ) and capture efficiency (α) have

been proposed. For example La Mer and Healy (1963) proposed:

α = θ(1-θ) 2-19

implying that the aggregation rate should be highest at 50 % coverage. Unfortunately,

this proposal is not supported by experimental evidence (Hsu et al. 1995, Glasgow &

Page 31: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

30

Lueche 1980). Hogg (1984) took into account the possibility of particle re-orientation

and proposed:

α = 1 - θ2n - (1-θ)2n 2-20

where n is the number of adsorption sites per particle. When n = 1, Equation 2-20

collapses to (Hsu et al. 1995):

α = 2θ(1-θ) 2-21 which Hsu et al. (1995) state is the correct formulation of 2-19 when the statistics of

collision between bare and occupied patches are properly taken into account. Various

even more complex relationships have been proposed, see for example Berlin and

Kislenko (1995) or Hsu et al. (1995).

There appears to be experimental evidence to support both the bridging and the

electrostatic patch models. Mabire et al. (1984) flocculated silica with a copolymer of

acrylamide and acrylate and found maximum flocculation performance at intermediate

flocculant dosages, concluding that the electrostatic patch model was appropriate. The

optimal performance was found at intermediate coverage due to steric stabilisation

(Figure 2-3b) where at high polymer dosages the particles become totally covered with

polymer (Gregory 1989). Adachi et al. (1994) found that polymer (PEO) flocculation

occurred faster than coagulation by salt and concluded that the polymer chains were

bridging the gap between particles and increasing the collision radius.

Lindquist and Stratton (1976), and Ditter et al. (1982) found that the bridging

mechanism dominated above pH 9 and 9.5 respectively, while the electrostatic

attraction dominated at lower pHs. Leu and Ghosh (1988) found that the optimal

flocculant dosage was predominantly related to the charge density of the polymer rather

than the length of the polymer, supporting the electrostatic patch model.

Ditter et al. (1982) suggest the mechanism can be determined by briefly increasing the

fluid shear and noting whether the aggregate size returns to its original value. If the

original aggregate size returns, the electrostatic patch mechanism can be assumed, if

not, bridging is assumed. Chen and Doi (1989) suggest a similar method involving

Page 32: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

31

modelling. Both these suggestions are based on the assumption that polymer

breakage/re-conformation affect bridging but not electrostatic patch aggregation.

In summary, although the electrostatic effect of the polymer is likely to contribute to the

bonding mechanism, polymer bridging is usually considered to be dominant. This is

supported by experimental evidence of the irreversible effect of shear on the aggregate

size (see Section 2.2.8) and the improved performance generally given by a flocculant

of higher molecular weight.

2.2.8 Loss of flocculant activity by polymer degradation

Although aggregates formed by coagulation with salt are comparatively weak,

aggregate breakage is generally found to be reversible, allowing the daughter fragments

to re-aggregate. However, with polymer flocculants there is considerable evidence that

broken aggregates have a reduced ability to re-aggregate (Pelton 1981, Sikora &

Stratton 1981, Glasgow & Lueche 1980, Chen et al. 1990) except when the mechanism

is charge neutralisation (Chen et al. 1990, Ditter et al. 1982). The bridging performance

of polymers decreases due to scission or re-conformation of the polymer due to shear

(Gregory 1987a, Stratton 1983) and Ditter et al. (1982) suggested that the breakage

reversibility in shear may be used to determine the mechanism.

The strength of the carbon-carbon bond is only about 10-8-10-9 N, and Gregory (1997)

suggests that several bridges are required between particles to provide sufficient

strength to prevent breakage. Pelton (1981) has proposed a decreasing aggregate

strength depending on the history of the particle surface through time. Initially, after

polymer addition all the surface is considered active. However, after aggregation and

subsequent breakage the previously bonded surface is considered inactive (Figure 2-5).

As the aggregation/breakage process continues the adsorbed flocculant progressively

loses its activity, decreasing the aggregate strength.

Page 33: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

32

Figure 2-5: Diagram showing polymer degradation after contact and subsequent breakage (adapted from Pelton 1981).

The loss of bonding strength after breakage has been observed experimentally by Yeung

et al. (1997), who measured the strengths of individual polymer flocs with a cantilever

balance arrangement and reported that the re-aggregated flocs had a strength of only

10 % that of the original flocs. The decrease in the aggregate size with continued shear

appears to have received little attention in the literature, particularly the modelling

literature that has typically considered coagulation rather than flocculation.

2.3 Aggregate porosity Aggregates are a loose assemblage of primary particles, enclosing considerable void

space within the aggregates structure (Figures 2-6 & 2-7). The porosity of the

aggregate has several important effects on the aggregation process. Firstly, it increases

the effective collision radius of the aggregate, thereby increasing the collision rate

(Flesch et al. 1999). Secondly, it decreases the settling rate of the aggregate, compared

to a more compact structure (Gregory 1989). Thirdly, it affects the strength of an

aggregate, since a more compact structure is likely to have more particle-particle

contacts (Gregory 1989, Dobias 1993, Bagster 1993). Finally, a porous aggregate may

also increase the thickener bed height for a given underflow solid (Healy et al. 1994,

Svarovsky 2000).

ATTACHED DETACHED

Large sphere Large sphere

Polymer flocculant Inactive

area

Page 34: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

33

Figure 2-6: Metallic aluminium particles (~ 5 μm) flocculated with high molecular weight polyacrylate, image is 880 × 660 μm.

Aggregate porosity is typically found to increase with the aggregate size (Figure 2-7)

(see also section 2.7.1) and is usually described by fractal geometry. Various

approaches are possible, however the one most widely used (and directly useful) is

based on the aggregate’s mass and diameter. In this case the aggregate volume is taken

to increase with the cube of the diameter as with solid objects:

3

aggagg kdV = 2-22 where: Vagg = Volume of aggregate (m3) k = Constant, (π/6 for spheres) dagg = Diameter of the aggregate (m) but the aggregate mass increases at some lower, fractional, (hence fractal) power,

giving:

fDaggs dm ∝ 2-23

where: ms = Mass of solid in aggregate (kg) Df = Mass-length fractal dimension (1 [thin rods] ≤ Df ≤ 3 [spheres etc])

The aggregate density (ρagg) is (see also Section 2.7.1 & Appendix, Figure 2-18) (Mills

et al. 1991, Jiang & Logan 1991, Kusters et al. 1997, Serra & Casamitjana 1998):

Page 35: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

34

3D

p

aggsagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ=ρ 2-24

where: dagg = Effective diameter of aggregate (m) dp = Primary particle diameter (m) Df = Fractal dimension

ρ = Density (agg = aggregate, p = primary particle) (kg m-3)

The decreased density with aggregate size is usually rationalised on the basis of

increasing voidage as large aggregates are formed from aggregate/aggregate collisions

later in the aggregation process (Figure 2-7).

Figure 2-7: Pictorial (2D) representation of self-similar fractal aggregation showing that the density drops with the ratio of the aggregate to primary particle size (adapted from Meakin 1988).

Typically, the fractal dimension is assumed to be D3 (3-dimensional, i.e. mass-length),

however other systems are used (e.g. area, D2, or perimeter based, Dpf) depending on

the method measurement or simulation. Clearly it is important to distinguish between

them, for example Serra and Casamitjana (1998a) calculated: D3 = 2.6, D2 = 1.98,

D1 = 1.06, Dpf = 1.12 for the same aggregate structure.

Computer simulation of fractal aggregates has been popular and several models have

been proposed. Vold (1963) used what has become referred to as a “diffusion limited”

type model (Meakin 1988) where a single primary particle has successive primary

particles randomly added to build up an aggregate using a Monte-Carlo type simulation.

Page 36: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

35

This type of simulation typically gives a Df of 2.75 (Gregory 1989), higher than

typically found in practice (Table 2-3).

Meakin wrote a series of papers describing a diffusion limited cluster-cluster

aggregation that leads to a lower Df of around 1.75 (Gregory 1989), the same as found

experimentally by Weitz et al. (1985) because extra voidage is introduced when

aggregates combine, as opposed to single particle addition. The fractal dimension is

somewhat lower than typical, and the model has been extended to allow for

reorganisation during or after aggregation (Meakin 1985), leading to higher density

aggregates. This has been supported experimentally by Sonntag and Russel (1986),

who formed aggregates very slowly under Brownian conditions, and then increased the

mass-size fractal dimension from 2.2 to 2.5 by shearing in a Couette device.

Higher fractal dimensions are also produced if the aggregation is chemically rather than

diffusion limited, as attaching particles have a chance to try various positions

(Weitz et al. 1985), presumably preferring a position that leads to higher particle-

particle contact and hence higher density. Aubert and Cannell (1986) aggregated silica

slowly to get Df = 2.08, and rapidly to give 1.75. Similarly, Weitz et al. (1985) found

Df = 2.05 and 1.75 for reaction or diffusion limited aggregation, respectively.

The above models assume that aggregation is performed under Brownian conditions.

Since shear aggregation is the dominant mechanism for flocculation on the

industrial-scale, various ballistic models have also been developed where the colliding

particles are assumed to travel in straight lines (Gregory 1989). Tence et al. (1986)

achieved good agreement between the measured Df (1.88) of aggregated iron particles

and a ballistic cluster-cluster model that predicted Df = 1.91. Conversely, Serra and

Casamitjana (1998a) found that the fractal dimension was independent of the shear rate

or volume fraction, and Gregory (1989) notes that it is not affected by polydispersity of

particles.

Fractal dimensions have also been measured experimentally for a variety of different

systems (Table 2-3), generally giving results in the range Df = 1.8-2.5. However,

detailed comparisons are difficult due to the wide range of measurement techniques and

substrates used.

Page 37: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

36

Table 2-3: Fractal dimensions reported in the literature Reference Fractal

dimension⊗ Type Method Material

Adachi and Ooi 1990

2.08-2.4 (area based) 2.00-2.27 (max diameter based)

D3 from total number of primary particles

Brownian DLA microphotography to measure area and max. diameter. Aggregate destruction by ultrasonication followed by particle counting

PSL particles 0.804 μm

Adachi & Kamico 1994

1.6-2.3 D2(?) from max. floc. φ

Video Latex, 0.93 & 2.02 μm

Allain et al. 1996 2.1-2.3 D3 from settling rates Aggregation by Brownian and differential settling. Settling rate (800 mm jar) vs. radius by microphotography

CaCO3.

Amal et al. 1990a 2.3 D3 from Malvern and rate change in PSD

Brownian DLA. Malvern and rate data to calculate Df

Haematite

Amal et al. 1990b 2.3 diffusion limited 2.8 reaction limited

As above As above As above

Danielson et al. 1991

1.58-1.72 D2 from number of primary particles and radius

Shear aggregation in chaotic and regular flows

Model

Haw et al. 1997 1.42 (average) D2 from weight (area) average and characteristic radius

2D CCDLA Model

Horne 1987 2.21 D3 from mass-length Brownian DLA. Light scattering

Skimmed milk/ calcium ions

Horwatt et al. 1992

2.5 - DLA 2.6 - RLA 2.0 - HCCA ∼ 3.0 - LTA ∼ 3.0 – Eden

D3 from mass and radius of gyration

Compares Df for variety of different models

Model

Huang 1994 1.83-1.97 D3 Couette flow, Malvern and settling data

Estuarine sediment

Jiang & Logan 1991

2.4-3.75 (not a typo) (shear) 1.61-2.31 (DS)

D3 from mass-diameter

Shear and DS using dimensional analysis of fractal scaling relationships

Model

Johnson et al. 1996

1.78-2.25 D3 from size - total particle No.

Gentle shear in stirred tank for different times to give different Df. Microscopy and ultrasonication/ particle counting

0.87 or 2.6 μm latex spheres/ NaCl

Jullien & Meakin 1989

1.89-2.13 D3 from 3D model

Ballistic CCA with restructuring

Model

Kolb et al. 1983

1.38 D2 from 2D model Brownian CCA Model

Kusters et al. 1997

2.5 ± 0.1 D3 from laser scattering

Stirred tank, Malvern particle sizer.

1 μm polystyrene spheres/ NaCl

Page 38: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

37

Table 2-3: Continued Reference Fractal

dimension⊗ Type Method Material

Lin et al. 1990 1.86 (DLCA) 2.13 (RLCA)

D3 from laser scattering

Brownian RLA or DLA

Colloidal gold or silica

Lin et al. 1989 2.10 (gold) 2.12 (silica) 2.13 (polystyrene)

D3 from laser scattering

Brownian RLA Colloidal gold, polystyrene or silica

Meakin 1983 1.45 -1.5 D2 from 2D model 2D Diffusion limited cluster-cluster/particle

Model

Meakin & Jullien 1988

1.80-2.18 (DL cluster-cluster) 1.95-2.19 (ballistic cluster-cluster) 2.09-2.24 (RL cluster-cluster)

D3 from 3D model Various models and effect of restructuring

Model

Neinmark et al. 1996

1.75 D3 from TEM. Combustion generated soot

Soot particles

Oles 1992 2.1 (initially) 2.5 (extended shear)

D3 from light scattering

Shear flow in Couette, Malvern Mastersizer

2.17 mm polystyrene spheres/ NaCl

Serra & Casamitjana 1998a

2.24 D3 from laser scattering and microphotography

Aggregation in shear (Couette) flow (laminar)

2 μm latex/ NaCl

Sonntag and Russel 1986

2.48 D3 from light scattering

Sheared aggregates in Couette (laminar)

Polystyrene/KCl

Spicer & Pratsinis 1996b

1.19-1.29 Dpf from microscopy and image analysis

Stirred tank (turbulent). Perimeter based fractal dimension

0.87 μm polystyrene/ alum.

Spicer et al. 1998 2.1-2.5 D3 from light scattering

Stirred tank. Malvern Mastersizer 0.87 μm polystyrene/ alum.

Stoll & Buffle 1998

2.03 (RLA - no polymer) 2.03-2.52 (polymer in rod conformation) 1.97-2.08 (polymer in coiled conformation

D3 from radius of gyration and number of primary particles

CCA with various polymer conformations

Model

Torres et al. 1991a

1.75-1.85 D3 PCDLA & CCDLA Model

Weitz et al. 1985 1.7-1.8 (DLA) 2.0-2.1 (RLA)

D3 calculated from TEM

Brownian aggregation either DLA or RLA depending on coagulant dosage

Gold particles/ pyridine

Wiesner 1992 Table Literature review ⊗. DLA = diffusion limited aggregation, RLA = reaction limited aggregation, PCDLA = particle cluster DLA, CCDLA = cluster-cluster DLA, LTA = linear trajectory aggregates, EDEN = Model for biological systems after Eden 1961.

Page 39: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

38

2.4 Fluid behaviour 2.4.1 Laminar flow, shear stress (τ), viscosity (μ), and fluid shear (γ),

The shear rate of a fluid is the difference in velocity between adjacent parts of the fluid

(Figure 2-8)(Perry & Green 1997) due to an applied shear stress.

x

y

Figure 2-8: Laminar shear

The shear rate (γ) is calculated from:

γ = dU/dy 2-25 where:

U = velocity (m s-1) i.e. dx/dt The shear stress (τ) is:

τ = F/A 2-26 where: F = Force (N or kg m s-2) A = Area (m2) and the viscosity (μ) is (Perry & Green 1997): μ = τ /γ 2-27 [units μ = kg m s-2 / m2 / m s-1 = kg m-1s-1, or N s m-2]

2.4.2 Turbulent flows, G, Kolmogoroff micro-scale, eddy spectrum

Turbulent flows occur when fluid inertia dominates viscous dissipation, leading to the

formation of turbulent eddies. Since turbulent shear is the dominant aggregation

mechanism on an industrial-scale, some way is required to describe the turbulence and

its effect on the flocculation process. Kolmogoroff (or Kolmogorov 1941) envisaged a

Page 40: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

39

cascade of turbulent eddies that transfer energy from the large, energy-containing

eddies, down to the small energy dissipating eddies (Figure 2-9).

1/(Eddy size)

Ene

rgy

spec

trum

Large eddies,Eulerian macro-scale

Kolmogoroff micro-scale

Viscous dissipation

Inertial subrange

Figure 2-9: Eddy energy spectrum (adapted from Shamlou, Titchener-Hooker 1993, see also Klute & Amirtharajah 1991, Glasgow & Lueche 1980).

The Kolmogoroff eddy scale is divided into three regions. The large macro-scale eddies

are of a similar size to the stirrer (Kusters 1991, Spicer 1997) or vessel (Gregory 1989)

and contain the bulk of the energy (Parker et al. 1972). However, they dissipate little

energy directly; instead transferring it down to smaller eddies in the inertial sub-range

(Kusters 1991). These eddies in turn pass the energy down to the smallest eddies

(Kolmogoroff micro-scale of turbulence), where the bulk of the energy is released as

heat through viscous dissipation (Cleasby 1984). The smallest eddies also have the

highest shear rate (Brunk et al. 1998). The Kolmogoroff micro-scale marks the

boundary between the inertial and viscous sub-ranges (Parker et al. 1972).

The smallest eddies have a characteristic length scale referred to as the Kolmogoroff

micro-scale of turbulence, which Kolmogoroff calculated from dimensional reasoning

(Kusters 1991):

41

3

⎟⎟⎠

⎞⎜⎜⎝

⎛∈ν

=η 2-28

Kolmogoroff time and velocity scales are also sometimes quoted (Kusters 1991):

Page 41: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

40

21

k ⎟⎠⎞

⎜⎝⎛∈ν

=τ 2-29

( )41

kv ∈ν= 2-30

where: η = Kolmogoroff micro-scale of length (m) τk = Kolmogoroff micro-scale of time (s) vk = Kolmogoroff micro-scale of velocity (m s-1)

ν = Kinematic viscosity (m2s-1)

∈ = Energy dissipation rate (m2s-3)

Similar relationships can be found by Delichatsios (1980), and Lu et al. (1998).

2.4.3 Deviations from ideal Newtonian behaviour

Fluids do not always display Newtonian behaviour, where the shear stress is

proportional to the shear rate. Various deviations are possible (Figure 2-10). This of

course means that it is important that the viscosity (τ/γ) is measured at an appropriate

shear rate, since the viscosity is only a constant for Newtonian fluids.

Figure 2-10: Rheograms of various fluid behaviour (adapted from Perry & Green 1997).

A variety of equations have been proposed to describe non-Newtonian behaviour

(Table 2-4), with the Herschel-Bulkey equation being widely used:

n

y Kγ+τ=τ 2-31

Where: τ = Shear stress (N m-2 or kg m-1s-2) τy = Yield stress (N m-2 or kg m-1s-2) K, n = Fitted parameters

Pseudo plastic (shear thinning)

Dilitant (shear thickening) Sh

ear

stre

ss

Newtonian

Bingham plastic

Shear rate

Page 42: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

41

Table 2-4: Equations used to describe stress(τ)/shear(γ) relationships.

Reference: Relationship: Source/comments:

Newtonian

μγ=τ

1, 2, 3 Newtonian fluids

Bingham 1922

γμ+τ=τ ∞y

1, 3 Bingham plastics

nKγ=τ 1, 2, 3 Shear thinning

ηγ=τ−τ y 2, Bingham plastic

Herschel-Bulkey ny Kγ+τ=τ

2, 3, Bingham plastic

Casson ( )2cy K γ+τ=τ

3, Non-linear plastic

Prandtl 1928, Eyring 1936 ⎟

⎠⎞

⎜⎝⎛ γ=τ −

BsinhA 1

2, Pseudo plastics

Philippoff 1935

( )2s

o

/1 ττ+

μ−μ+μ

τ=γ

∞∞

2, Pseudoplastics

Powell and Eyring 1944 ( )

nPE

PEoa t

)1tln(γ+γ

μ−μ+μ=μ ∞∞ 3, Time dependent

Sisko 1958 cba γ+γ+τ

2, Hydrocarbon greases

Moore 1959

( )

( )λγ+−=θλ

γλ+μ=τ

baadd

co

2, Thixotropic

Ritter 1961, Ritter & Govier 1970 DR

,ss

,so,sD

s2

o,s

,ss Kloglogo,

Klog −θ⎟⎟⎠

⎞⎜⎜⎝

τ−τ

τ+τ−=⎟

⎟⎠

⎞⎜⎜⎝

τ−τ

τ−τ

∞∞ 2, Thixotropic

Meter 1963 ( ) 1

m

oa

/1 −α∞

∞ττ+

μ−μ+μ=μ

2, 3 Pseudoplastics in laminar pipe flow

White & Metzner 1963 t

d2 j,irj,iaj,i ∂

τ∂θ+μ−=τ

2, Viscoelastic

Cross 1965

32

1o

aαγ+

μ−μ+μ=μ ∞

∞ 2, 3 Pseudoplastics

Cheng and Evans 1965 ( ) ( )λγ=

θλ

γγλ=τ 21 fdd ,f

2, Thixotropic

Ellis 1967 ( )ττφ−φ=γ −α 11o

2, Empirical

Sestak 1988 ( )λγ+τ+γ+τ=τ n11y

nyo KK

3, Thixotropic Liu & Masliyah 1996 μγ+γ=τ G 3, Viscoelastic Kranenburg 1999

)d3/(2

pym

f

25

⎟⎟⎠

⎞⎜⎜⎝

φφ

μγ=τ Fractal aggregates

1: Perry and Green 1997 2: Govier and Aziz 1972 3: Liu and Masliyah 1996

Page 43: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

42

2.4.4 Effect of suspended solid on viscosity

The viscosity of a suspension increases as a function of the solid volume fraction. In

1908 Einstein proposed a relationship based on the first terms of a series expansion:

( )...K1 Eos +φ+μ=μ 2-32

where: μs = Viscosity of suspension (kg m-1s-1, N s m-2) μo = Fluid viscosity (kg m-1s-1, N s m-2) KE = Constant (2.5) φ = Solid volume fraction [0,1] Einstein originally proposed that KE = 1, but later, famously, revised it to = 2.5.

Equation 2-32 is only correct for very dilute solutions, and various alternative equations

(Table 2-5) have been formulated by adding additional terms.

For suspensions of high solid fraction the viscosity is typically given as a function of

the maximum possible solid fraction (φm) in the form (Govier & Aziz 1972, Liu &

Masliyah 1996):

k

mos 1

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

−μ=μ 2-33

The maximum solid volume fraction is typically taken as 0.6-0.7 (Fleer & Scheutjens

1993, Strenge 1993, Schramm 1996, Bustos et al. 1999). At this point the particles are

taken to be in close contact (gelation) and form a continuous network that resists shear,

causing the viscosity to rise exponentially.

The particle shape and surface charge may also affect the viscosity. Various workers

(Greenburg et al. 1965, Mishra et al. 1970, Hiemenz 1986) report that smaller particles

tend to increase the viscosity, due to a film of water bound to the particle that increases

the effective volume occupied by the particle. Smaller particles have a larger surface

area (per unit mass), leading to an increased viscosity.

Irregularly shaped particles also tend to increase the viscosity (Shaw 1993, Hiemenz

1986), because of the increase in surface area and reduced packing efficiency. Ionic

strength and pH may also alter the viscosity of a suspension by modifying the surface

charge and particle-particle interaction (Eirich 1960, Greenburg 1965, Horsley et al.

1984, Farrow et al. 1989, Johnson et al. 2000).

Page 44: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

43

Table 2-5: Equations relating solid fraction to viscosity Reference: Equation: Solid fraction: Source/comments: Einstein 1908 ( )φ+μ=μ k1os k = 2.5 φ < 2-3 % Govier and Aziz 1972 Kunitz 1926

( ) ⎟⎟⎠

⎞⎜⎜⎝

φ−φ+

μ=μ 4os 15.01.0

φ < 10 %

Purchas 1981

Guth et al. 1936 ⎟⎟

⎞⎜⎜⎝

φ−φ−φ−φ+

μ=μ 2

2

os 6.9215.05.01

φ < 8 %

Vand 1949

Guth et al. 1936 ( )2os 1.145.21 φ+φ+μ=μ φ < 20 % Govier and Aziz 1972, Kruyt

1952 Vand 1947 ( ) ( ) ...KkrQ2KK

!21k1 2

122111os +⎟⎟⎠

⎞⎜⎜⎝

⎛φ⎥⎦⎤

⎢⎣⎡ −+++φ+μ=μ

low-intermediate Vand 1949

Vand 1948 ( )2

os 349.75.21 φ+φ+μ=μ Govier and Aziz 1972

Burgers & De Bruyn 1949

⎟⎟⎠

⎞⎜⎜⎝

⎛φ+

φ−μ=μ

5.111

os Kruyt 1952

Vand 1948 ( )32

os 2.1617.75.21 φ+φ+φ+μ=μ Kruyt 1952

Chong et al. 1971 ⎟⎟

⎞⎜⎜⎝

⎛φφ−

φφμ=μ

m

mos /1

/

Chiu and Don 1989

Gillespie 1983

( ) ⎟⎟⎠

⎞⎜⎜⎝

φ−φ+

μ=μ 2os 12/1

Thomas 1965

( ) ⎟⎟⎠

⎞⎜⎜⎝

φ+φ+φ+

μ=μ6.16exp00273.0

05.105.21 2

os 0 < φ < φm

Govier and Aziz 1972, Mishra et al. 1970

Frankel & Acrivos 1967

( )( ) ⎟

⎟⎠

⎞⎜⎜⎝

φφ−

φφμ=μ

31

31

m

mos

/1/125.1

0.8φm < φ < φm

Govier and Aziz 1972

Maron-Pierce-Kitano 1956, 1981

2

mos 1

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

−μ=μ

Liu and Masliyah 1996

Landel et al. 1963

5.2

mos 1

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

−μ=μ 0.4φm < φ < 0.95φm

Govier and Aziz 1972

Turian et al. 1997 ( )

( ) ⎟⎟⎠

⎞⎜⎜⎝

φφ−φφ

μ=μ β

α

m

mos /1

/

Mooney

φ−φ

=μs1

kln s Liu and Masliyah 1996, Kao

et al. 1975, Van Diemen et al. 1985

Krieger 1972

φ−

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

−μ=μ5.2

m

fos 1

Liu and Masliyah 1996, Van Diemen and Stein 1984

Taylor 1932

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

μ+μ

μ+μφ+μ=μ

p

52

pos 5.21

emulsions < 2-3 %

Govier and Aziz 1972

Shaw 1993 ( ) ( )φ+μ=μ k1ratio axial 2os non-spherical

Hiemenz 1986

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛⎥⎦

⎤⎢⎣

⎡ρρ

+φ+μ=μ1

2

2

b,1os m

m15.21

bound water

Phan-Thien & Graham 1991 ( ) ⎟

⎟⎠

⎞⎜⎜⎝

φφ−φφ−

φ+μ=μ 2m

mEos /1

/5.01K1

Liu and Masliyah 1996

Kranenburg 1999

)D3/(2

m

ps

f

52

⎟⎟⎠

⎞⎜⎜⎝

⎛φ

φμ=μ

Fractal aggregates

Liu & Masliyah 1996 ⎟

⎜⎜

φ−φ

φ+μ=μ

m

3y

osa

1

Liu and Masliyah 1996, Flocculated suspensions

Page 45: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

44

2.4.5 Multi-phase mixing

Mixing can be divided into macro-scale and micro-scale mixing. Macro-scale mixing is

the large-scale mixing of the bulk-phases in the mixing vessel. It is dependent on the

vessel size and geometry, and also the flow behaviour in the vessel. Most macro-scale

mixing theory has been developed for stirred tanks where the pumping capacity of the

impeller (Perry & Green 1997) determines the macro-scale mixing.

Due to the effect of geometry, macro-scale mixing is difficult to scale-down from

industrial to laboratory scale. Laboratory experiments in small vessels have short

circulation times and good macro-scale mixing (Perry & Green 1997). Laboratory

vessels also have lower spatial variation in shear rates and have far lower Reynolds

numbers (typically 5-25 times lower) (Perry & Green 1997).

However, to get good mixing on a molecular or small particle scale (< 100 μm),

micro-scale turbulence is also required (Perry & Green 1997). This is primarily

determined by the energy input for a given volume, simplifying scale-up from

laboratory to full-scale. Since micro-scale mixing is determined primarily by the power

input, laboratory systems are typically scaled on the mean energy dissipation rate: the

familiar G. While this may adequately describe the collision rate of particles in a

homogeneous suspension (micro-scale mixing), the rate of macro-scale mixing of the

flocculant and feed suspensions is likely to be overestimated by laboratory-scale

studies.

Mixing can also be described as a blend time. This is somewhat arbitrary (Perry &

Green 1997), but can be measured experimentally, for example by adding an acid or

alkali to a transparent vessel already containing a pH indicator. This is essentially

limited to laboratory situations, but for plant studies an electrolyte can be added and

conductivity measurements taken. Hence the blend time would be the time taken to

reach a given stability of the conductivity probe, for example 5 % of the original change

when the electrolyte was added (Perry & Green 1997). Electrical impedance

tomography (Brown et al. 1985, Webster 1990, Salkeld 1991, Dickin et al. 1992,

Williams & Simons 1992) systems are now becoming available and are likely to

Page 46: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

45

become useful for validating computational fluid dynamics predictions of macro-scale

mixing.

Uhl and Gray (1967) suggest the usual rule of thumb: “If the reaction half-time is five

times greater than the average loop circulation time, then perfect mixing is justified for

usual engineering precision”.

Nagata (1975) defines a mixing index as:

0

1Mσσ

−= 2-34

where: M = Mixing index σ = Concentration variation σo = Concentration variation at time = 0 and:

( )∫ −=L

0

22 dx)m(C)x(Cσ 2-35

where: C(x) = Concentration at point x C(m) = Mean concentration through vessel

Macro-scale mixing appears to have received little attention in the flocculation

literature, although there is general recognition that the polymer has to be well mixed

with the pulp (Gregory 1987, Bajpai & Bajpai 1995, Wigsten & Stratton 1984) to

prevent local overdosing and possibly steric re-stabilisation (Figure 2-3b). This usually

requires a reasonable level of fluid shear, however this tends to result in aggregate

breakage, and Elimelech et al. (1995) concluded that an intermediate shear level is the

best.

Elimelech et al. (1995) also note that aggregation is sensitive to the flocculant addition

point and suggest staged addition to improve flocculant distribution (as do Gregory &

Guibai 1991). Tapered shear rates are also suggested (Guibai & Gregory 1991,

Gregory & Guibia 1991, Ives & Bhole 1973). Flocculant solution used on an industrial-

scale is usually quite dilute, 0.1-1.0 % w/v, (Gregory 1987), which reduces its viscosity

and helps improve mixing (Stratton 1983).

Page 47: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

46

2.5 Aggregate breakage 2.5.1 Cause of aggregate breakage

Aggregate breakage is generally assumed to be caused by hydrodynamic forces on the

particle due to the turbulent flow, although occasionally breakage is attributed to

particle collision (Ham & Christman 1969, Burban et al. 1989). Two mechanisms are

proposed for breakage by fluid shear; fragmentation or surface erosion. As with

aggregation, the mechanism is thought to be dependent on the aggregate’s size in

relation to the Kolmogoroff micro-scale. When the aggregate is larger than the micro-

scale, fragmentation is assumed, due to the fluctuating dynamic pressure (Gregory

1989, Glasgow & Lueche 1980). Below the micro-scale, viscous forces dominate and

cause either fracture or erosion of small particles from the aggregate surface (Gregory

1989, Glasgow & Lueche 1980).

Healy and La Mer (1964) suggested primary particle erosion based on the surface drag

using Stokes’ equation. Glasgow and Lueche (1980) point out that Stokes’ equation

was derived for creeping flows, not the turbulent flow that Healy and La Mer

considered. Around the same time, Thomas (1964) proposed bulgy pressure

deformation based on existing theories of emulsion droplet breakage. Later, Parker

et al. (1972) suggested that aggregation would continue up to a characteristic stable

aggregate size for a given system. Breakage was assumed to be via erosion, with the

maximum aggregate size (dmax) and rate equations being:

nmax GCd = 2-36

where: G = Mean shear (s-1) n = Exponent, (typically 0.5-2, Tambo & François 1991) C = Floc strength coefficient Or, alternatively as a differential:

mBGK

dtdN

= 2-37

where: KB = Floc breakage coefficient (equations given by Parker et al. 1972) m = Exponent (typically 1.5-2, see text) N = Number of particles

Page 48: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

47

Kim and Glasgow (1987) argue against the erosion mechanism, suggesting that a range

of fragment sizes is more likely, and that eddies comparable to the Kolmogoroff micro-

scale contain too little energy to remove primary particles.

However, the general form of Equation 2-37 has proved popular in the literature (Akers

et al. 1987, Bemer & Zuiderweg 1980, Brakalov 1987, Chung et al. 1998, ∅degaard

1979, Serra & Casamitjana 1998b, Spicer & Pratsinis 1996a). Experimental values of

m vary from 1.6 (Spicer et al. 1996), 2.07 (Chung et al. 1998) and 1.89-5.62 depending

on solid fraction (Serra & Casamitjana 1998b).

This has a major impact on the aggregation process, because the aggregation rate is

generally regarded as being proportional to Gn where n ≤ 1 (i.e. Saffman & Turner

1956). That is, an increased shear rate increases the rates of both aggregation and

breakage, but breakage is dominant. This results in the typical observation

(Figure 2-11) that the aggregate particle size drops with increased fluid shear (Fair &

Gemmel 1964, Smith & Kitchener 1978, Spicer & Pratsinis 1996a, Serra & Casamitjana

1998b, Chung et al. 1998).

Data of Oles, 1992Population balancemodelα = 1y = 1.6±0.18

0 20 40 600

5

10

15

20

25

G=25 s-1

50 s-1

75 s-1

100 s-1

125 s-1

150 s-1

Dimensionless Time, Gφt

Dim

ensi

onle

ss M

ass M

ean

Floc

Dia

met

er, d

mm

/d1

A' = 0.0047±0.0002

Figure 2-11: Effect of shear on the aggregate size adapted from Spicer and Pratsinis 1996. Polystyrene particles coagulated with KCl.

Page 49: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

48

Aggregate porosity also has a significant effect on the strength of aggregates. As the

aggregates become more porous (lower fractal dimension) there are fewer particle-

particle contact points and the aggregate becomes more fragile (Gregory 1989). For a

given number of primary particles the aggregate also becomes larger and more likely to

break. Tomi and Bagster (1975) used a structural link model to simulate the

hydrodynamic forces on various aggregate structures and suggested dmax ∝ s-0.5, where s

is the velocity gradient. They also report that both Fair and Gemmel (1964) and Ritchie

(1955) suggested dmax ∝ 1/s.

Horwatt et al. (1992) modelled the breakage of a variety of fractal structures (DLA,

RLA, HCCA, LTA, EDEN – see footer Table 2-3) in simple shear flow. Fracture was

modelled to give either a planar or irregular fracture. Not surprisingly, the planar model

predicted a higher aggregate strength, and was considered less realistic. In general,

aggregates became weaker as the fractal dimension was reduced (more porous), but the

arrangement of the primary particles also influenced the aggregate strength.

Blunt (1989) simulated the hydrodynamic forces on fractal aggregates using a steady-

state Navier-Stokes model. The model used a series of small-amplitude sinusoidal

buckles and suggested that the hydrodynamic forces act mostly on the outer protruding

tips of the aggregate, with little effect on the inner core. This generated very high local

forces at specific points of the aggregate making “breakage or deformation of the

structure very likely”. Blunt also comments that as the shear force acts mostly on the

surface of the aggregate, and since there is a broad distribution of forces on the

aggregate, the aggregate may continue to break down over a considerable length of

time.

Figure 2-12 shows the results by Lee and Brodkey (1987), who viewed wood pulp

aggregate breakage in Couette flow with a variable speed camera (400-8000 fps). They

describe two main mechanisms. Firstly, global deformation, breakage and

fragmentation; and secondly a surface erosion mechanism. In 1991 Glasgow and Liu

used a stainless steel mesh to break polymer aggregates and followed the process using

a high-speed camera. Their results essentially mirrored those of Lee and Brodkey

(1987), and they observed that in virtually all cases aggregate breakage was initiated by

Page 50: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

49

vertical compression. This was often followed by surface erosion, however Glasgow

and Liu report that the eroded fragments were too large to be primary particles.

break

fragment

shed

stretch

disintegrate

string

doublet

thread

dispersed

fragments

aggregate

Figure 2-12: Breakage phenomena observed by Lee and Brodkey (1987).

In another photographic study, Glasgow and Lueche (1980) studied the breakage of pre-

formed clay-polymer aggregates by pumping them through a rectangular section glass

pipe. They modelled the decrease in number of the largest aggregate size range as

measured by a camera. Based on the fit between the modelled and experimental data

Glasgow and Lueche concluded that shear was responsible for breakage.

∅degaard (1979) used population balances to model aggregate breakage and concluded

that the shear breakage model based on that by Parker et al. (1972) (i.e. erosion) gave a

better fit to experimental data than a collision model based on Ham and Christman

(1969).

2.5.2 Effect of aggregate size on breakage

In Section 2.4.1 the concept of a maximum stable aggregate size (Parker et al. 1972)

was mentioned briefly (Equation 2-36). Parker et al. envisaged that aggregation would

continue until some point where the forces on the aggregate exceeded its strength, at

which point the aggregate would break. While the concept of a maximum stable

aggregate size is attractive in its mathematical simplicity, it has several limitations. It is

Page 51: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

50

not supported by experimental evidence, where particle size distributions invariably

form some sort of smooth bell-shaped curve, and not a distinct upper size limit. It is

also inconvenient for use in a population balance model, since the particles will tend to

pile up at the largest stable size.

An alternative approach takes the breakage rate to be a continuous function of aggregate

size. This can be rationalised since there will be a wide range of fluid fluctuations and

particle orientations through time, leading to a spectrum of breakage events. Spicer and

Pratsinis (1996) and Chung et al. (1998) both used the breakage kernel (2-38) based on

earlier work by Kapur (1972), Broadway (1978) and Peng and Williams (1994):

a

iY

i VAGS = 2-38 where: Si = Breakage rate of the ith sized particle (s-1) A = Breakage rate constant (cm-3ay) G = Shear rate (s-1) Vi = Volume of ith particle a = Constant (⅓) y = Fitted parameter

The constant of ⅓ is in agreement with the model result from Pandya and Spielman

(1982).

Various attempts have been made to directly measure the strength of aggregates. Smith

and Kitchener (1978) used MgSO4 or polymer to attach glass spheres to a silica surface.

The surface was either inverted or rotated in a centrifuge and the number of spheres

remaining counted. By calculating the force on the spheres by gravity (or centrifugal

acceleration) they concluded that the force of adhesion was roughly proportional to the

size of the particle.

More recently Yeung et al. (1997) trapped individual aggregates between two capillary

tubes and measured the force (around 100 nN) required to break the aggregates using a

calibrated cantilever beam. They reported to previously (Yeung et al. 1996) finding no

correlation between aggregate size and strength in the size range 15-48 μm. They did,

however, note that the maximum aggregate strength occurred when the aggregates were

formed at intermediate shear rates. They proposed this was due to an initial

Page 52: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

51

densification of the aggregate at low shear rates leading to stronger aggregates,

followed by weakening of the polymer links at higher shear rates as the polymer

became degraded. They suggest that the re-flocculated aggregates had a strength only

10 % that of the original aggregates.

2.5.3 Daughter particle distribution

Modelling aggregate breakage requires both a knowledge of the rate of breakage, and of

the size and number of daughter fragments (Ilievski 1996). Most workers suggest that

aggregates break into two or three similarly sized daughter aggregates (Fair & Gemmel

1964, Coulaloglou 1977, Glasgow & Lueche 1980, Pandya & Spielman 1982, Lu &

Spielman 1985, Hill 1996, Chung et al. 1998).

Breakage is frequently assumed to result in two equally sized fragments (Spicer &

Pratsinis 1996a, Flesch et al. 1999, Randolph & Larson 1988), simplifying the solution.

Alternatively, the fragments can be taken to be distributed according to some

distribution such as: normal (Valentas & Amundson 1966, Coulaloglou & Tavlarides

1977), log-normal (Valentas et al. 1966, Spicer & Pratsinis 1996a, Lu & Spielman

1985, Cheng & Redner 1988, Ducoste & Clark 1998), or binomial (Lick & Lick 1988,

Burban et al. 1989, Serra & Casamitjana 1998b).

The erosion of primary particles from the aggregate surface is occasionally suggested as

an alternative breakage mechanism. Experimental evidence using cameras has shown a

variety of breakage mechanisms (Lee & Brodkey 1987, Glasgow & Liu 1991)

(Figure 2-12) including breakage, fragmentation and surface erosion. Glasgow and Liu

(1991) reported seeing small particles eroded from the surface of a clay-polymer

aggregate, but note that the eroded particles were too large to be primary particles. This

is in qualitative agreement with Kim and Glasgow (1987), who suggest that turbulent

eddies of the size of the Kolmogoroff micro-scale have too little energy to erode

primary particles from the aggregate. Pandya and Spielman (1982) found an increase in

particles in a lower size range (although above the size of primary particles) and argued

that it was evidence of erosion.

Page 53: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

52

Table 2-6: Breakage equations reported in the literature Reference Kernel Comment/References Thomas 1964

25

23

29

p

fi D

S −− ∈ν⎟⎟⎠

⎞⎜⎜⎝

⎛ σ∝ Inertial

212

1

p

fi D

S ⎟⎠⎞

⎜⎝⎛∈ν

⎟⎟⎠

⎞⎜⎜⎝

⎛ σ∝ Viscal

σ = surface tension term

Parker et al. 1972

mB

i Gkdt

dN=

Inertial, m = 4,Viscal, m = 2

Coulaloglou & Tavlarides 1977

⎟⎟

⎜⎜

ν∈ρ

σ−∈ν= −

95

52

31

92

d

21

kexpk

dtdn

Droplet splitting

Tomi & Bagster 1978

( ) )d(dC)d(p 3

2

1 ∈ρ= η < d < L 21

2C)d(p ⎟⎠⎞

⎜⎝⎛ν∈

μ= d < η

32

23 dC)d(p ⎟

⎠⎞

⎜⎝⎛ν∈

ρ= d ∼ η

η = Kolmogoroff micro-scale L = Eulerian macro-scale

∅degaard 1979

p

Bi GkS φ= P = 2 or 4, Attributed to Parker, Kaufman & Jenkins

Glasgow & Lueche 1980

i2ii

ii NdfS

ψτ

= Viscal

( )i

ii

NdS

34

65

21

ψ∈μρ

≈ Inertial

τ = shear stress ψ = strength parameter fi = eddy freq. scale

Pelton 1981

( ) 2i

i r32kNS

θ−θ−θ=

θ = surface thickness of flocculant covering

Pandya & Spielman 1982

mii VS ∝

M fitted to ≈ 0.33 from experimental data, i.e. breakage ∝ diameter

Lu & Spielman 1985

ii KGVS =

Akers et al. 1987

kGdtdv

−=

Erosion kernel attributed to Spielman 1970

Leu & Ghoshi 1988

c3DG

dtdN r1

nv2

nv

π=

+

Shamalou & Titchener-Hooker 1993

v

2

7i CAS ⎟⎠⎞

⎜⎝⎛ν∈

= Inertial

v7i CAS ⎟⎠⎞

⎜⎝⎛ν∈

= Viscal

Cv = aggregate volume

Page 54: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

53

Table 2-6: Continued

Huang & Helums 1993

m

ii 1xkS ⎟⎟

⎞⎜⎜⎝

⎛φ−

= m = 0.1 φ = void fract.

Spicer & Pratsinis 1996

y

ii GAVS 31

= y = 1.6 ± 0.18 G = 25-150 s-1, (Oles 1992)

Berlin et al. 1997

( )

⎟⎠⎞

⎜⎝⎛ −= − N

oN1b e1NK

dtdN

Serra & Casamitjana 1998

31

ib

Bi VGkS =

From Spicer & Pratsinis (1996)

Lu et al. 1998

32

31

21

iSη

∈δ−=

δ = double layer η = K.M

Chung et al. 1998

ya

ii GAVS = From Chen (1990), Spicer & Pratsinis (1996)

Ducoste & Clark 1998

⎟⎟⎟

⎜⎜⎜

ρ=

21

21

21

dNDN

CNerfcNCS

p

3Qii

Stirred tank reactor

Kobayashi et al. 1999

21-

4max,f fNcGCd ⎟⎠⎞

⎜⎝⎛ μ= Viscal

83

32 -

5max,f fNcCd ⎟

⎜⎜

⎛ ∈ρ= Inertial

Kramer & Clark 1999

( ) fD1

)k(ra2kNS 1b'

max'fsi μ=

Reich & Vold, Lu & Spielman, Thomas, Ray & Hogg

Flesch et al. 1999

y

i G'AS = Kusters (1991), Spicer & Pratsinis (1996)

Young et al. 2000

L'

Bi GkS = L = 2 (<KM) or 4 (>KM)

Page 55: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

54

2.6 Population balance models 2.6.1 Smoluchowski’s equation, aggregate breakage

In earlier sections, kinetic equations (kernels) for aggregation and breakage were

described. In order to simulate the overall aggregation process, some way is needed to

keep track of the number and size of the aggregates through time. In 1917

Smoluchowski proposed the first population balance model, describing aggregation

only:

∑∑∞

=

=+=

β−β=1i

kiik

1k

kji,1ijiij2

1k NNNNdt

dN 2-39

where: Nk = Number of k sized particles βij = Rate of collision between i and j sized particles (aggregation kernel) t = Time (s) The first term on the right of Equation 2-39 describes the increase in number of k-sized

particles through the aggregation of i and jth particles. The constant of ½ is included to

avoid double-counting. The second term describes the loss of k-sized particles through

aggregation of a k-sized particle with another particle.

To model aggregate breakage another two terms are needed (Argaman & Kaufman

1968, Randolph & Larson 1988, Lick & Lick 1988, Spicer & Pratsinis 1996a, Serra &

Casamitjana 1998b, Kramer & Clark 1999):

∑∑∑∞

+=

=

=+=

Γ+−αβ−αβ=1kl

lllkkk1i

kiik

1k

kji,1ijiij2

1k NSNSNNNNdt

dN 2-40

where: Sk = Breakage rate (kernel) of kth sized particles

Γlk = Breakage distribution function (number of k size particles produced from the breakage of a l sized particle) αij = Capture efficiency

In this case the third term on the right describes the loss of k-sized particles due to

breakage. The final term describes the gain in k sized particles through the breakage of

an i sized particle to give a k-sized daughter fragment.

Page 56: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

55

A population balance is a system of differential equations, with one equation for each

size particle. However, a particle size distribution is effectively a continuous function

(many particles), and covers a wide range of sizes (typically several orders of

magnitude). Various analytical solutions have been formulated, but usually only for

breakage or aggregation separately (Patil and Andrews 1998), and only for the simplest

aggregation or breakage kernels e.g. size-independent (Hounslow et al. 1988). Patil and

Andrews (1998) describe an analytical solution for simultaneous aggregation/breakage,

but the solution requires the assumption that the total particle number does not change

with time.

2.6.2 Discrete models, numerical solution

Analytical solutions are generally unavailable for physically realistic kernels, but

simplified kernels (e.g. size-independent) are used to check the accuracy of numerical

solutions (Hounslow et al. 1988, Litster et al. 1995, Smit et al. 1994). For a numerical

solution, the continuous spread of particle sizes is sectioned (discretised), in much the

same way that particle sizing instruments give histograms where each bar covers a

range of sizes.

Discretisation allows the population balance to be written as a closed set of ordinary

differential equations, which can be readily solved numerically. Initially workers

writing population balances had to write their own numerical routines, typically

Runge-Kutta methods coded in Fortran (Ilievski 1996), but recently commercial

packages have become available e.g. SPEEDUP (Johnson & Cresswell 1996), NIMBUS

(Ilievski 1996), MATLAB (Biggs & Lant 2002), gPROMS (this work).

In 1964 Fair and Gemmel coded one of the earliest discretised population balances

based on Equation 2-38 by Smoluchowski (1917). The size range was discretised into

20 equally spaced bins by volume. The balance was solved using an iterative technique

in Fortran. Aggregate sizes were limited either by stopping aggregation that would

cause oversize particles, or by breakage of oversized particles.

Batterham et al. (1981) wrote a population balance for pelletization kinetics

(aggregation only) in balling circuits, which Hounslow et al. (1988) simplified to:

Page 57: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

56

( )∑ ∑−

=

−−−−−−−−−

β−β++

β+β+β=

2i

1j j

2jijjij,i

ij

21i1i,1ii1ii,1i4

31i2i1i,2i8

3i

NNN21

NNNNNdt

dn

2-41

Hounslow et al. (1988) noted that the model correctly predicts particle volume (third

moment) but does not accurately predict total particle number (zeroth moment). Kumar

and Ramkrishna (1996) pointed out that Batterham’s model double counts equal sized

particle aggregation, although the errors are somewhat self-cancelling. Despite the

criticisms, Batterham et al. proposed an elegant method of discretisation, following a

simple numerical progression where volume (or mass) doubles for each size range:

2V

Vi

1i

=+

2-42

or: 3

i

1i

2L

L=

+

2-43

where: Vi = Volume of the ith channel Li = Length of the ith channel This method of discretisation has several advantages. The spacing is finer for smaller

particles, giving additional accuracy for the large number of small particles which

would be almost completely lost if constant width discretisation was used (Hounslow

et al. 1988). The discretisation also covers a wide range of particle sizes with

comparatively few channels, reducing the computation time. This discretisation was

also previously consistent with the output from a Coulter counter (Hounslow et al.

1988) and Reynard (R10) series sieves (Allen 1990, p. 192), simplifying comparisons

between modelled and experimental data. Batterham et al.’s discretisation method has

been used by other workers, notably Hounslow et al. (1988), and Spicer and Pratsinis

(1996a) (Equations 2-44 & 2-45).

A variety of population balance models have subsequently been described (Pandya and

Spielman 1982, Lu & Spielman 1985, Hill 1996, Randolph & Larson 1988, Hounslow

et al. 1988, Marchal et al. 1988, Lick & Lick 1988, Serra & Casamitjana 1998a, Litster

et al. 1995, Wynn 1996, Chung et al. 1998, Nickmanis & Hounslow 1998).

Page 58: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

57

Hounslow et al. (1988) used the discretisation described by Batterham et al. (1981) and

derived (see Appendix):

∑∑∑∞

=

=−−−

=−−

+− β−β−β+β= −

ijjj,ii

1i

1jjj,ii

21i1i,1i

2i

1j21

j1ij,1i1iji NNN2NNNN2

dtdN ij 2-44

Hounslow et al.’s model describes aggregation only. The first two terms on the right

describe the aggregation of smaller particles into the ith size range, and the second two

terms describe the loss of particles in the ith size range as they aggregate with other

particles (see Appendix).

Hounslow et al.’s balance has been widely used and referenced (Atteia 1998, Cresswell

et al. 1994, Ilievski & White 1994, Ilievski & White 1995, Kumar and Ramkrishna

1997a, b, c, Patil & Andrews 1997, Simons 1996, Spicer & Pratsinis 1996a, Spicer et al.

1996, Spicer 1997, Flesch et al. 1999) primarily because it accurately predicts both

particle number (zeroth moment) and volume (third moment) (Hounslow 1988, Kumar

& Ramkrishna 1997) and is simple to solve, with a minimum of terms containing

summations rather than integrals.

However, Hounslow et al.’s population balance has a relatively coarse discretisation,

and gives slight errors in the first and second moments (length and area weightings).

Kostoglou and Karabelas (1994) compared population balances by Batterham 1981,

Hounslow et al. 1988, Marchal et al. 1988, and Gelbard et al. 1990 and found that

Hounslow et al.’s balance gave the best performance. Similar results are shown by

Litster et al. (1995), who showed that for particles coalesced from 50 primary particles

the errors in the first (length) and second (area) moments were in the order of 2 %, with

no error in the zeroth or third moments.

Hounslow et al.’s population balance was written primarily to describe crystallisation

systems, and the paper also describes crystal growth kinetics. For flocculation

applications growth terms are not needed, but breakage terms are. Spicer and Pratsinis

(1996a) added two terms to Hounslow et al.’s balance:

Page 59: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

58

∑∑∑∞

=

=

=−−−

=−−

+−

Γ+−

β−β−β+β= −

ijjjj,iii

ijjj,ii

1i

1jjj,ii

21i1i,1i

2i

1j21

j1ij,1i1iji

NSNS

NNN2NNNN2dt

dN ij

2-45

These terms describe the loss of particles from the ith channel by breakage (second last

term) and increase in the number of ith particles due to the breakage of larger aggregates

(final term). In this case Si is the breakage kernel and Γi,j is the breakage distribution

(see Section 2.4.3). The simplicity of the discretisation used by Hounslow et al. makes

the coding of Equation 2-45 relatively straightforward.

Solving population balances numerically allows physically realistic kernels to be solved

relatively simply. However, solving discretised systems of differential equations

numerically typically leads to numerical errors. Using finer particle size discretisation

and decreasing the step size is likely to help, but other methods are required to check

the accuracy of the results. With aggregation, the mass of the solid is constant through

time (as opposed to a crystallisation with surface deposition), and the population

balance should conserve mass (third moment).

However, a model may correctly conserve mass but not correctly predict particle

number (e.g. Batterham et al. 1981). Particle sizing is frequently performed using

Coulter counters, hence the correct prediction of the total particle number (zeroth

moment) is frequently described in the population balancing literature (Kumar and

Ramkrishna 1996, Hounslow et al. 1988, Litster 1995). Typically the moments from

zero through to the third (particle number, length weighted, area weighted and volume

or mass weighted) are compared to analytical solutions, usually for a very simple

formulation of the aggregation kernel, e.g. size independent (Hounslow et al. 1988,

Litster et al. 1995).

Higher moments are also sometimes checked, usually the sixth. The physical

interpretation of such a high moment is obscure, but it is used to check the mathematical

stability and accuracy of the model. Weighting the system so heavily with size tends to

reveal the errors in the few larger particles in their wider channels. Smit et al. (1995)

Page 60: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

59

used the sixth moment as an indication of mathematical gellation, where the population

balance rapidly “runs away” producing a very few super-sized aggregates. In practice,

the use of a breakage kernel effectively limits the aggregate size.

Population balance models are usually compared to experimental data, with the

experimental data used to estimate unknown parameters of the model. This is usually

performed with a sum-of-squares minimisation routine (Ilievski et al. 1993, Ilievski &

White 1994). The data is also used as evidence for various formulations of aggregation

and breakage kernels; the “inverse problem” (Muralidar & Ramkrishna 1986, Ilievski &

White 1994). However, caution is required in interpreting the fit between modelled and

experimental data, because even a physically unrealistic model may give a reasonable

agreement with enough fitted parameters, or if there is limited experimental data.

In conclusion, the architecture of population balances is well established, with several

successful models already being available, in particular the models by Hounslow et al.

1988, and Spicer and Pratsinis 1996a. Although other, slightly more accurate,

population balance models are also available, they are also considerably more complex

reducing the incentive to use them, especially as a sub-model within CFD simulations.

In addition, the level of inaccuracy in the numerical population balance models is

probably relatively minor in comparison to the inaccuracy in the experimental data

(Ilievski & White 1995). The availability of existing population balance models, and

commercial simulation packages capable of solving the differential equations, will

allow future research efforts to focus on the determination of appropriate aggregation

and breakage kernels for specific systems.

2.7 Solid-liquid separation by gravity settling The settling of solid particles in a viscous fluid can occur in several different regimes,

depending on the solid volume fraction and the aggregate size and density (Figure

2-13). At very low solid fractions, individual particles are well separated in the fluid

and settle freely with little, or no, interaction with each other. The settling velocity of

solid spherical particles in creeping flow was first described by Stokes (1851), but

Stokes’ equation (Equation 1-3) has since been extended to suit other particle

geometries and higher settling velocities.

Page 61: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

60

Figure 2-13: Effect of solid fraction and degree of flocculation on settling behaviour (adapted from Perry and Green 1997).

At higher solid fractions (Figure 2-13) individual particles begin to interact with each

other (Bhatty et al. 1982, Perry & Green 1997), until a critical point when the particles

are constrained to settle together as a loosely connected mass (Govier & Aziz 1972,

Tong et al. 1998, Perry & Green 1997, p. 18-60). This mass subsidence is referred to as

hindered or zone settling, (or even more confusingly, sometimes free-settling, Fitch

1975A, Pearse 1977) and is characterised by a distinct mudline separating clear

supernatant liquor from the settling mass of solids (Tong et al. 1998, Pearse 1977,

p. 15). Although the settling velocity is slower than the average free-settling velocity of

the same particles (Wesely 1985, Dahlstrom & Fitch 1985, Pearse 1977),

hindered-settling has the advantage that it captures fine particles that would otherwise

report to the overflow (Wesely 1985). In this region, solid settling is still restrained

only by hydrodynamic forces, and the settling velocity is a function of solid fraction and

aggregation state (Pearse 1977).

Eventually, the mass of settling solid will contact the base of the vessel or thickener and

sediment compaction begins. The rate of solid settling is now restrained by both

hydrodynamic and mechanical forces. Some sediment compaction is usually necessary

to achieve the required underflow concentration. Compaction is often encouraged by

Solid

frac

tion

Highly aggregated

particles

High concentration

Low concentration

Degree of particle coherence

Hindered settling regime

Particulate settling regime

Compression regime

Totally discrete particles

Page 62: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

61

allowing the bed to build up (typically to ∼ 1 m, Galvin & Waters 1987, Pearse 1977,

Perry & Green 1997) to increase the overburden weight of the bed. The action of

raking the solids towards the underflow pump at the base of the thickener also helps de-

water the sediment by disrupting the bed structure, and by forming channels that

increase the drainage rate of the displaced fluid (Perry & Green 1997, Farrow et al.

2000).

2.7.1 Particulate settling (low solid fraction)

Solid spherical particles

The terminal settling velocity of individual particles is determined by the various forces

acting on the particle, as shown by Figure 2-14 (Stokes 1851, Clift et al. 1978, Seville

et al. 1997, Happel & Brenner 1973).

Figure 2-14: Forces acting on a free-settling particle (adapted from Wesely 1985).

where: Fg = force of gravity (N, or kg m s-2) Fd = force of drag (N, or kg m s-2) Fb = force of buoyancy (N, or kg m s-2) The general equation for the settling velocity of solid spheres is:

dl

ls2

C3)(dg4

ρ−ρ= 2-46

where: U = Settling velocity (m s-1) d = Particle diameter (m) g = Gravity (9.8 m s-2)

ρ = Density (s = solid, l = liquid) (kg m-3) Cd = Coefficient of drag

Fb

Fg

Fd

Particle

Page 63: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

62

Stokes solved the Navier-Stokes equation and showed that for creeping flows past rigid

spheres (see Clift et al. 1978):

Re24Cd = 2-47

Where Re, Reynolds number for spheres is:

μρ

= lUdRe 2-48

where: μ = Fluid viscosity (kg m s-2) Combining Equations 2-45, 2-46 and 2-47 gives Stokes’ law:

μρ−ρ

=18

)(gdU ls

2

2-49

Stokes’ law rapidly breaks down (Concha & Almendra 1979) at higher Reynolds

numbers as fluid inertial effects become significant in comparison to the viscous

resistance. Alternative drag coefficients have been proposed for higher Reynolds

numbers, usually as a function of the Reynolds number (Figure 2-15). The settling

velocity is overestimated by Stokes’ law by 2 % at Re = 0.1, 4 % at Re = 0.2 and 20 %

at Re = 1 (Seville et al. 1997, Rushton et al. 1996). Figure 2-16 shows the effect of

particle diameter and settling velocity on the Reynolds number, indicating regions

(Re < 1) where Stokes’ law can be used to describe the settling velocity.

Page 64: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

63

ψ = 0.125ψ = 0.220ψ = 0.600ψ = 0.806

ψ = 1.000

10000

100060040020010060402010

6421

0.60.40.20.1

0.10.010.001 1 10 100 1000 10000 105 106

Reynolds number Dpvρ=NRe

μ

Dra

g co

effic

ien t

2F

d g c

=C

dv2ρS

Figure 2-15: Effect of Reynolds number and sphericity (ψ) on the drag coefficient (adapted from Wesely 1985).

0

10

20

30

40

50

60

70

80

90

100

0 200 400 600 800 1000 1200 1400 1600 1800 2000

Particle size (µm)

Sett

ling

rate

(m h

r-1)

Re = 10

Re = 3

Re = 1

Re = 0.1

Re = 0.3

Re = 0.01

Figure 2-16: Iso-plot of particle Reynolds number as a function of size and settling rate, calculated from Stokes’ law for the free-settling rate of spherical particles. Stokes’ law is only accurate for low Reynolds numbers (see text) and hence is only appropriate for particles bounded by the lower lines in the figure. (constructed from Equation 1-3, 2-49).

Page 65: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

64

A variety of alternative drag coefficients have been published (Table 2-7), generally in

the (dimensionless) form of Equation 2-50 for particles settling at higher Reynolds

numbers when inertial effects become significant (Pearse 1977).

nd RekC = 2-50

where: Cd = Coefficient of drag (dimensionless) Re = Reynolds number for spheres (dimensionless) k = Constant n = Exponent Very high Reynolds number settling (1000 < Re < 350 000) is described by Newton’s

(1687) equation (Equation 2-51), where the drag coefficient is a constant of 0.445,

independent of the viscosity and settling velocity, but highly shape dependent

(Figure 2-15) (Clift et al. 1978, Perry & Green 1997):

( )lN

ls2N C3

dg4U

ρρ−ρ

= 2-51

where: UN = Settling velocity given by Newton’s equation (m s-1)

ρ = Density (s = solid, l = liquid) (kg m-3) g = Acceleration due to gravity (9.8 m s-2) CN = Drag coefficient given by: 5.31 - 4.88ψ ψ = Sphericity, ratio of surface area of equivalent sphere to actual surface area (see also: Amirtharajah et al. 1991)

Gregory (1997) comments that in most cases porous aggregates settle sufficiently

slowly for Stokes’ law to apply.

Page 66: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

65

Table 2-7: Coefficient of drag (Cd) as a function of Reynolds number (Re)

Reference Cd Re Source Stokes 1851 24/Re Re → 0 1 Oseen 1927 (24/Re)(1+3Re/16) Re << 1 1 Goldstien 1929

(24/Re)(1 +3Re/16 –19Re2/1280 + 71Re3/2480 – 30179Re4/34406400 + 122519Re5/56742400)

Re << 1

1

Clift et al. 1978

3/16 + 24/Re

Re < 0.01

3

Clift et al. 1978

(24/Re)(1 + 0.1315Re(0.82-0.05logRe))

0.01 ≤ Re ≤ 20

1, 3

Pruppacher & Steinberger 1968, Beard & Pruppacher 1969

(24/Re)(1 + 0.102Re0.955)

0.02 ≤ Re ≤ 3

1

Kurten et al.1966

0.28 + 6/Re0.5 + 21/Re

0.1 < Re < 4000

3

Gilbert et al. 1955 0.48 + 28Re-0.85

0.2 < Re < 2000 3

Landmuir & Blodgett 1948 (24/Re)(1 + 0.197Re0.63 + 2.6x10-4Re1.38)

1 < Re < 100 3

Lin & Lee 1973 (24/Re)(1 + 0.2207Re0.5 + 0.0125Re)

1 ≤ Re ≤ 1000 1

Allen 1900 30Re-0.625

1 < Re < 1000 3

Allen 1900 10Re-0.5 2 < Re < 500

3

Schiller & Naumann 1933

(24/Re)(1 + 0.15Re0.687) 2 ≤ Re ≤ 800 1, 3

Pruppacher & Steinberger 1968, Beard & Pruppacher 1969

(24/Re)(1 + 0.115Re0.802) 3 ≤ Re ≤ 20 1

Proudman & Pearson 1957

(24/Re)(1 + 3Re/16 + 9Re2ln(Re/2)/160) Re ≤ 1.3 1

Kurten et al.1966

2 + 24/Re Re < 10 3

Clift et al. 1978 (24/Re)(1+ 0.1935Re0.6305) 20 ≤ Re ≤ 260

1, 3

Pruppacher & Steinberger 1968, Beard & Pruppacher 1969

(24/Re)(1 + 0.189Re0.632)

0 ≤ Re ≤ 4000

1

Kurten et al.1966

1 + 24/Re Re < 100 3

Lapple 1951

(24/Re)(1 + 0.125Re0.72) Re < 1000 3

Abraham 1970

0.2924(1 + 9.06Re-0.5)2 Re < 6000 3

Matsumo & Mori 1975

K/NRe0.82 2, Fractal

Tambo & Watnabe 1979, Moudgil & Vasudevan 1989

K/NRe

2, Fractal

Dirican 1981, Klimpel & Hogg 1986, Hogg et al. 1987

K(1+K’/NRe

0.5)2

2, Fractal

Newton 1687 0.445 (accurate to 13%) 1E+3 ≤ Re ≤ 3.5E+5 4 1: Nguyen-Van et al. 1994 2: Moudgil and Vasudevan 1989 3: Clift et al. 1978 4: Perry and Green 1997

Page 67: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

66

Other shapes

Stokes’ equation was written for solid spherical particles, however aggregates formed in

thickeners have highly irregular shapes. Various workers (Pettyjohn and Christiansen

1948, Happel and Brenner 1973, Kousaka et al. 1981, Seville et al. 1997) have studied

the effect of shape on free-settling. Pettyjohn and Christiansen followed the settling

behaviour of a wide range of regular metal shapes in glucose/water solution and

calculated a wide range of correction factors (k) under various conditions. The

correction factors were typically in the range 0.7-1.0, i.e. the particles settled slightly

slower than predicted by Stokes’ equation (Heiss & Coull 1952). Happel and Brenner

(1973) and Seville et al. (1997) considered shapes of various aspect ratio, settling with

various orientations (Figure 2-17). Again, both showed that shapes with aspect ratios

typical of aggregates settle slightly slower than predicted by Stokes’ equation.

0.1 0.5 5 10

0.50

0.60

0.70

0.80

0.90

1.00

Constant circularity ofprojected area

Variable circularity ofprojected area

- Cylinders- Rectangular parallelograms

- Cylinders- Rectangular parallelograms

Spheroidedgewise

Spheroidflatwise

Height-diameter ratio, h/d

Settl

ing

f act

or, K

1.0

Figure 2-17: Effect of particle shape on free-settling velocity in creeping flows (Heiss and Coull 1952, adapted from Happel and Brenner 1973).

At high Reynolds numbers the particle shape has a larger impact on the particle drag

(Figure 2-15).

Aggregate Porosity

Aggregate porosity has two counteractive effects on the settling velocity of an

aggregate. Firstly, it may allow fluid to pass through the structure, reducing the drag.

Secondly, it dramatically reduces the aggregate density.

Page 68: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

67

Veerapaneni and Wiesner (1996) compared various models that assume constant

internal permeability that suggest that the flow through the aggregate structure only

affects the settling velocity for fractal dimensions lower than approximately 2.0, less

than found in practice (Table 2-3). Gregory (1997) used Brinkman’s (1948)

permeability model to model the effect of fractal dimension under typical industrial

conditions and also showed that flow through the structure had little impact with a

fractal dimension of over 2.

The major effect of porosity on particle settling is from the reduction in the effective

density at lower fractal dimensions (Figure 2-18), given by (Gonzales & Hill 1998,

Atteia 1998, Ellis & Glasgow 1999, Manning & Dyer 1999):

( ) ( )3D

p

agglslagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ−ρ=ρ−ρ 2-52

where: d = Diameter (agg = aggregate, p = primary particle)(m) Df = Mass-length fractal dimension

ρ = Density (agg = aggregate, l = liquid, s = solid) particle (kg m-3)

0

500

1000

1500

2000

2500

3000

0 10 20 30 40 50 60Relative aggregate size (dagg/dp)

Agg

rega

te d

ensi

ty (k

g m

-3)

Figure 2-18: Effect of aggregate porosity on density. Assuming: ρs = 2710 kg m-3, ρl = 1000 kg m-3, Df = 2.4.

Numerous experimental studies have related the particle settling velocity to the particle

size under various conditions, flocculant/coagulant types, particle material etc. These

studies invariably show significant experimental scatter, but usually produce a straight

line relationship between log (settling velocity) vs. log (particle diameter) (Sternberg

Page 69: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

68

et al. 1996, Matsumoto & Mori 1975, Kusuda et al. 1981, Allain et al. 1996, Adachi &

Kamiko 1994, Huang 1994, Adachi & Tanaka 1997, Dyer & Manning 1999), although

some workers have plotted settling velocity vs. size (Farrow & Warren 1989, 1993).

2.7.2 Hindered settling (intermediate solid fraction)

Effect of solid fraction on hindered settling

The transition from free-settling to hindered (or zone) settling is gradual, and generally

poorly defined (Bhatty et al. 1982, Perry & Green 1997). However, hindered settling is

characterised experimentally by a distinct settling line (mudline), leaving a clear

supernatant above (Fitch 1975a, 1987, Tong et al. 1998). This situation may be brought

about by two separate conditions (Pearse 1977), firstly, if the bulk settling velocity of

the suspension is slower than the free-settling velocity of the slowest (probably

smallest) particles. In this situation the small particles will catch up to the mudline,

leaving a clear supernatant. The onset of hindered-settling is a function of the particle

size distribution and the bulk-settling velocity at a given concentration. In the artificial

situation where all the particles are the same size and shape, a distinct settling line will

appear regardless of the solid fraction.

The second condition that will cause a clear supernatant is when differential settling

results in aggregation (Pearse 1977). In this situation large, fast-settling aggregates

collide with slower settling particles and sweep them from solution (Fitch 1975a). The

onset of a distinct mudline in this situation will be a function of the capture efficiency

(Bhatty et al. 1982), the collision radius, the fractal dimension and the solid

concentration.

The hindered-settling velocity decreases as a function of solid fraction (Table 2-8), due

to the increase in the hydrodynamic force (Govier & Aziz 1972, Perry & Green 1997),

which increases as the suspension behaves increasingly like a permeable network.

Settling is also slowed by an increase in the rise velocity of displaced fluid (Perry &

Green 1997). Flocculation has the effect of creating channels that ease the fluid flow

through the solid network, increasing the settling velocity (Govier & Aziz 1972).

Page 70: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

69

Various empirical and theoretical relationships between settling velocity and solid

fraction have been proposed (Table 2-8), with the relationship by Richardson and Zaki

(1955) being the most widely used:

( )n

o 1UU φ−= 2-52 where: φ = Solid volume fraction (m3m-3)

n = Exponent, usually taken as 4.65 (Table 2-9) U = Suspension settling velocity (m s-1) Uo = Free settling velocity of same size particle (m s-1)

Richardson and Zaki’s (1955) equation has been fitted successfully to a wide range of

suspensions, including flocculated suspensions (Table 2-9), with the exponent typically

found to be 4.65 in Stoke’s regime (Re < 0.2).

Page 71: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

70

Table 2-8: Effect of solid fraction on the hindered-settling velocity Reference Settling velocity (U, m s-1) Particle

size (μm) Solid fraction (φ)

Comments

Burgers 1941, 1942

φ+=

88.61UU o

0 < φ < 0.1 1 , 2

Steinour 1944a,b,c )(UU 2

s εφε= 13.5 & 1740

3, widely used

Uchida 1952

31

1.21

UU o

φ+=

0 < φ < 0.1 2

McNown & Lin 1952

31

6.11

UU o

φ+=

0 < φ < 0.1 2

Richardson & Zaki 1955, Maude and Whitmore 1958

( )no 1UU φ−= 100-1000

0 < φ < 0.35

3, widely used

Michaels & Bolger 1962 ( )nfloc k1UU φ−=

0.2-2 3

Thomas 1963

φ−= 9.5oeUU

0 < φ < 1 1, Empirical

Famularo & Happel 1965

31

1

UU o

γφ+= γ = 1.3 ± 0.24

0 < φ < 0.1 1, 2, 3

Smith 1965, 1966 )(3FU)(2FU)(1FU iiiiii i∑ φ+φ+φφ= φ

2900-6000

0.008 < φ < 0.250

3

Dollimore & McBride 1968 ( )noUU φ=

widely used

Batchelor 1972 )66.61(UU o φ−=φ

3 Govier & Aziz 1972

( )φφ−

=101U

U3

o

0 < φ < 1 1

Lockett & Al-Habbooby 1973 s,ls1o1 )1(UU φ−φ−= 600-2000 0.05 < φ < 0.5

Barnea & Mizrahi 1973 [ ])1(3/5xpε)1(

)1(UU31

o

φ−φφ+

φ−=φ

5-1740

3

Garside & Al-Dibouni 1977

( ) 2.0

R

R Re06.0U)(B

AU +∈=−∈

∈−

170-3100

0.05 < φ < 0.5

3

Ishii & Zuber 1979 Re),('FUU o φ=φ 360-6000 0.05 < φ < 0.5 3

Masliyah 1979 μ

∈ρ−ρ∈=

18)(F)(gxU ms

2i

70-400

0.1 < φ < 0.5

3

Concha & Alemendra 1979 )(f1)(fx0921.01

x52.20U 1

2

2dd

* 21

32

φ⎥⎦⎤

⎢⎣⎡ −⎟

⎠⎞⎜

⎝⎛ φ+=

3

Reed & Anderson 1980 )0.41(

)88.11(UU o

φ+φ−

3

Bhatty et al. 1982 [ ])11/(1oUU ∈−∈∈= 0.05 < φ < 0.6 3

Buscall et al. 1982 P'k

P1UU o ⎟

⎠⎞

⎜⎝⎛ φ−=φ

3

0.01 < φ < 0.5

3

Glendinning & Russel 1982 )55.61(UU o φ−=φ φ < 0.2 3

Patwrdhan & Chi Tien 1985 ⎟⎟

⎞⎜⎜⎝

⎛ρ−ρρ−ρ

=∈ −φ

fsi

msi)2n(i

oUU

i

0.15 < φ < 0.5 3

Patwrdhan & Chi Tien 1985 φφ μμ∈= /UU o 2000-5000 0.1 < φ < 0.5 3

Mandersloot et al. 1987

D/x8.1770.4so

*)K1(UU +φ+=

0.05 < φ < 0.8 3

Daves & Gecol 1994 ( ) ( )( )∑ φ−+φ−= jiiij

Ss SS11UU ij

U = settling velocity (m s-1), Us = settling velocity given by Stokes’ law (Equation 1-3, 2-46) (m s-1), Uo = settling velocity at infinite dilution (m s-1) (may = Us), φ = solid fraction (1-ε) (dimensionless), ε = liquid volume fraction (1-φ) (dimensionless), Sij = function of polydispersity. 1: Govier and Aziz (1972). 2: Famularo and Happel (1965). 3: Williams and Amarasinghe (1989).

Page 72: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

71

Table 2-9: Exponents reported for Richardson & Zaki’s Equation Reference Exponent Comment Bhatty et al. 1982 Not stated

Davies & Gecol 1994 5.1 Attributed to Garside & Al-Dibouni 1977

Dollimore & Horridge 1971 Not stated

Farrow et al. 1985 Not stated

Fitch 1975B 4.65 References Michaels & Bolger 1962, can use for

aggregates provided use effective solid fraction

Font 1991 4.65

Hogg & Bunnaul 1992 Not stated Suggests maximum effective solid volume fraction approx 0.5-0.6

Kanungo & De 1970 4.7

Maude & Whitmore 1958 ∼ 5 Large table of exponents for different systems, 4.15-9.35

Michaels & Bolger 1962 4.65 Aggregates

Perry & Green 1997 4.65

2.33 Stokes’ regime Newton’s regime

Turian et al. 1997

4.8 5.4 5.8

Spheres Cube Brick shape

Di Felice et al. 1993

4.65 4.4Re-0.03

4.4Re-0.1 2.4

Re < 0.2 0.2 < Re < 1 Re = particle Reynolds number 1 < Re < 500 Re > 500

Rushton et al. 1996

4.65 4.35Re-0.03

4.44Re-0.1

4.45Re-0.1 2.39

Re < 0.2 0.2 < Re < 1 1 < Re < 200 200 < Re < 500 Re > 500

Estimating the continuous settling flux from batch data

Thickeners usually operate under approximately steady-state conditions, with

continuous feed, underflow, and overflow rates. However, experimental settling data is

normally collected from a batch-settling test in a measuring cylinder, or similar settling

column. Various methods are available for estimating the continuous settling flux (and

hence required thickener area) from batch data.

Page 73: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

72

Most methods for sizing thickeners assume that some critical solid fraction will limit

the settling flux. The settling flux is defined (Pearse 1977, Rushton et al. 1996) as the

settling velocity multiplied by the solid fraction:

φ=ψ U 2-54 where: ψ = Settling flux (m s-1) U = Settling velocity (m s-1) φ = Solid volume fraction (m3m-3) Multiplying by the solid density (ρs) gives the mass settling flux (kg s-1m-2, or t m-2h-1).

Mishler (1912) pioneered the use of laboratory tests to predict settler performance by

assuming that the initial feed settling velocity was balanced by the overflow fluid rise

velocity. Coe and Clevenger (1916) suggested measuring the initial settling velocity of

various suspensions, ranging from the feed concentration to the desired underflow

concentration. The settling flux relative to the underflow withdrawal flux is then

calculated from:

u

11U

φ−

φ

=ψ 2-55

where: U = Settling velocity (m s-1) φ = Solid fraction (m3m-3) φu = Underflow solid fraction (m3m-3)

Coe and Clevenger assumed that the settling velocity was a unique function of the solid

fraction, which may not be true for flocculated suspensions if the degree of flocculation

is also dependent on the solid fraction (Williams & Simons 1992, Svarovsky 2000).

Equation 2-55 is then used to calculate the settling flux from feed to underflow dilution,

with the concentration giving the minimum flux rate used to calculate the required

thickener area from:

l

ul

U

11FA

⎟⎟⎠

⎞⎜⎜⎝

⎛φ

−φ

= 2-56

Page 74: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

73

where: A = Thickener area (m2) F = Solid feed rate (m3s-1) Ul = Settling velocity of limiting concentration (m s-1) φl = Limiting solid fraction (m3m-3) φu = Underflow solid fraction (m3m-3)

The Coe and Clevenger method for sizing thickeners was the preferred method for

nearly 40 years (Pearse 1977) until Kynch (1952) produced his analysis of the batch

settling curve, allowing the calculation of the flux curve from a single settling test. In

this method a large (e.g. 1 litre) measuring cylinder is filled with a suspension of the

expected feed concentration and the fall of the mudline is measured through time.

Kynch proposed that layers of higher concentration would propagate upwards from the

base of the container (Figure 2-19) at constant rates until they intersected the mudline,

at which point the mudline settling velocity would drop to that given by the new

interface concentration. As with Coe and Clevenger, Kynch also assumed that the

settling velocity is a unique function the solid fraction. However, since Kynch’s

method starts with a suspension at feed concentration, it is less influenced by the effect

of solid fraction on the aggregation process.

Time (min.)

Sedi

men

t hei

ght (

m)

Constant settling region (concentration constant at φf)

Kynch isoconcentraion lines (concentration increasing at interface)

H1

Ho

40 % w/v45 % w/v50 % w/v55 % w/v60 % w/v

Figure 2-19: Kynch analysis of the batch settling curve, where layers of higher concentration propagate upward at constant rates until they intersect the interface where they cause a reduction in the settling velocity (see Figure 2-20). Line drawn from H1 is used in Equation 2-58.

Page 75: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

74

The rate of upward propagation of the isoconcentration lines is given by (see Figures

2-19 & 2-20):

φψ

−=dd

dtdHk 2-57

where: HK = Height of Kynch isoconcentration line ψ = Settling flux (m h-1), see Figure 2-20

The batch experimental test will not give settling flux data covering the entire solid

fraction range shown in Figure 2-20. Initially the solid fraction just below the mudline

will be the initial feed concentration, provided the suspension is initially mixed to

homogeneity. This concentration will initially be maintained as the mudline falls.

Eventually the mudline will meet higher concentration Kynch layers propagating

upwards from the base of the vessel (Figure 2-19). At this point the solid fraction at the

mudline will increase rapidly and the settling velocity of the mudline will begin to

decrease.

0 10 20 30 40 50 60 70 80 90 100

Solid concentration (% w/v)

Sett

ling

flux

(m h

r-1)

φf

Kynch isoconcentration lines

40 % w/v45 % w/v50 % w/v55 % w/v60 % w/v

Figure 2-20: Kynch analysis of the settling flux curve, where layers of higher concentration propagate upward at rates given by lines drawn at a tangent to the flux curve. The lowest concentration Kynch line is found by drawing a line at a tangent to the flux curve such that it intersects the flux curve at the feed concentration (φf) (see Figure 2-19).

The solid fraction of the Kynch layers is likely to be considerably greater than the feed

concentration. The first Kynch layer intersected by the mudline is the Kynch layer with

Page 76: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

75

the fastest rise velocity. This is given by drawing a tangent to flux curve such that it

also intersects the flux line at the feed concentration (e.g. φ = 40 % w/v in Figure 2-20).

Faster rising Kynch layers are not possible because insufficient feed flux is available to

maintain them, i.e., φ = 40 % w/v is the most direct pathway in this case.

Talmage and Fitch (1955) extended Kynch theory to thickener sizing by calculating the

concentration at each point on the settling curve by drawing a tangent to the batch

settling curve (Figure 2-19), with the intercept giving H1 (Pearse 1977):

1

oo

HHφ

=φ 2-58

where: φ = Solid fraction (φo at t = 0)

H = Height of suspension (Ho at t = 0)

Drawing the tangent to the settling curve accurately from experimental data is difficult

(Pearse 1977), often leading to significant errors in the final result. Kynch did not

consider sediment compression (see Section 2.6.3), so analysis of the settling curve is

limited to concentrations lower than the compression point (φg). Talmage and Fitch

(1955) suggested that the compression point gives the lowest settling velocity within the

hindered-settling region (Pearse 1977).

Yoshioka (1957) and Hasset (1965) extended the ideas of Kynch by using the flux curve

rather than the settling velocity curve (Figure 2-21). This simplifies the interpretation

compared to the Talmage and Fitch method by allowing the experimental and required

flux (also, and somewhat misleadingly, referred to as the operating line) curves to be

plotted together.

The operating line (see Figure 2-21) is drawn from the intercept of the solid feed rate

(y-axis) to the desired underflow concentration (x-axis), such that it is never above the

observed flux rate over the expected range of solid fraction (φf to φv). The operating

line is described by (Pearse 1977):

φ−ψ=φ ur UU 2-59

Page 77: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

76

where: Ur = Settling velocity required to prevent solid reaching the overflow (m s-1)

Uu = Settling velocity due to underflow removal (m s-1)

Equation 2-59 uses the solid settling flux due to bulk fluid flow to the underflow to

calculate the required additional flux from hindered particle settling. If the feed flux or

required underflow concentration (φu) are raised further, a bed of concentration φcrit will

build up and overflow the thickener (Fitch 1975a). In this situation the underflow solid

fraction will also be affected, since not all of the feed solid reaches the underflow.

Figure 2-21: Yoshioka/Hasset combined flux curve, showing the effect of underflow concentration on the possible throughput.

The Yoshioka/Hasset construction has several advantages. By changing the desired

flux and underflow concentration the relationship between throughput and underflow

density can be rapidly investigated, allowing the selection of a suitable compromise.

This construction also identifies a lower conjugate concentration, limiting the possible

feed dilution. If the feed is diluted further the overflow will contain solid, because feed

dilution increases the rise velocity in the clarification zone, even if the thickener

throughput has not otherwise been exceeded. As discussed above, the feed should

ideally be close to the lower conjugate concentration (rather than the maximum of the

flux curve, for example) because a lower solid fraction typically improves the mixing

during flocculation (Pearse 1977).

Page 78: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

77

Fitch (1975a) and Pearse (1977) argue that the lower conjugate concentration is

unstable, requiring thorough mixing of fresh feed with displaced liquor flowing

upwards towards the overflow. In practice it is more likely that the feed slurry will

plunge from the feedwell and spread out as a layer at its hydrostatic equilibrium.

Caution is required when using the combined flux curve (Figure 2-21) to interpret

thickener operation. The only hindered settling concentrations possible in steady-state

operation are the upper and lower conjugate concentrations. These are the only points

where the required, and available, flux rates are equal. Pearse (1977) describes steady

state thickener operation including the behaviour when the unit is under or overloaded.

When the thickener is under-loaded, the effective settling area will naturally contract to

its steady-state equilibrium position, as the hindered settling and compression zones

contract into the thickener’s conical base. Hindered settling has also been modelled by

dynamic simulation (Fitch 1990, Diehl 1997, Bustos et al. 1999).

2.7.3 Compression (high solid fraction)

The settling solid will eventually build up on the base of the vessel, forming a sediment

bed. In this region settling is resisted by both hydrodynamic and mechanical forces.

The critical concentration (φg – gelation) where compression begins is the concentration

where the suspension forms a continuous network that can transmit mechanical force

through the structure (Perry and Green 1997, Chandler and Hogg 1987, Healy et al.

1994).

Highly flocculated suspensions will enter compression at relatively low solid fractions

because highly porous aggregates have a large effective volume. In extreme cases,

flocculated suspensions may be in compression from the onset (Pearse 1977).

However, due to their high porosity and few particle/particle contact points, the

compressive yield strength may initially be very low, and settling may essentially still

follow hindered-settling behaviour as predicted by Kynch type theory (Pearse 1977). In

many cases the transition from hindered to compression settling is so gradual that it is

difficult to detect. In these cases Pearse (1977) and Fitch (1975b) suggest plotting

log(H - H∞) vs. time to help detect the transition (Figure 2-22).

Page 79: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

78

It was realised very early in the development of thickener technology (Pearse 1977) that

increasing the height of the bed increased the underflow concentration, although the

mechanism remained unclear. Coe and Clevenger (1916) proposed that the rate of

compression was a function of time only, and that increasing the height of the bed

merely increased the residence time in the bed at a constant underflow rate. Roberts

(1949) suggested that the rate of removal of interstitial water was also proportional to

the amount present, and proposed:

( )KD

dtDDd

−=− ∞ 2-60

where: D = Units of water per unit solid D∞ = D at infinite time K = Constant

giving a plot as shown by Figure 2-22.

Figure 2-22: Roberts’ plot (adapted from Pearse 1977).

Roberts’ assumptions were not correct (Yoshioka et al. 1957, Fitch 1975b), however

most suspensions (Fitch 1975b, Pearse 1977) give linear relationships on a Roberts’ plot

and Roberts’ work forms the basis of some of the current thickener sizing techniques

(Dahlstrom & Fitch 1985).

Michaels and Bolger (1962) proposed that both the compressive yield strength (P) and

permeability (κ) are functions of the concentration:

Time

Hindered settling regime

Compression regime L

og (H

-H∞)

Page 80: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

79

)(PP y φ= 2-61 )(φκ=κ 2-62 where: P = Bed overburden (N m-2) Py(φ) = Compressive yield strength (N m-2) κ = Permeability (m kg-1) κ(φ) = Dynamic compressibility

They proposed that when the weight of the bed exceeds the compressive yield stress of

the sediment, it will compress at a rate limited by the rate of upward fluid drainage for

the given sediment permeability. These assumptions form the basis of current

compression theory (Buscall & White 1987, Healy et al. 1994, Green et al. 1996):

( )φ≤=φ

yPP ; 0dtd 2-63A

( )[ ] ( )φ>φ−φκ=φ

yy PP ; PP)(dtd 2-63B

The permeability of the sediment is typically calculated from Darcy’s (Wakeman &

Holdich 1984, Perez et al. 1998, Diplas & Papanicolaou 1997, Concha et al. 1996,

p. 62) or Kozeny’s (Fitch 1975b) law of fluid flow through packed beds. Permeability

may also be a function of the dynamic pressure gradient (dP/dL)(Fitch 1975b,c, Kos

1980), which may cause the formation of drainage channels in the sediment structure.

Rakes with vertical pickets have also been used to improve sediment compression

(Dahlstrom & Fitch 1985).

Equation 2-63 is not solvable analytically (Perez et al. 1998), but various numerical

solutions have been described (Wakeman & Holdich 1984, Williams & Amarasinghe

1989, Barker & Grimson 1990, Font 1988, 1990, 1991, Concha et al. 1996, Diehl 1997,

Diplas & Pypanicolaou 1997, Burger & Concha 1998, Perez et al.1998, de Kretser et al.

2000).

Simulating bed behaviour is currently an active topic in settling research. This subject

has become far more important since the introduction of polymer flocculants that

dramatically improve free and hindered-settling velocities, but also bring forward the

Page 81: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

80

onset of compression by forming voluminous porous aggregates (Kos 1980, Hogg &

Bunnaul 1992). Also, the modelling of compression has only been possible since

efficient computerised numerical methods have become available.

The investigation of the compressive yield stress (Py(φ)) is currently an active research

area. Two methods of measuring Py(φ) are popular, centrifugation or pressure filtration.

The use of centrifugation was suggested by Buscall and White (1987), who proposed a

method for determining Py(φ) by measuring the equilibrium sediment height at various

rotation speeds (Landman & White 1992). Green et al. (1996) propose an alternative,

and apparently more accurate, numerical analysis.

Alternatively, the sediment concentration profile can be measured, either by x-ray or

γ-ray gauge, or manually by sectioning the sediment (Berström et al. 1992). This

simplifies the calculation of Py(φ) from experimental data (Green et al. 1996).

Liu and Masliyah (1996) used data from Auzerais et al. and proposed:

φ−φ

φ=φ

max

no

yP

)(P 2-64

where: Po = 30 - 160 kPa φmax = 0.72 n = 3 - 6 Py(φ) can also be determined by pressure filtration at various pressures (Landman et al.

1995), a method that has the additional advantage of also giving sediment permeability

data (via the fluid flow rate).

An increasing solid fraction increases the yield strength rapidly (Fitch 1987). Polymer

flocculants are likely to affect both the sediment yield strength and permeability (Healy

et al. 1994, Figure 2-23).

Page 82: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

81

120000

100000

60000

40000

20000

00.10 0.15 0.20 0.25 0.30 0.35 0.40

Volume Fraction

Com

pres

sive

Yie

ld S

tres

s (Pa

)

80000

ZrO2(0.2 µm diam.)

pH=6.0 7.1 2.9

polyacrylic acidMw~750000

polyacrylic acidMw~2000

no additive

Figure 2-23: Effect of flocculant type on Py(φ) (adapted from Healy et al. 1994).

Increasing the bed height to aid compression is the basis for the deep cone thickener,

which may have a bed height up to 5 m (Healy et al. 1994). However, these units may

require rakes with 3 to 4 times the torque, and underflow pumps capable of pumping

against 3 to 4 kPa m-1 of pipe (Perry & Green 1997).

While high dosages of polymeric flocculants may improve the initial sedimentation rate

by increasing the permeability, they also increase the yield stress, leading to lower final

underflow concentrations (Healy et al. 1994). In this case the action of rakes may be

beneficial, by disrupting the sediment structure. Rakes with vertical pickets have been

used (Dahlstrom & Fitch 1985) to create drainage channels in the sediment, increasing

its permeability.

2.8 Conclusions to literature review Aggregation and settling are both mature areas of research with most phenomena

already well understood. However, the incentive to further improve unit performance

remains, and research and development continues in several key areas, particularly

computer modelling. Although the basic architecture of the various models (e.g. CFD,

population balancing, settling simulations) is already well established, many current

models assume simplified behaviour and do not adequately describe real systems. This

is largely hampered by a lack of experimental data suitable for model construction, with

few of the models supported by credible experimental data, and with many modelling

papers containing no experimental data at all.

Page 83: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

82

The population balance approach to aggregation kinetics modelling was first proposed

by Smoluchowski in 1917, however the majority of the work in this area has been done

in the last decade, since efficient numerical techniques have become available to solve

the equations. This has allowed the population balances to become sufficiently

complex to describe real systems. However, the full potential of population balancing

has yet to be realised and several aspects require more work.

Population balances have already been used successfully to simulate the

aggregation/breakage of aggregates formed by coagulation (e.g. Spicer & Pratsinis

1996a), but apart from the limited work by Pelton (1981) polymer flocculation has been

largely ignored. Flocculation and coagulation are similar in many respects, but the

breakage of flocculated aggregates is generally found to be irreversible, with the

aggregate size decreased permanently by extended shear. This may be a significant

practical issue in some cases, for example in feedwells which subject fully formed

aggregates to high levels of shear (perhaps due to less-than-optimal baffle placement).

Fluid shear has a major impact on the rates of both aggregation and breakage, and most

current population balance models (e.g. Spicer & Pratsinis 1996a) are written to

accommodate changes to the shear rate. However, other process variables (e.g.

flocculant dosage, primary particle size and solid fraction) also impact on the

aggregation process, and require consideration.

Population balance modelling has previously been treated as an isolated research area,

with little effort made to link the predicted aggregate size to important performance

parameters like the settling rate. There also exists the opportunity to begin to couple

various models (e.g. CFD-population balance, or population balance-settling

simulation) and create larger process models to describe the wider process. This would

then allow the unit performance to be optimised as a whole, for example feedwell

design could be altered to give optimal settling performance. In addition to

optimisation, sophisticated dynamic process models are likely to have other

applications, for example in the development of suitable process control strategies.

Page 84: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

83

3. On-line particle sizing

3.1 Particle sizing techniques The measurement of particle or aggregate size distributions has already been well

reviewed elsewhere (Herbst & Sepulveda 1985, Allen 1990, Barth & Sun 1991,

Stanley-Wood & Lines 1992, Farrow & Warren 1993, Perry & Green 1997) and only

brief general comments are offered here.

In practice, particle size distributions tend to be broad, usually several orders of

magnitude, and although there may be only a few large particles, they may contain the

bulk of the mass. For these reasons a particle size distribution is usually plotted as the

particle volume weighted counts per channel (y-axis) against log (size) (x-axis). This

usually produces a distribution that may be approximated by a log-normal distribution,

although alternative distributions are available (Perry & Green 1997, Dahlstrom & Fitch

1985, Randolph & Larson 1988).

Although numerous techniques and instruments are available for determining particle

size, few can be applied in-situ/on-line, or over a wide range of solid concentrations

(Williams 1992). In-situ ultrasonic techniques have been developed (Coghill et al.

1997), but require lengthy measurement times, making them unsuitable for dynamic

systems. Laboratory instruments like the Malvern Mastersizer have also been modified

(Bale & Morris 1987, Hobbel et al. 1991, Bale 1996, Phillips & Walling 1995a) to suit

in-stream sizing, but are bulky and relatively expensive for field or plant use.

Alternatively, in-stream measurements can be made by focussed beam reflectance

measurement (FBRM). The FBRM probe has a wide range of applications, frequently

in the process environment (Table 3-1). The instrument gives a particle chord length

distribution, which is a function of the true particle diameter distribution. For many

process applications (e.g. control) this distinction is not important because the relevant

process variable (e.g. crystal size) can be correlated with some aspect of the chord data,

and accurate particle sizing is not specifically required.

Page 85: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

84

Table 3-1: Summary of common FBRM process applications Application Reference Control

Heffels et al. 1998, Fakatselis 1994, Farrell and Tsai 1995, Peng and Williams 1993, Dijkstra et al. 1996, Reid and Stachowicz 1990

Crystallisation/ precipitation

Sparks and Dobbs 1993, Farrell and Tsai 1995, Dijkstra et al. 1996, Monnier et al. 1997, Johnson et al. 1997, Poilov et al. 1997, Bongartz et al. 1999, Barrett and Glennon 1999, Hildred et al. 2000

Estuarine solids Bale 1996, Phillips and Walling 1995a,b, 1996, 1998, 1999, Law et al. 1997

Flocculation Williams et al. 1992, Peng and Williams 1994, Spears and Stanley 1994, Sengupta et al. 1997, Caron-Charles and Gozlan 1996, Murphy et al. 1994,1995, Sharma et al. 1994, Fawell et al. 1997, Hokanson and Preikschat 1992, Honaker et al. 1994, Hecker et al. 1999, Alfano et al. 2000

Grinding Hokanson and Preikschat 1992, Reid and Stachowicz 1990, Thomas et al. 1998

Paper manufacture Hanseler and McKean 1988, Blanco et al. 1996, Anderson 1996, Alfano et al. 1999

Petroleum Cowie and Schmoll 1994, El-Hamouz and Stewart 1996, Simmons et al. 1999

Pharmaceutical Vo 1994, Daniels and Barta 1992

The instrument’s rapid on-line analysis, and conventional log-normal size distribution

make it attractive to workers modelling the kinetics of flocculation (Peng & Williams

1994, Sharma et al. 1994) or crystallisation (Farrell & Tsai 1995, Monnier et al. 1997,

Poilov et al. 1997) by population balance. For dynamic modelling, accurate sizing and

counting of particles is desired for model parameter estimation. Both theoretical

(see Section 3.3.1) and empirical (see Section 3.3.2) methods for estimating particle

diameter distributions from FBRM chord distribution are reported in the literature.

The aim of the work described in this section was to find a calibration method for

FBRM, allowing the aggregate chord data to be converted to aggregate diameter data

suitable for the population balance modelling described later (Section 5).

Page 86: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

85

3.2 Principle of FBRM The Lasentec FBRM instrument directs a 789 nm laser through a lens rotating at

4410 rpm (Sparks and Dobbs 1993). This results in a rotating beam of light highly

focussed at a point near the instrument’s window. The position of the focal point is

adjustable, and sometimes (Monnier et al. 1996) adjusted depending on the particle

size. The diameter of the circular path of the laser beam is 8.5 mm and the beam is

focussed to a section of approximately 0.7 × 2 μm with a focal depth of 10 μm

(Hokanson & Preikschat 1992). However, the exact width of the beam appears to be

dependent on both the focal position, and the size of the particle and its distance from

the focal position (Reid & Stachowicz 1990, Sparks & Dobbs 1993).

Laser beam

Opticsrotating ata fixed highvelocity

Sapphirewindow

Scanning directionof focusedlaser beam

Reflec

ted ch

ord

Reflec

ted ch

ord

Reflec

ted ch

ord

Probe

Laser

Figure 3-1: Schematic of the FBRM probe.

When the moving laser beam crosses the path of a particle, some of the light is reflected

back to the instrument detector. Since the tangential velocity of the beam is known

(∼ 1.9 m s-1), the duration of the reflected light pulse is directly proportional to the width

of the particle intersected. The tangential beam velocity is assumed (Section 3.3.1) to

be much larger than the particle velocity relative to the instrument window. However,

since it is unlikely that the laser beam will pass directly across the centre of the particle

(i.e. diameter), a chord length is measured.

Several descriptions are given of previous versions of the FBRM instrument and data

processing regimes (Hokanson & Preikschat 1992, Sparks & Dobbs 1993, Reid &

Stachowicz 1990). Data processing of the detector response uses both the signal

strength and slope. Current instrument data processing is adjustable through the use of

either F (fine) or C (coarse) electronics modules depending on the particle size.

Page 87: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

86

3.3 Estimating particle size from FBRM data 3.3.1 Theoretical methods

Several papers deal with the transformation of chord to diameter distributions from first

principles (Clark & Turton 1988, Jakeman & Anderssen 1975, Hobbel et al. 1991,

Williams et al. 1992, Liu & Clark 1995, Liu et al. 1998, Tadayyon & Rohani 1998,

Barrett & Glennon 1999, Simmons et al. 1999). Most workers assumed that the

particles are spherical to simplify the mathematics, and most analysis of non-spherical

particles is generally limited to two dimensional ellipses (Jakeman & Anderssen 1975,

Clark & Turton 1988, Liu & Clark 1995, Liu et al. 1998, Tadayyon & Rohani 1998).

Assuming the particles are spherical dramatically simplifies the mathematics, because

regardless of the particle’s orientation it always presents a circular profile (Figure 3-2).

This assumption may be valid for suspensions of spherical latex or glass beads, or

emulsions of two immiscible liquids. The spherical-particle assumption may also be

essentially valid for suspensions of particles with an aspect ratio (ratio major:minor

axis) approaching 1:1. As the aspect ratio increases and the particles or crystals tend

towards needles/plates, the chord distribution broadens (Williams et al. 1992, Barrett

and Glennon 1999).

Most descriptions assume that the focussed laser beam has no width, and that light is

reflected immediately after the beam reaches the edge of the particle, however Hobbel

et al. (1991) introduce a parameter (k) to account for the “arch of rejected reflection”.

Sparks and Dobbs (1993) considered the physics of the laser/lens interaction and

suggest that the beam may be broadened in some conditions. Spreading of the beam

away from the optimal focal point may also affect the chord measurement (Reid &

Stachowicz 1990), and beam spreading is likely to be a function of solids loading, fines,

solution and particle surface characteristics.

Page 88: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

87

Figure 3-2: Method of determining the chord distribution expected from a spherical particle of diameter Dp.

Where: Dp = Diameter of pth particle (μm) Ci,u = Length of the upper boundary of the ith chord channel (μm) Ci,l = Length of the lower boundary of the ith chord channel (μm)

Wp,i = Width of pth particle giving chords in the ith channel (μm) Figure 3-2 shows that the likelihood of a particle being detected is proportional to its

diameter, introducing a bias. Therefore, for modelling purposes FBRM is assumed to

measure the first diameter moment of the chord distribution. Calculating a chord

distribution from a given spherical particle distribution is relatively straightforward,

however the reverse algorithm is more difficult, usually requiring a numerical solution

in the form (Tadayyon & Rohani 1998, Simmons et al. 1999):

( )∑=

−=φmax

1 i

2*ii

NNNmin

*p

3-1

where: Ni = Number of i sized chord counts observed (s-1)

Np* = Number of pth sized particles predicted (s-1)

Ni* = Number of i sized chords predicted (s-1):

( ) ( )∑=

−−−=max

1i

2u,i

2p

2l,i

2p

*p

*i

21

21

CDCDNN 3-2

where:

Ci,u Dp Ci,l

2 × 0.5Wp,i

Particle

Path taken by laser beam at Ci,l

Page 89: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

88

Dp = Diameter of pth particle (μm) Np

* = Number of pth sized particles predicted (s-1) Ci,u = Length of the upper boundary of the ith chord channel (μm)

Ci,l = Length of the lower boundary of the ith chord channel (μm) The resulting calculated diameter distribution is a number distribution, and is converted

to the conventional particle volume distribution by applying a length-cube weighting in

the form of Equation 3-3 (see Section 3.5.1).

3.3.2 Empirical methods

FBRM results are also compared with alternative particle sizing techniques (Table 3-2),

for a variety of particle sizes and materials. Early Lasentec FBRM software included a

spherical-equivalent size algorithm (Sparks & Dobbs 1993, Peng & Williams 1993,

1994, Caron-Charles & Gozlan 1996, Law et al. 1997, Phillips & Walling 1998).

Comparisons between the resulting average size and alternative sizing methods showed

that the FBRM oversized small particles (<150 μm Law et al. 1993, <300 μm Phillips &

Walling 1998) and undersized larger ones (>500 μm Phillips & Walling 1998).

Current software computes various chord statistics: median, mean and mode with

various chord weightings. Daymo et al. (1999) found good agreement between the

cube-weighted mean size and sieve analysis of kaolin suspensions with a mean size of

around 55 μm. Alfano et al. (2000) show a good correlation between Malvern

Mastersizer volume-average size and un-weighted FBRM chord lengths for suspensions

of aggregated silica 20-200 μm. However, the un-weighted FBRM results

underestimated the size of the larger particles. El-Hamouz and Stewart (1996)

compared FBRM and Malvern square-weighted mean drop sizes of oil/water emulsions

and found the FBRM overestimated smaller (-180 μm) drops, but underestimated larger

(+180 μm) drops.

Page 90: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

89

Table 3-2: FBRM comparisons with other sizing techniques Reference Particle material Alternative sizing

method (size) Size by FBRM

Hobbel et al. 1991

Glass beads

Helos, laser forward scattering (66 μm)

170 μm

Sparks & Dobbs 1993

Duke standards, carbon, sand, alumina, hydrate.

Duke certified standards, Microtrac (laser forward scattering) (0.8-410 μm)

Spherical equivalent, FBRM(μm)=52.353+1.0362Duke(μm) 10-162 μm

Murphy et al. 1994

Ground silica.

BET, SEM, Micrometrics sedigraph, (2.1 μm)

Oversized considerably, could not be calibrated in this range

Monnier et al. 1995

Glass beads, latex, micronized pharmaceutical product. various supporting fluids

Cilas (laser diffraction) Coulter Counter Microscopy (4.1-396 μm)

Oversizes < 100 μm Undersizes > 100 μm

Dijkstra et al. 1996

Sugar. Sieved (30-250 μm)

Sieve(μm)=2.044E-3(FBRM μm)2.2844

El-Hamouz & Stewart 1996

Oil/water emulsions Malvern Mastersizer (d32 = 60-350 μm)

Sauter mean (d32 = 130-210 μm), Undersized > 180 μm Oversize < 180 μm

Law et al. 1997 Pollen, glass and latex Duke standards, fractionally settled sediment, sieved sand. 6.5-766 μm

Coulter Counter Malvern Mastersizer Certified standards Optical Microscopy

FBRM Par-Tec spherical equivalent, Undersized > 500 μm Oversized < 150 μm

Phillips & Walling 1998

Sieved river sediment

Coulter laser diffraction (1-355 μm)

FBRM Par-Tec spherical equivalent, Coulter d50(μm)=-19.239+1.064FBRM(μm)

Tadayyon & Rohani 1998

Sieved ion-exchange resin.

Sieve (212-800 μm). FBRM undersized, better with cube-weighting

Daymo et al. 1999

Graphite, gibbsite, bentonite, mica, silica, kaolin, plastic beads

Sieve, Horiba sedigraph. (0.3-663 μm)

5 % w/v kaolin = 53 μm by cube-weighted FBRM, 57 μm by sedigraph. Other samples gave week correlation (order of magnitude)

Alfano et al. 2000

Mircrocrystalline cellulose.

Malvern Mastersizer (20-180 μm)

FBRM(μm)=10.09+0.586Malvern(μm)

Page 91: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

90

3.4 Experimental 3.4.1 Equipment

Suspensions of various particle size and solid fraction were presented to a Lasentec

M500 FBRM probe as shown by Figure 3-3. The focal point was set to 0 μm (the

window) as calibrated by the manufacturer unless otherwise stated. This position was

also confirmed experimentally by placing a film of marking pen ink on the window and

adjusting for maximum counts. The impeller (Figure 3-3) was set to 650 rpm unless

otherwise stated. Data was collected using 90 log-channels unless otherwise stated.

Window cleaning was performed regularly (∼ every 5 minutes measurement time) to

give a total count rate < 50 s-1 in air.

Figure 3-3: Schematic of the laboratory equipment.

A MS10 Malvern Mastersizer and a Coulter Multisizer II were used as alternative sizing

techniques. Particles were dry sieved using conventional Endecotts screens (38, 45, 53,

63, 75, 106, 125, 150, 180, 212, 250, 300, 425, 500, 600, 710, 850, 1000, 1180 μm) in a

Rotap RX-29-10 mechanical shaker, and smaller size fractions obtained by cyclosizing

with a Warman M-4 cyclosizer. SEM (scanning electron microscope) micrographs are

from a GEOL JSM-5800LV.

FBRM probe Stirred suspension

Instrument

Focal position adjusting knob

Stirrer

Fiber optic cable

Page 92: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

91

3.4.2 Samples

Most samples were either aluminium (Valimet, Stockton California) or calcite (Omya

Southern, Commercial Minerals) particles that were sized to narrow fractions by

screening or cyclosizing. The solids were weighed out on a 0.1 mg balance and made

up to concentrated stock suspensions in a volumetric flask. A series of more dilute

suspensions were prepared by serial dilution of the stock suspension from a stirred and

baffled tank to prevent stratification. The calcite samples suffered from fines and where

possible were washed and decanted. Suspensions of aluminium particles were reactive,

producing a gas (presumably hydrogen) and were kept on ice to reduce the reaction rate.

A single suspension of mono-sized latex spheres (Dynaspheres, Duke Scientific) with a

specified size of 19.9 μm was also measured.

3.5 Results and discussion 3.5.1 Effect of applying weightings to FBRM chord distributions

The raw chord lengths measured by the FBRM are sorted into a series of size intervals

(channels) covering the range 1 – 1000 μm (Figure 3-4). The channel widths (μm) are

discretised according to a numerical progression (following Ci,l = 1.08i-1, i = [1, 90]),

although the FBRM software has various options, including constant width channel

spacing.

0

200

400

600

800

1000

1200

1 10 100 1000Chord length (µm)

Cou

nts p

er c

hann

el (s

-1)

Sieved 38-45 µmSieved 45-53 µmSieved 53-63 µmSieved 63-75 µmSieved 75-106 µmSieved 106-125 µmSieved 125-150 µm

Figure 3-4: Effect of particle size on chord distribution. Sieved aluminium particles, 10 % w/v, F-electronics.

Page 93: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

92

Figure 3-5 shows the effect of applying length weightings (moments) to the chord

distribution given by the FBRM, i.e.:

n

A,io,in,i CNN = 3-3 where: Ni,n = number of n-weighted counts in the ith channel (μmn s-1) Ni,o = raw un-weighted counts in the ith channel (s-1)

n = exponent, 0 = un-weighted, 1 = length-wtd, 2 = square-wtd, 3 = cube-wtd, etc Ci,A = geometric average length of ith channel (μm) = (Ci,u × Ci,l)1/2 Ci,u = length of the upper boundary of the ith channel (μm)

Ci,l = length of the lower boundary of the ith channel (μm)

0

1

2

3

4

5

6

7

8

1 10 100 1000Length (µm)

% c

ount

s per

cha

nnel

(FB

RM

)

0

5

10

15

20

25

30

% c

ount

s per

cha

nnel

(M

alve

rn)

FBRM (un-weighted)

Malvern

FBRM (square-weighted)

FBRM (length-weighted)FBRM (cube-weighted)

Figure 3-5: Effect of applying various length weightings (moments) to the FBRM chord distribution. Malvern Mastersizer volume weighed distribution also shown. Aluminium particles, sieved 45-53 μm.

This is analogous to applying a length-cube weighting to a particle number distribution

to give a conventional volume-weighted distribution. For particles between

approximately 50 to 400 μm, Figures 3-5 and 3-6 show that the square-weighted (n =

2) chord distribution gave the best correlation with Malvern Mastersizer results,

although the FBRM distributions remained broader.

A previous version of the FBRM software (Sparks & Dobbs 1993, Peng & Williams

1993, 1994, Caron-Charles & Gozlan 1996, Law et al. 1997, Phillips & Walling 1995A)

included a chord-to-spherical-equivalent algorithm that both narrowed the distribution

and increased the length weighting (Phillips & Walling 1995a).

Page 94: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

93

3.5.2 Comparison of average particle size by various techniques

Particle size distributions can be weighted and averaged in a variety of ways (Randolph

& Larson 1988, Rushton et al. 1996, Seville et al. 1997, Perry & Green 1997). FBRM

software provides a range of options including: mean, median (d50) and mode averages

of length-weighted chord distributions.

Figure 3-6 compares various length-weighed median FBRM average sizes with Malvern

D50 and sieve sizing for a range of sieved or cyclosized fractions. Generally, smaller

particles (< 50 μm) were best estimated by a length-weighed FBRM average. Particles

in the range approximately 50 to 400 μm were best estimated by the square-weighted

average. Suspensions of very large particles (> 400 μm) benefited from cube-

weighting, which has the effect of further weighting the distribution towards larger

particles. The Malvern (as opposed to the sieve) results are plotted on the x-axis

because the smaller fractions were prepared by cyclosizing.

y = 1.1032x + 21.386R2 = 0.9881

y = 0.8936x + 12.453R2 = 0.9927

y = 0.8847x + 4.8684R2 = 0.9918

y = 0.7492x + 5.6964R2 = 0.9921

y = 0.5304x + 2.5324R2 = 0.9808

0

50

100

150

200

0 50 100 150 200 250

Malvern D50 (µm)

Mea

sure

d si

ze (µ

m)

FBRM chord cube-weighted

FBRM chord un-weighted

FBRM chord length-weighted

FBRM chord square-weighted

Sieve

Figure 3-6: Comparison of median average sizes given by various techniques and chord weightings, showing the square-weighted chord gives the best agreement with Malvern and sieve analysis. Sieved/cyclosized aluminium fractions. F-electronics, FBRM suspensions 5 % w/v.

The mean, and to a lesser extent also the median, averages were affected by the FBRM

instrument’s sensitivity to fine particles (Figures 3-11 & 3-13). However, the mode

(distribution maxima) were less affected (Figure 3-7). Figure 3-7 shows direct

comparison of average sizes by the techniques for a series of sieved or cyclosized

Page 95: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

94

suspensions. Linear correlation equations (Figure 3-7) could be re-arranged to give a

2-parameter calibration, to give reasonable agreement over 20-500 μm.

y = 0.8108x + 8.0231R2 = 0.9983

y = 0.8752x - 10.783R2 = 0.9903

y = 0.8157x + 0.5694R2 = 0.9927

0

100

200

300

400

500

600

700

800

0 100 200 300 400 500 600 700 800

Malvern D50 (µm)

Mea

sure

d av

erag

e pa

rtic

le si

ze (µ

m)

FBRM mode of square-weight

Coulter

Sieve

Figure 3-7: Comparison of mode average sizes from the FBRM square-weighted distributions, with median values from Malvern and Coulter analysis. Sieved/cyclosized calcite particles, 5 % w/v. F-electronics.

Large particles can be obscured if the suspension contains large quantities of fines

(see Section 3.7). In this case even the square-weighted distribution may become

bimodal, and applying the cube-weighting and manually assigning the mode to the right

hand peak may extend the measurable range. In some situations, decanting the fines

may be feasible.

3.5.3 Effect of solid volume fraction on indicated size

Increasing the solid volume fraction has the effect of increasing the FBRM instrument’s

sensitivity to fine particles, probably due to fine particles crowding the measurement

zone near the window (see Section 3.5.7). However, this effect is largely overcome by

applying the square-weighting, (Figures 3-8 & 3-9). Applying a cube-weighting may

lead to further improvements, especially for suspensions of large particles.

Page 96: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

95

0

50

100

150

200

250

300

350

400

1 10 100 1000

Chord length (µm)

0.0E+00

5.0E+05

1.0E+06

1.5E+06

2.0E+06

2.5E+06

3.0E+060.1 % w/v, un-weighted0.5 % w/v, un-weighted1 % w/v, un-weighted2 % w/v, un-weighted5 % w/v, un-weighted10 % w/v, un-weighted20 % w/v, un-weighted0.1 % w/v, square-weighted0.5 % w/v, square-weighted1 % w/v, square-weighted2 % w/v, square-weighted5 % w/v, square-weighted10 % w/v, square-weighted20 % w/v, square-weighted

Figure 3-8: Effect of solid fraction on the un-weighted and square-weighted FBRM distributions. Aluminium particles sieved 125-150 μm, F-electronics.

Previous FBRM software calculated a spherical-equivalent size. Various studies

showed that the spherical-equivalent size was essentially unaffected by solid fraction

(Sparks & Dobbs 1993, Law et al. 1997). Daymo (1999) used the chord cube-

weighting and also found little variation in the indicated size as the solid fraction was

varied.

0

100

200

300

400

500

600

0.1 1 10 100Weight fraction (% w/v)

Cal

cula

ted

size

from

FB

RM

(µm

) sieved 38-45 µmsieved 45-53 µmsieved 53-63 µmsieved 63-75 µmsieved 75-106 µmsieved 106-125 µmsieved 125-150 µmsieved 150-180 µmsieved 180-212 µmsieved 212-250 µmsieved 250-300 µmsieved 300-425 µmsieved 425-500 µm

Figure 3-9: Effect of solid fraction on the square-weighted mode of the chord length, calcite particles, F-electronics.

Figure 3-8 shows the influence of solid fraction on the un-weighted FBRM chord

distributions. The chord-to-diameter models described in Section 3.3.1 (Equation 3-2)

Page 97: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

96

rely on the un-weighted distribution, and gave results that were dependent on the solid

fraction (Figure 3-10). This effect limits the usefulness of algorithms like Equation 3-

2, and further model parameters would be required to compensate for this effect. The

calculated distributions are also noticeably broader than the Malvern volume-weighted

distribution, despite the particles (Figure 3-16) approaching the sphericity assumed by

the model.

A high flow velocity past the probe window could be suspected to cause this effect, via

both positive and negative Doppler shifts depending on the direction of the particle

relative to the circular path of the laser beam. However, changing the impeller speed

had little effect on the FBRM distribution, despite the impeller passing 2 mm from the

instrument window at a velocity of approximately 2 m s-1 at 750 rpm. The cause of

FBRM distribution broadening remains unclear, but may be a combination of particle

shape/reflectivity, width of the viewing zone (affected by particle concentration), and

data processing (C or F electronics).

0

5

10

15

20

25

30

1 10 100 1000Particle size (µm)

% p

er c

hann

el (M

alve

rn)

0

1

2

3

4

5

6

7

8

9

10

% P

er c

hann

el (F

BR

M)

Malvern

FBRM calc. vol. dist. (0.1 % w/v)

FBRM calc. vol. dist. (0.5 % w/v)

FBRM calc. vol. dist. (1 % w/v)

FBRM calc. vol. dist. (2 % w/v)

FBRM calc. vol. dist. (5 % w/v)

FBRM calc. vol. dist. (10 % w/v)

FBRM calc. vol. dist. (20 % w/v)

Figure 3-10: Effect of solid concentration on transformed chord data, showing that the model (Equation 3-2) overestimated the particle size in dilute suspensions. Aluminium particles sieved 38-45 μm, F-electronics.

Page 98: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

97

3.5.4 Comparison between C and F-electronics

As mentioned previously, the FBRM instrument is available with either C (coarse) or F

(fine) electronics modules. As the names suggest, F-electronics is preferred for

suspensions of fine particles, and C-electronics for coarse fractions or aggregates.

Modified signal processing using C-electronics allows aggregates to be measured as

single, large particles, rather than a series of fine particles.

Figure 3-11 shows the results from F and C-electronics using the same suspension of

calcite particles sieved between 38 and 45 μm. Fine particles abraded from the solid

surface during mixing are readily detected with F-electronics, with the C-electronics

recording more counts in the higher channels.

0

100

200

300

400

500

600

700

800

1 10 100 1000

Length (µm)

Un-

wei

ghte

d co

unts

per

ch

anne

l (s-1

)

0.0E+00

5.0E+05

1.0E+06

1.5E+06

2.0E+06

2.5E+06

Squa

re-w

eigh

ted

coun

ts p

er

chan

nel (

µm2 s-1

)

F-electronics, unweighted

C-electronics,square-weighted

F-electronics, square-weighted

C-electronics, unweighted

Figure 3-11: Calcite particles sieved 45-53 μm, 5 % w/v, F-electronics.

Figure 3-12 shows the square-weighted median sizes given by F and C-electronics for a

range of sieved or cyclosized calcite particles. As expected, smaller particles

(< ∼250 μm) were more accurately sized by F-electronics, with C-electronics being

slightly preferable for larger fractions.

Page 99: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

98

y = 0.6837x + 36.903R2 = 0.994

y = 0.7161x + 15.658R2 = 0.9953

y = 0.8119x - 2.5398R2 = 0.9961

0

50

100

150

200

250

300

350

400

0 50 100 150 200 250 300 350 400 450

Malvern D50 (µm)

Mea

sure

d si

ze (µ

m)

Sieve (calcite)

Calcite, F-electronics

Calcite, C-electronics

Figure 3-12: Comparison between square-weighted median FBRM results given by C and F-electronics for various sieved or cyclosized calcite fractions. FBRM suspensions 5 % w/v.

3.5.5 Effect of particle material

Particle shape (Williams et al. 1992, Barrett & Glennon 1999) and reflectivity (Reid &

Stachowicz 1990, Sparks & Dobbs 1993) were expected to affect the reflection of the

laser beam to some extent. Figure 3-13 shows a comparison between aluminium and

calcite particles of the same size and solid fraction. Aluminium and calcite powders

have the advantage of similar densities (ρal. = 2 702, ρcal. = 2 710 kg m-3), giving them

similar volume fractions when made up by weight, and similar flow requirements to

keep the particles suspended.

The increased fine counts shown by the (un-weighted) calcite distribution (Figure 3-13)

are again probably caused by fine material abraded from the surface during suspension

mixing. The counts rate increased in this region with time, but was reduced by washing

and decantation. Aluminium particles were also experimentally inconvenient, as they

were reactive, producing a gas, (presumably hydrogen) and required careful

preparation, storage and disposal (see Section 3.4.2).

Generally however, little difference was observed between either the square-weighted

distributions (Figure 3-13), their median sizes (Figure 3-14), or the overall count rate

(Figure 3-13), despite the two samples having different morphologies, and presumably

Page 100: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

99

also different reflectivities (see SEM images, Figures 3-15 & 3-16). A colour

difference between the aluminium and calcite particles was visible to the naked eye, but

not clear in the SEM images (Figures 3-15 & 3-16) where the calcite particles were gold

plated for imaging. Daymo et al. (1999) added food colouring to a suspension of 4 %

w/v bentonite and reported no change in the average size. However, Sparks and Dobbs

(1993) report that transparent particles (oil drops, latex particles) were generally poorly

sized by FBRM.

0

100

200

300

400

500

600

700

800

1 10 100 1000

Length (µm)

Un-

wei

ghte

d co

unts

per

ch

anne

l (s-1

)

0.0E+00

2.0E+05

4.0E+05

6.0E+05

8.0E+05

1.0E+06

1.2E+06

1.4E+06

1.6E+06

Squa

re-w

eigh

ted

coun

ts p

er

chan

nel (

µm2 s-1

)

Calcite particals, un-weighted

Calcite particals,square-weighted

Aluminium particals,square-weighted

Aluminium particals, un-weighted

Figure 3-13: Comparison between FBRM distributions of aluminium or calcite. Sieved 45–53 μm, 5 % w/v, F-electronics.

0

50

100

150

200

250

300

350

400

0 50 100 150 200 250 300 350 400 450

Malvern D50 (µm)

Mea

sure

d si

ze (µ

m)

Sieve (aluminium)Sieve (calcite)Aluminium, F-electronicsCalcite, F-electronics

Figure 3-14: Comparison between square-weighted median FBRM results given by various sieved or cyclosized calcite and aluminium fractions. FBRM suspensions 5 w/v, F-electronics.

Page 101: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

100

Figure 3-15: SEM secondary electron image of the un-sieved calcite particles.

Figure 3-16: SEM secondary electron image of un-sieved aluminium particles

Mono-sized (19.9 μm) spherical latex spheres (Duke Dynaspheres) were also sized by

FBRM, and gave mean sizes of 26 μm (F-electronics, un-weighted), 51 μm

(F-electronics, square-weighted), 36 μm (C-electronics, un-weighted), 66 μm

(C-electronics, square-weighted). Similar observations have been reported previously

Page 102: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

101

(Sparks & Dobbs 1993), where transparent particles (laser = 789 nm) are sized poorly

by FBRM, presumably due to complex reflection/refraction effects.

3.5.6 Minor effects

A number of additional minor effects may alter the instrument response. The position

of the focal point of the laser can be manually adjusted in relation to the instrument’s

window. This has the effect of moving the position of the measurement region further

into the pulp, or, as negative measurements indicate, behind the window/suspension

interface. Larger particles may have trouble entering the measurement region when it is

set to the window (= 0 μm, Figure 3-17), leading to an exaggerated sensitivity to fine

particles. Monnier (1996) and Law et al. (1997) suggest setting the focal position

further into the suspension for larger particles. In some cases setting the focal position

approximately 20 μm behind the window’s outer face will also improve the sizing of

large particles (Becker 1999).

0

50

100

150

200

250

300

350

400

1 10 100 1000Chord length (µm)

Cou

nts p

er c

hann

el (s

-1)

focal posn. -100 µm, un-weighted

focal posn. -50 µm, un-weighted

focal posn. -20 µm, un-weighted

focal posn. -10 µm, un-weighted

focal posn. 0 µm, un-weighted

focal posn. 10 µm, un-weighted

focal posn. 20 µm, un-weighted

focal posn. 50 µm, un-weighted

focal posn. 100 µm, un-weighted

Figure 3-17: Effect of laser focus position on chord distribution calcite, sieved 106-125 μm, 20 % w/v, 500 rpm impeller speed. Un-weighted. (See Figure 3-18 for square-weighted).

However, as was the case previously, applying a square-weighting (Figure 3-18)

normalises the response considerably. Adjusting the focal position to the window

( = 0 μm) increased the count rate and so improved the statistical accuracy. Except

Page 103: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

102

Figures 3-17 and 3-18, all other data presented is with the focal position set to the

window.

0.0E+00

1.0E+06

2.0E+06

3.0E+06

4.0E+06

5.0E+06

1 10 100 1000Chord length (µm)

Cou

nts p

er c

hann

el (µ

m2 s-1

)focal posn. -100 µm, square-weighted

focal posn. -50 µm, square-weighted

focal posn. -20 µm, square-weighted

focal posn. -10 µm, square-weighted

focal posn. 0 µm, square-weighted

focal posn. 10 µm, square-weighted

focal posn. 20 µm, square-weighted

focal posn. 50 µm, square-weighted

focal posn. 100 µm, square-weighted

Figure 3-18: Effect of laser focus position on chord distribution of calcite, sieved 106-125 μm, 20 % w/v, 500 rpm impeller speed. square-weighted. (See Figure 3-17 for un-weighted).

The tangential velocity of the focussed laser beam is approximately 1.9 m s-1, and is

assumed (see Section 3.3.1) to be far greater than the particle velocity with respect to

the instrument window. The tip velocity of the supplied 50 mm diameter pitched blade

impeller reaches 2 m s-1 at approximately 750 rpm. Other flow geometries such as pipes

or process streams may also give flow velocities in this range. However, since the

viewing region is adjacent to the window, the flow velocities in the viscous sub-layer

are likely to be considerably lower than for the bulk of the fluid flow.

In practice, changing the impeller speed had little effect (Figure 3-19) on the observed

square-weighted FBRM distribution. Larger particles (> 500 μm) tended to stratify at

low stirring speeds (< 500 rpm), leading to a reduction in the average measured size.

Extended high-speed stirring of the calcite particles lead to increased fines via surface

abrasion. Except Figure 3-19, all measurements were made with the impeller set to

650 rpm.

The effect of fluid flow rate on the measured size has already been investigated, with

contradictory results reported. Law et al. (1997) showed no significant effect of stirring

Page 104: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

103

speed on the indicated size in the range 500-1000 rpm. Monnier et al. (1996) found that

increasing the impeller speed from 500-1000 rpm dramatically increased both the

measured size and total counts of glass beads. Daymo et al. (1999) reported that

increasing the flow velocity of a kaolin suspension from 1.3 to 2.4 m s-1 in a 76 mm ID

pipe reduced the indicated size by 13 %. However, these changes may have been due to

particle setting at low flow rates, or particle breakage at high flow rates.

0.0E+00

1.0E+07

2.0E+07

3.0E+07

4.0E+07

5.0E+07

1 10 100 1000

Length (µm)

Squa

re-w

eigh

ted

coun

ts

per

chan

nel (

µm2 s-1

)

250 rpm375 rpm500 rpm750 rpm1000 rpm

Figure 3-19, extra figure showing effect rpm. Calcite particles sieved 106-125 mm, 20 % w/v. Duplicate results performed with rpm decreasing with time showed less fines at higher rpm.

A short (30 s) period of stirring was sufficient to disperse the solid. The measurements

were then taken rapidly before excessive fines were produced from surface abrasion.

The effect of dispersants on the suspension was also investigated, but gave no

significant improvement over the short period of stirring. In other systems the use of

dispersants (Monnier et al. 1995) can be advantageous.

Some particles are prone to adhere to the window (Coghill et al. 1997), especially fine

particles, or when flocculant is added. The particles build up on the window, increasing

the count rate and influencing the measured size. All results shown in this work were

obtained with regular (approx. every 5 min.) window cleaning to give a total

background count (in air) of less than 50 s-1. In many situations such regular cleaning

may not be feasible, although it is recommended that the probe is positioned so that the

Page 105: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

104

fluid flow helps keep the window clear (Sparks & Dobbs 1993, Lasentec Users

Manual).

3.5.7 Effect of solid fraction on particle counts

Estimation of total particle number is desirable for mass balancing or estimation of

aggregate porosity. Unfortunately, total FBRM counts did not correlate well with solid

fraction, tapering off at high particle concentrations (Figure 3-20). Similar results are

reported by Barrett and Glennon (1999) and Daymo et al. (1999). This effect is

probably due to an increase in the instrument dead-time at higher solid fraction, because

the instrument only counts one particle at a time. At a low solid fraction only a small

portion of the scan time is used detecting particles, and the likelihood of particles

overlapping in the viewing zone is low. However, at a higher solid fraction, a

significant portion of the time is spent measuring particles, reducing the time available

to detect other particles in the measurement zone (e.g. behind the measured particle).

This dead-time will be slightly larger than the time spent traversing the particle, because

the reflected light must return to zero for a short period between particles. The signal

processing algorithm also requires that the reflected light pulse to rise and fall quickly

before and after each particle, and counts from overlapping particles (or particles

outside the viewing zone) are rejected, further increasing the dead-time.

0

5000

10000

15000

20000

25000

0 5 10 15 20 25 30

Solid fraction (% w/v)

Tot

al c

ount

s (s-1

) Sieved 38-45 µmSieved 45-53 µmSieved 53-63 µmSieved 63-75 µmSieved 75-106 µmSieved 106-125 µmSieved 125-150 µmSieved 150-180 µmSieved 180-212 µmSieved 212-250 µm

Figure 3-20: Relationships between total counts and solid fraction for various sieved calcite fractions. F-electronics.

Page 106: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

105

Figure 3-21 shows the effect of correcting for instrument dead-time according to:

( )

⎟⎟

⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

⎛+−×

×=

∑=

max

1 iA,ii

6

6

t*t

k2CN109.1

109.1NN 3-4

where: Nt

* = Corrected total particle counts (s-1) Nt = Total observed particle counts (s-1) 1.9×106 = Velocity of laser spot (μm s-1) Ni = Number of ith sized chord counts observed (s-1) Ci,A = Geometric average length of ith channel = (Ci,u × Ci,l)1/2 (μm) k = Additional dead time required before (and after) particle (μm) Equation 3-4 includes a parameter (k) to estimate the additional dead time required

before and after the laser intersects the particle. In this case k = 26.3 μm (calcite,

F-electronics) was found to produce the most linear relationships, and was found

numerically by maximising the sum of the correlation coefficients. Performing the

same optimisation on the other possible data sets gave values of k = 33.3 μm (calcite,

C-electronics), 10.9 μm (aluminium, F-electronics), 21.3 μm (aluminium,

C-electronics).

0

20000

40000

60000

80000

100000

120000

0 5 10 15 20 25 30

Solid fraction (% w/v)

Cor

rect

ed to

tal c

ount

s (s-1

)

Sieved 38-45 µmSieved 45-53 µmSieved 53-63 µmSieved 63-75 µmSieved 75-106 µmSieved 106-125 µmSieved 125-150 µmSieved 150-180 µmSieved 180-212 µmSieved 212-250 µm

Figure 3-21: Relationships between corrected total counts and solid fraction for various sieved calcite fractions. F-electronics. k = 26.3 μm.

Page 107: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

106

3.6 Conclusions The FBRM instrument gives rapid, on-line, in-situ, particle size analysis of concentrated

pulps. While the instrument gives a particle chord distribution, as opposed to a

conventional diameter based distribution, mean or mode averages of the square-

weighted FBRM sizes are comparable to conventional sizing techniques over the range

approximately 50-400 μm.

This range may be extended by using the length-weighted average for smaller particles,

and by applying a cube-weighting to larger particles. However, using a lower

weighting increases the instrument’s susceptibility to changes in solid fraction, while

large particles may be difficult to size accurately due to obscuration by fines. This

effect also appears to result in a poor correlation between total particle counts and solid

fraction. However, FBRM square-weighed average sizes were found to be essentially

independent of suspension bulk flow velocity, focal position, or solid fraction in the

range 0.1-20 % w/v. Although the FBRM technique may have slightly less accuracy

and require more calibration compared to other sizing techniques, it is currently the

most suitable commercially available on-line and in-stream sizing probe. It is therefore

the best technique for aggregation studies where sub-sampling and dilution will disrupt

fragile aggregates (see Section 4).

Page 108: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

107

4. Calcite flocculation in turbulent pipe flow

This section describes experimental work to measure aggregation kinetics in turbulent

flow, providing data suitable for the population balance model described in Section 5.

Experimental studies of aggregation kinetics are relatively difficult to perform, and may

suffer a range of scale-related problems (Keys & Hogg 1978, Shamlou 1993). Several

laboratory-scale reactor designs have been used successfully, with a simple stirred and

baffled tank being the most popular (Ives 2000). A stirred tank reactor has the

advantage of readily suspending solid particles at a comparatively low average shear

rate (Shamlou 1993), however the shear rate varies considerably through the tank, and

is typically 5-10 times higher around the impeller (Koh et al. 1984, Shamlou &

Titchener-Hooker 1993, Perry & Green 1997). The spatially averaged shear rate (G) is

typically calculated from the mean energy dissipation rate (Cleasby 1984, Bird et al.

1960, p. 157, Davies 1972, p. 7, Clark 1985, p. 749, Perry & Green 1997, p. 6-45, Chen

& Jaw 1998, p. 6-18):

ν∈

=G 4-1

where: ∈ = Energy dissipation rate per unit mass (J s-1kg-1, m2s-3) ν = Kinematic viscosity (m2s-1)

In a stirred tank the energy input can be estimated from the impeller power number

(Spicer & Pratsinis 1996, Perry & Green 1997, Ives 2000), or measured directly from

the motor torque and rpm (Lu et al. 1998, Svarovsky 2000, p. 141). However, a stirred

tank reactor is not suited to continuous flow for aggregation kinetics experiments, due

to the broadening of the residence time distribution (Nauman & Clark 1991).

Restricting the laboratory vessel to a batch process leads to difficulties if aggregate

sizing or settling rate measurements have to be taken ex-situ. The action of removing

the suspension of fragile aggregates (e.g. by syringe or peristaltic pump) can alter the

aggregate size (Williams 1992, Spicer et al. 1998). Further, if the required sub-sample

is large the volume remaining in the tank reactor may be significantly reduced,

increasing the power dissipation rate per unit mass.

Page 109: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

108

At the relatively high fluid shear (~10-1000 s-1) and solid fraction (~1-10 % w/v)

(Dahlstrom & Fitch 1985, Williams & Simons 1992) found in mineral processing

thickener feedwells, aggregation proceeds rapidly after flocculant/feed mixing and

adsorption. The short reaction time serves as a further incentive to avoid time-

consuming sub-sampling.

Another popular laboratory vessel for aggregation studies is a Couette device, where the

suspension fills the gap between concentric rotating cylinders (Fair & Gemmell 1964,

Ritchie 1965, Akers et al. 1987, Oles 1992, Shamlou 1993). Either the inner or outer,

or both, cylinders may rotate. In this case the laminar fluid shear rate can be calculated

directly from the relative rotation speed, the cylinder radii, and the width of the gap

(Mühle 1993, Krutzer et al. 1995). However, the Couette device may suffer end effects

(Ives 2000), although an open, upright design effectively removes one end. Both batch

(Ives 2000) and continuous (Farrow & Swift 1996) Couette reactors are reported, and

tapered shear rate can be generated using tapered cylinders (Ives & Bhole 1975, 1977)

with continuous flow.

Aggregation studies have also been performed in pipe flow, either laminar (Gregory

1981, Eisenlauer & Horn 1984, Whittington & George 1992, Suharyono & Hogg 1996),

or turbulent (Delichatsios & Probstein 1975, Klute & Amirtharajah 1991, Wigsten &

Stratton 1994, Ives 2000). However, particles tend to settle out of laminar flow unless

the pipe is vertical, or the particles are neutrally buoyant. Turbulent pipe flow has the

advantage of relatively homogeneous and isotropic turbulence in the core (Delichatsios

& Probstein 1975, Koh et al. 1984), and the mean shear rate can be estimated from the

pressure drop along the pipe (Thomas 1964, Matsuo & Unno 1981, Gould 1985,

Krutzer et al. 1995, Ives 2000), for example by manometer. The average residence time

is readily calculated from the flow rate and the pipe length/diameter, by assuming plug

flow (Nauman & Clark 1991 p. 127) in the turbulent regime. However, small scale

turbulent pipe flow generates a relatively high mean fluid shear rate (Suharyono &

Hogg 1996, Ives 2000, p. 152), and requires a large feed volume, which cannot be re-

cycled because it is irreversibly altered by the flocculant.

Page 110: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

109

As outlined before, the aim of this section is to provide aggregation kinetics data

suitable for a population balance model (see Section 5) of flocculation. In addition to

fluid shear, several other process variables may impact on the industrial flocculation

process, and the effects of flocculant dosage, solid volume fraction and primary particle

size are also been investigated.

4.1 Experimental Aggregation/breakage kinetics were studied in turbulent pipe flow (Figures 4-1, 4-2 &

4-3). A combination of pipe diameters (25.4 or 38.1 mm ID) and flow rates

(14-40 L min-1, 0.21-1.29 m s-1) were used to produce a range of fluid shear rates. A

series of flocculant injection nipples along the pipe produced a range of effective pipe

lengths, and hence mean residence times assuming plug flow.

The pipe reactors were assembled from lengths (1.83 m, BW Plastics) of transparent

acrylic pipe, with care taken to produce seamless joints and ensure a smooth pipe inner

surface. Conveniently, the pipes are produced in a range of sizes where the outside

diameter (OD) of each pipe is the same as the inside diameter (ID) of the next largest

size. This allowed a butt joint to be easily fabricated by gluing the ends together inside

a short length of larger pipe. Disassembly was by saw.

Similarly, the flocculant injection nipples were attached to the outside of the pipe so

that the inside surface of the pipe was only altered by only drilled (2.4 mm) holes. The

pipe was suspended approximately 1 m above the floor from a tensioned horizontal

suspension cable arrangement (Figure 4-1), ensuring that the pipe reactor was kept

straight and horizontal.

Page 111: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

110

Figure 4-1: Schematic of the laboratory equipment.

Figure 4-2: View of manometer bank and Figure 4-3: View of pipe reactor experimental apparatus

Flocculant dosing points

Sizing probe

Pipe reactor 15 m x 25.4 mm ID 20 m x 38.1 mm ID

Feed suspension storage tank

To waste

Settling column

Flocculant pump

Flocculant storage tank

Manometer

Feed pump

Linked ball valves

Page 112: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

111

A Lasentec FBRM (C-electronics) particle sizing probe was placed in the flow at the

end of the pipe reactor, followed by a vertical column used to take hindered settling rate

measurements. The FBRM instrument was described in Section 3, and was calibrated

with the same solid substrate (Calcite, Commercial Minerals).

The feed suspension was made up quantitatively in a 0.8 m3 stirred/baffled feed tank,

and confirmed gravimetrically by drying. A range of feed suspensions of different solid

fraction (3.33 – 16.7 % w/v, φExp = 0.012-0.062, ρcalcite = 2710 kg m-3) and/or mean

primary particle size (2.36 – 24.3 μm) were used to simulate changes to the feed

(Table 4-1). The various grades of calcite were sized independently by Laser

diffraction (Malven Mastersizer), which also provided particle surface area

measurements. Surface area measurements were also made by an alternative technique

(BET) for comparison.

The feed suspension was delivered to the pipe reactor by a variable ratio drive (metal

belt type) positive displacement pump (Mono-pump), and the flow rate confirmed by an

on-line (Admac AE) flow meter. The pressure drop along the pipe was measured by a

manometer bank (Figures 4-1 & 4-2). The manometers were equilibrated from above,

after flushing with water. This removed any solid from the manometer tubes, ensuring

a known density (taken as 1000 kg m-3) and accurate pressure drop measurement.

Flocculant (Nalco 9902, 30 % anionic acrylate/acrylamide copolymer with a nominal

molecular weight of 15 million g mol-1) stock solution (0.02 % w/v) was made up

reproducibly (stirrer intensity, volume and time) the day before use, minimising any

effect from flocculant aging (Owen et al. 2002). The flocculant was dosed

quantitatively on a mass (flocculant) for mass (solid) basis into the pipe reactor with a

peristaltic pump (Masterflex) and confirmed by flow meter (Magflow MAG 300).

Settling rate measurements were made on the aggregated suspension using an isolatable

(note valves) and detachable settling column, also made from acrylic tube (38.1 mm

ID). The column was marked at regular intervals (1 & 5 cm) and lit from behind by

fluorescent tube, allowing the measurement of the hindered settling rate by following

the fall of the mudline through time with a stopwatch.

Page 113: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

112

The pipe work and fittings were designed to keep the top of the settling column as close

as possible to the particle sizing probe, reducing the additional residence time. This

was the primary reason for mounting the pipe reactor above floor level (note Figure 4-

1). The fittings around the bend and valve were also constructed to produce a smooth

inner surface in an effort to reduce additional turbulence and pressure drop. Despite

this, the additional effective residence time was still considerable (see Section 6).

The addition of flocculant irreversibly alters the suspension, precluding the possibility

of re-cycle. Consequently, after size and settling rate measurement the flocculated

suspension was fed to a 10 m3 waste storage sump.

A matrix of experimental runs was performed according to Figure 4-4. The process

variables of fluid shear, flocculant dosage, solid fraction and primary particle size were

altered independently away from a common baseline point: 25.4 mm ID pipe,

14.01 L min.-1. (mean velocity 0.461 m s-1), 10 % w/v (3.69 % v/v) solid, Omya-carb 5

(volume-weighted average diameter 6.59 μm by Malvern), and a flocculant dosage of

20 g t-1 (solid).

Figure 4-4: Experimental matrix

Flocculant dosage (g t-1)

5

80

3.33

10

40

2.36

3.47

15.0

24.3

Mean primary particle size (μm)

Flow velocity (m s-1) [pipe ID (mm)]

6.67

16.67

0.554 [38.1]

0.207 [38.1]

Solid volume fraction (% w/v)

Baseline: 0.461 m s-1 [25.4 mm], 10 % w/v, 6.59 μm, 20 g t-1

0.343 [38.1]

1.294 [25.4]

0.781 [25.4]

13.33

Page 114: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

113

Aggregate size and settling rate measurements were made with flocculant injected at 13

positions (including a zero residence time - no flocculant) along the pipe, producing a

range of residence times. The injection points were spaced according to a numerical

progression, so that they were closer together at short residence times where the

aggregate size increases rapidly (e.g. Figure 4-5).

This range of pipe lengths was used for each of the experimental conditions described

by the sparse matrix described by Figure 4-4, giving a total of 13 × (6 + 4 + 4 + 4) data

points. A complete 13 × 5 × 4 × 4 × 4 experimental matrix was clearly not feasible.

However, data was also collected under additional conditions, in the gaps of the matrix,

by changing two process variables simultaneously. These additional data points were

used to test the predictivity of the population balance model described in Section 5.

Table 4-1: Matrix of pipe reactor experimental runs Run No. Pipe ID Flow rate Solid φ Mean dp Floc. dose Viscosity Susp. Density Kin. visc. Re ∈ G f Comment

(m) (m3s-1) (m3m-3) (m) (kg kg-1) (N s m-2) (kg m-3) (m2s-1) (m2s-3) (s-1)211100A 0.0254 2.34E-04 0.037 6.59E-06 2E-05 0.0024 1063 2.21E-06 5291 0.071 179.6 0.0093 Baseline # 1211100B 0.0254 2.34E-04 0.037 6.59E-06 4E-05 0.0165 1063 1.56E-05 752 0.116 86.4 0.0151 Settling211100C 0.0254 2.34E-04 0.037 6.59E-06 1E-05 0.0016 1063 1.52E-06 7681 0.065 206.6 0.0084221100A 0.0254 2.34E-04 0.037 6.59E-06 8E-05 0.0509 1063 4.79E-05 245 0.154 56.7 0.0200 Settling221100B 0.0254 2.34E-04 0.037 6.59E-06 5E-06 0.0014 1063 1.34E-06 8759 0.063 217.0 0.0082231100A 0.0254 2.34E-04 0.025 6.59E-06 2E-05 0.0018 1042 1.72E-06 6821 0.067 197.6 0.0087231100B 0.0254 2.34E-04 0.062 6.59E-06 2E-05 0.0030 1105 2.73E-06 4294 0.075 166.1 0.0098231100C 0.0254 2.34E-04 0.012 6.59E-06 2E-05 0.0015 1021 1.42E-06 8222 0.064 211.9 0.0083241100A 0.0254 2.34E-04 0.049 6.59E-06 2E-05 0.0027 1084 2.52E-06 4637 0.074 170.9 0.0096271100A 0.0254 2.34E-04 0.037 6.59E-06 2E-05 0.0026 1063 2.47E-06 4739 0.073 172.3 0.0095 Baseline # 2281100A 0.0254 3.96E-04 0.037 6.59E-06 2E-05 0.0015 1063 1.45E-06 13700 0.274 435.0 0.0073281100B 0.0254 6.56E-04 0.037 6.59E-06 2E-05 0.0012 1063 1.17E-06 28173 1.039 944.1 0.0061281100C 0.0381 6.32E-04 0.037 6.59E-06 2E-05 0.0019 1063 1.80E-06 11736 0.068 194.2 0.0076011200A 0.0381 3.91E-04 0.037 6.59E-06 2E-05 0.0330 1063 3.10E-05 421 0.037 34.5 0.0174 Settling011200B 0.0381 2.36E-04 0.037 6.59E-06 2E-05 0.1675 1063 1.58E-04 50 0.014 9.4 0.0297 Settling041200A 0.0254 2.34E-04 0.037 1.51E-05 2E-05 0.0058 1063 5.42E-06 2158 0.089 128.3 0.0116041200B 0.0254 2.34E-04 0.037 2.36E-06 2E-05 0.0019 1063 1.81E-06 6478 0.068 193.8 0.0088051200A 0.0254 2.34E-04 0.037 3.47E-06 2E-05 0.0016 1063 1.46E-06 7996 0.064 209.7 0.0084051200B 0.0254 2.34E-04 0.037 2.43E-05 2E-05 0.0101 1063 9.46E-06 1237 0.103 104.1 0.0133 Settling061200A 0.0254 2.34E-04 0.025 1.51E-05 2E-05 0.0028 1042 2.72E-06 4307 0.075 166.3 0.0098061200B 0.0254 3.96E-04 0.037 6.59E-06 4E-05 0.0012 1063 1.17E-06 16934 0.260 471.0 0.0069071200A 0.0381 3.91E-04 0.037 6.59E-06 1E-05 0.0029 1063 2.77E-06 4715 0.020 85.3 0.0095071200B 0.0381 3.91E-04 0.025 6.59E-06 2E-05 0.0136 1042 1.31E-05 1000 0.030 47.7 0.0140 Settllng

Page 115: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

114

4.2 Results and discussion 4.2.1 Effect of mean fluid shear rate

Figure 4-5 shows the effect of changing the flow rate and pipe reactor inside diameter

(ID) on the volume-average aggregate size. The flow rate and pipe ID were altered to

produce a range of spatially-averaged shear rates (~100-1000 s-1, Table 4-1) calculated

from the pressure drop measured by the manometer bank. A small pipe and high flow

rate increases the flow resistance, the energy dissipation, and hence also increases the

mean fluid shear rate. This initially leads to an increased aggregation rate due to faster

flocculant/suspension mixing and particle collision, but ultimately leads to a reduced

aggregate size due to an increased aggregate breakage rate. The reduction in final

aggregate size at higher shear rates is typical, and other systems show a similar response

(Curtis & Hocking 1970, Oles 1992, Keys & Hogg 1978, Mühle 1993, p. 382, Spicer &

Pratsinis 1996, Serra & Casamitjana 1998, Chin et al. 1998, Flesch et al. 1999).

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 10 20 30 40 50 60 70

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

38.1 mm ID, 0.343 m sֿ¹, G = 52.6 sֿ¹, (settling obs. in pipe)

25.4 mm ID, 0.461 m sֿ¹, G = 174.1 sֿ¹, (Baseline)

38.1 mm ID, 0.554 m sֿ¹, G = 175.2 sֿ¹

25.4 mm ID, 0.781 m sֿ¹, G = 488.1 sֿ¹

25.4 mm ID, 1.294 m sֿ¹, G = 935.5 sֿ¹

Figure 4-5: Effect of pipe size and fluid flow rate on mean aggregate size.

The gentle reduction in the aggregate size at extended residence times is typical

behaviour when polymer flocculants are used (Keys & Hogg 1978, Williams et al.

1992, Leu & Ghosh 1988, Bagster 1993, Hogg 2000), and is usually taken as evidence

of flocculant degradation by scission or rearrangement (Sikora & Stratton 1981,

Gregory 1993, Stratton 1983, Ditter et al. 1982). Conversely aggregates formed by

coagulation with soluble salt typically reach a steady state aggregate size, because in

Page 116: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

115

that case breakage is reversible (Oles 1992, Flesch et al. 1999, Spicer & Pratsinis

1996a, Serra & Casamitjana 1998).

A manometer bank (Figures 4-1 & 4-2) was used to measure the pressure drop along the

pipe reactor. This allowed the calculation of the energy dissipation rate per unit mass of

fluid (∈) from (Perry & Green 1997):

LPV

fρΔ

∈= 4-2

where: V = Mean flow velocity along pipe (m s-1)

△P = pressure drop along pipe (N m-3) given experimentally from: ghP ρ=Δ ρf = Density of the fluid (kg m-3) L = Pipe length (m)

The mean pipe flow velocity (V) is known experimentally via the volumetric flow rate

and pipe ID. The fluid density (ρf) is readily calculated from the solid fraction and

density:

( )φ−ρ+φρ=ρ 1wsf 4-3

where: ρ = Density (f = fluid, s = solid, w = water) (kg m-3) φ = Solid volume fraction [0,1]

In this case the solid volume fraction (φ) is known experimentally because the feed is

made up quantitatively in batches in the feed tank, and confirmed gravimetrically by

drying.

The mean dissipation rate is readily calculated from Equation 4-2, however in order to

calculate the mean fluid shear rate (Equation 4-1), the viscosity is also required. In

most studies (Williams & Simons 1992) the viscosity is taken to be unchanged from

pure water at the same temperature. This assumption is usually valid because most

studies (Oles 1992, Spicer et al. 1996, Serra & Casamitjana 1998, Flesch et al. 1999,

Manning & Dyer 1999) have simulated conditions in water treatment clarifiers or river

estuary system, which are characterised by a low solid fraction (Dahlstrom & Fitch

1985, Perry & Green 1997).

However, feed streams to thickener units used in mineral processing have a

considerably higher solid volume fraction, typically in the percent range (Pearse 1977,

Dahlstrom & Fitch 1985, Perry & Green 1997), leading to an increased fluid viscosity

Page 117: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

116

(Williams et al. 1992, see Section 5). The volume of liquid enclosed within porous

aggregates has the effect of further increasing the effective solid volume fraction (Mills

et al. 1991, see Section 6). This effect is demonstrated experimentally (Figure 4-7) by

an increased flow resistance and larger pressure drop along the pipe. The use of a

positive displacement feed and flocculant pumps ensured constant flow rates, with the

additional required energy readily supplied by the pumps since the overall backpressure

along the pipe was comparatively minor at ∼ 1-10 × 103 N m-2 (0.01- 0.1 atm.).

Estimating the pressure drop through pipes is a common engineering problem, e.g.

when specifying pumps for pipelines. This has lead to the development of the concept

of a pipe Fanning friction factor (f), given by (Bird et al. 1960, Govier & Aziz 1972,

Wasp et al. 1977, Perry & Green 1997):

32 V2D

LV2PDf ∈

=ρΔ

= 4-4

where: f = Fanning friction factor D = Pipe diameter (m) ΔP = Pressure drop along pipe (N m-3)

ρ = Fluid density (kg m-3) ∈ = Energy dissipation rate (J s-1kg-1, m2s-3) V = Mean flow velocity (m s-1) L = Pipe length (m)

An alternative system is based on the Darcy friction factor (Daugherty et al. 1989),

differing only by a factor of 4.

Equation 4-4 is usually under-specified (ΔP or ∈), and efforts have been made

(Table 4-2) to find alternative descriptions of the friction factor, usually as a function of

the pipe roughness and pipe Reynolds number (Re), given by:

μρ

=DVRe 4-5

For turbulent flow (4000 < Re < 100 000) in smooth pipes the friction factor is described by the

Blasius (1913) equation (Table 4-2):

41

Re

0791.0f = 4-6

Page 118: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

117

Table 4-2: Equations relating pipe friction factor to Reynolds number Reference Equation Comment Govier & Aziz 1972 LV2

PgDf 2ρΔ

≡ Definition of friction factor (similar to Darcy’s friction factor)

Blasius 1913

41

Re

079.0f = 4 000<Re<100 000, smooth pipes (1)

Colebrook 1939 (see below)

⎥⎦

⎤⎢⎣

⎡+

∈−=

fRe256.1

D7.3log4

f1

Re>4 000, rough pipes, (1)

Churchill 1977

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠⎞

⎜⎝⎛+

∈−=

9.0

Re7

D27.0log4

f1

Re>4 000, rough or smooth pipes, ∈ = surface roughness (mm), (1)

Knudsen & Katz 1958 2.0Re

046.0f = (2)

Nikuradse 1932

[ ] 4.0fRelog4f

1−=

3 000<Re<3 000 000, (2)

Drew, Koo & McAdams 1932 32.0Re

125.000140.0f +=

3 000<Re<3 000 000, (2)

Govier & Aziz 1972

[ ] 6.0fRelog06.4f

1−=

(2)

Nikuradse 1932

48.3k2

Dlog4f

1+⎥⎦⎤

⎢⎣⎡=

Rough pipes (2)

Von Karman 1930 36.3

k2Dlog06.4

f1

+⎥⎦⎤

⎢⎣⎡=

Rough pipes (2)

Colebrook 1939 (see above)

⎟⎟⎠

⎞⎜⎜⎝

⎛+−+⎥⎦

⎤⎢⎣⎡=

fRek2D35.91log448.3

k2Dlog4

f1

Partially rough pipe (2)

Buckingham 1921 ⎟

⎟⎠

⎞⎜⎜⎝

+−ρ= 3

31

34 xx1

1DV16f where:

PDL4

x y

w

y

Δ

τ=

τ

τ=

Non-Newtonian fluid (2)

Govier & Winning 1948

⎟⎟⎠

⎞⎜⎜⎝

⎛μ

τ

μρ

φ=V

gD,DVf cy

1 Non-Newtonian fluid (2)

Hedstrom 1952

⎟⎟⎠

⎞⎜⎜⎝

μ

ρτ

μρ

φ= 2cy

2

2gD

,DVf Non-Newtonian fluid (2)

Tarrance 1963 n

65.20.6kRlog

n07.4

f1

−+⎥⎦⎤

⎢⎣⎡=

Non-Newtonian fluid (2)

Clapp 1961 [ ] 3.2fRelog53.4

f1

−= Non-Newtonian fluid (2)

(1) Perry & Green 1997 (2) Govier & Aziz 1972

Page 119: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

118

Equation 4-6 was used to calculate the pressure drop along the pipe reactor with water

only (no solid), allowing the estimation of the fluid viscosity (0.00105 N s m-3, compared

to 0.00102 N s m-3 at 20 oC, Daugherty 1989) from the experimental pressure drop.

y = x - 1.0461R2 = 1

0

100

200

300

400

500

600

700

800

900

0 100 200 300 400 500 600 700 800 900

Measured pressure drop (N m-3)

Est

. pre

ss. d

rop

Bla

sius

eq.

(N m

-3)

Figure 4-6: Comparison of measured and predicted (Equation 4-7) pressure drop of water only under various flow regimes (as per Figure 4-5).

Equation 4-6 is tentatively assumed to hold for the turbulent flow of the aggregated

suspensions, allowing the calculation of the fluid viscosity by combining Equations 4-4,

4-5 & 4-6:

4734

54

LV)0791.0*2(DPρ

Δ=μ 4-7

where: μ = Suspension viscosity (N s m-2) D = Pipe diameter (m)

ΔP = Pressure drop along pipe (N m-3) ρ = Fluid density (kg m-3) ∈ = Energy dissipation rate (J s-1kg-1, m2s-3) V = Mean flow velocity (m s-1) L = Pipe length (m)

Figure 4-7 shows the increase in the average suspension viscosity flocculated at

different shear rates, as calculated from the pressure drop using Equation 4-7 (see also

Table 4-1). The viscosity increases at lower shear rates, because the increased

aggregate size (Figure 4-5) leads to increased aggregate porosity (see Sections 5 & 6)

and hence effective solid fraction and suspension viscosity. The data is truncated to

Page 120: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

119

lower shear rates because stratification/settling in the pipe reactor throttled the flow,

leading to an increased pressure drop.

0

0.0005

0.001

0.0015

0.002

0.0025

0.003

0 100 200 300 400 500 600 700 800 900 1000Mean shear rate (s-1)

Vis

cosi

ty (N

s m

-2)

Figure 4-7: Aggregated suspension fluid viscosity estimated (Equation 4-7) for the pressure drop along the pipe (see Figure 4-5 for aggregate size data).

Solid particles are suspended by turbulent eddies, however, if the turbulence is

insufficient for a given particle settling rate, the solid may stratify, or form a stationary

bed on the bottom of the pipe reactor (Figure 4-8, Table 4-3). This has the effect of

throttling the flow, increasing the flow velocity and turbulence in the remaining clear

portion of the pipe, preventing further settling. Eventually a steady state condition is

reached with a uniform stationary layer. The use of a transparent acrylic pipe allows the

observation of stratification/settling, especially if a fluorescent tube is mounted behind

the pipe to cast a shadow. Data collected when there was visible stratification/settling is

of limited value, because the settled solid reduced the effective pipe diameter and

decreased the average residence time.

Minimum suspension flow velocities in pipelines are usually described by the Durand

(1953) equation as a function of the mean flow velocity in the pipe, and the particle size

and density, although there is a variety of alternative relationships to choose from

(Spells 1955, Zandi & Govatos 1967, Kao & Wood 1974 , Turian & Yuan 1977, Wasp

et al. 1977, Oroskar & Turian 1980, Anon. 1989, Shah & Lord 1991, Table 4-3).

Page 121: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

120

( )[ ] 5.0L2M 1sgD2FV −= 4-8

where: VM2 = Minimum velocity required to keep solid suspended (m s-1) (Figure 4-8) d = Particle diameter (m) D = Pipe diameter (m)

g = Gravity (9.8 m s-2) s = ratio of the solid to liquid density FL = function of solid volume fraction and particle size

Equation 4-8 is typically used for estimating the minimum transport velocity of

relatively coarse particles (>1000 μm) in full-scale pipelines, and has limited

applicability to the current work with aggregated suspensions of fine particles. In the

current case, stratification/settling was observed around the transition to laminar flow

(Re ∼ 2000-4000, see Table 4-1).

Figure 4-8: Critical flow velocities, see Table 4-3 below (adapted from Perry and Green 1997).

Log suspension velocity

Log

Pre

ssur

e gr

adie

nt VM3

VM1 VM4

MixturePure liquid

Sym

met

ric

Susp

ensi

on

Sym

met

ric

Susp

ensi

on

Stat

iona

ry B

ed

Mov

ing

Bed

VM2

Page 122: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

121

Table 4-3. Equations describing critical minimum flow velocities of slurries Reference Equation Comments Durand 1953

( )[ ] 5.0

L2M 1sgD2FV −=

Widely referenced

Spells 1955 )1s(gd

DV075.0V 85

775.0M1M2

1M −⎟⎟⎠

⎞⎜⎜⎝

⎛μρ

=

Spells 1955 )1s(gd

DV025.0V 85

775.0M1M2

2M −⎟⎟⎠

⎞⎜⎜⎝

⎛μρ

=

Oroskar and Turian 1980

( )[ ] 30.009.0Re

378.050

3564.0v

1536.0v

5.0502M N)D/d()C1(C1sgd85.1V χ−−= − [ ]

f)1S(gd2fD

N21

50Re μ

−ρ=

Wasp et al. 1977 ( )( )

31

21

dDF1sgd2V L2M ⎟⎠⎞

⎜⎝⎛−=

Zandi and Govatos 1967

( )( )21

21

dD

CC401sgdVd

2M ⎟⎠⎞

⎜⎝⎛⎟⎟⎠

⎞⎜⎜⎝

⎛−=

Shah and Lord 1991

( )[ ] zRe

w50

3564.0v

1536.0v

5.0502M N)D/d()C1(C1sgdYV −−−=

Turian and Yuan 1977

( )( )21

3225.0065.1w

5906.d2M CfC1sgD4608.0V −−−−=

Kao and Wood 1974 ( )( )

21

n1

21

dD

Dd

381sgdV 22M ⎟

⎠⎞

⎜⎝⎛

⎟⎠⎞

⎜⎝⎛⎟⎟⎠

⎞⎜⎜⎝

αβ−=

D = Pipe diameter (m) d = Particle diameter (m) ρ = Density (kg m-3) s = Ratio of the solid to liquid density Cv = Particle drag coefficient FL = Function of solid volume fraction and particle size The relatively high shear rate produced by small-scale turbulent pipe flow

(approx 100-1000 s-1 here) acts as an incentive to reduce the flow velocity and increase

the pipe size. However, reducing the flow reduces the Reynolds number, causing the

solid to settle out, while increasing the pipe size dramatically increases the required

feed and disposal volumes (~ 5 tonnes in this case). Alternatively, settling can be

suppressed by substituting a smaller, or less dense primary particle to reduce the settling

rate. In this case a range of relatively fine primary particles were used

(mean dp = 2.36 - 24.3 μm). Using a lower density solid substrate (latex perhaps) would

have compromised the hindered settling rate measurements (Section 6), and may also

have given a poor approximation of a mineral surface.

Page 123: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

122

4.2.2 Effect of flocculant dosage

Figure 4-9 shows the increase in aggregate size with flocculant dosage. The increase in

size can be attributed to either an increased capture efficiency (α) and/or increased

aggregate strength leading to reduced breakage. Although the data show a faster initial

aggregation rate at higher dosages, the difference must be at least partially attributed to

the increased initial mixing at higher dosages. The dilute (0.02 %) flocculant stream is

injected perpendicularly into the pipe reactor via a small hole in the pipe wall

(see Section 4.1), causing a visible (acrylic pipe) increase in the turbulence around the

injection point. The additional initial turbulence increased with flocculant dosage,

because the dosage was altered by changing the flocculant stream flow rate (rather than

by increasing the flocculant stream concentration).

Flocculant/suspension macro-scale mixing is likely to be highly scale dependent, and its

importance is evidenced by the industrial practice of using a very dilute flocculant

stream (0.01-0.1 %, Perry & Green 1997, Dahlstrom & Fitch 1985) and various

proprietary feedwell designs (e.g. multi-point addition) to produce efficient mixing. It

is difficult to replicate industrial-scale mixing conditions on a laboratory-scale, however

the use of emerging techniques like computational fluid dynamics (Lainé et al. 1999,

Ducoste & Clark 1999, Farrow et al. 2000), or electrical impedance tomography

(Brown et al. 1985, Webster 1990, Salkeld 1991, Dickin et al. 1992, Williams &

Simons 1992) is likely to increase the understanding in this area.

Figure 4-9 shows an increased aggregate size with flocculant dosage. There is evidence

to suggest (Healy 1961, Fleer & Scheutjens 1993, Svarovsky 2000) that very high

flocculant dosages may lead to steric re-stabilisation as the particle becomes overloaded

with flocculant. However, it is not clear that such a situation would occur

commercially, where flocculant is the major operating cost for many units (Perry &

Green 1997, Hogg 2000), and excessive flocculant may reduce the unit’s compression

performance (Healy et al. 1994, Svarovsky 2000).

Page 124: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

123

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

2.5E-04

3.0E-04

3.5E-04

0 5 10 15 20 25 30 35

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Dosage = 80 g tֿ¹, (settling obs. in pipe)Dosage = 40 g tֿ¹, (settling obs. in pipe)Dosage = 20 g tֿ¹, (Baseline)Dosage = 10 g tֿ¹Dosage = 5 g tֿ¹

Figure 4-9: Effect of flocculant dosage on the aggregate size.

Figure 4-10 shows the increase in viscosity and corresponding reduction in fluid shear

as a function of the flocculant dosage. Again, this is attributed to the increased effective

solid volume fraction as the aggregate size is increased. The possibility of the

flocculant alone increasing the viscosity was discounted by performing an additional

experiment performed under the same conditions, except without solid (water and

flocculant only). Somewhat surprisingly, the pressure drop actually decreased

fractionally (not shown), a phenomenon widely reported (Paterson and Abernathy 1970,

Moussa and Tiu 1994, Rho et al. 1996, Perry & Green 1997) and exploited to reduce

the pumping costs in long pipelines.

0

50

100

150

200

250

0 5 10 15 20 25Flocculant dosage (g t-1)

Shea

r ra

te (s

-1)

0

0.001

0.002

0.003

0.004

0.005

Vis

cosi

ty (N

s m

-2)

Shear rateViscosity

Figure 4-10: Effect of flocculant dosage on the measured fluid viscosity and shear rate. Data is truncated due to settling in pipe reactor.

Page 125: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

124

4.2.3 Effect of primary particle size

Figure 4-11 shows the increase in aggregate size with primary particle size. This may

be caused by a combination of effects. As the primary particle size is increased, the

total particle surface area per unit mass is decreased, leading to an increased flocculant

surface coverage per area (Hogg 1999a). The increase in primary particle size will also

increase the solid packing efficiency within the aggregate, decreasing the porosity,

hence increasing the aggregate strength (Dobias 1993, Bagster 1993). The decreased

porosity also leads to a decrease in the effective suspension solid fraction (Equation 6-

9), leading to a decreased fluid viscosity, dissipation rate and breakage.

Overall however, the increased strength of aggregates formed from larger primary

particles leads to an increased overall degree of aggregation, a higher suspension

viscosity (Figure 4-12) and dissipation rate. In this case the viscosity is the dominant

term of Equation 4-1, so the fluid shear is reduced with a larger primary particle.

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

2.5E-04

0 5 10 15 20 25 30 35

Time (s)

Mea

n A

ggre

gate

size

(m)

dp = 24.3 µm, (settling obs. in pipe)dp = 15.0 µmdp = 5.59 µm, (Baseline)dp = 3.47 µmdp = 2.36 µm

Figure 4-11: Effect of mean primary particle size on the measured fluid viscosity.

Page 126: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

125

0

50

100

150

200

250

0 5 10 15 20 25 30

Mean primary particle diameter (µm)

Shea

r ra

te (s

-1)

0

0.002

0.004

0.006

0.008

0.01

0.012

Vis

cosi

ty (N

s m

-2)

Shear rateViscosity

Figure 4-12: Effect of primary particle size on the measured fluid viscosity and shear rate. The mean size is the volume weighted mean as determined by laser diffraction (Malvern Mastersizer).

4.2.4 Effect of suspension solid fraction

Figure 4-13 shows the reduction in aggregate size with increased solid fraction, a result

that has been observed elsewhere (Lick & Lick 1988, Williams et al. 1992), although

not well understood. Figure 4-14 shows the increase in suspension viscosity and

reduction in the shear rate as determined by manometer measurements. The reduction

in the aggregate size at high solid fraction is attributed to a reduced particle collision

rate at lower shear, and an increased breakage rate due to the increase in viscosity and

energy dissipation. At very low solid fraction, there is evidence (Thomas 1964, de Boer

et al. 1989, Kobayashi et al. 1999) to suggest that the aggregate size may be reduced,

perhaps due to a decreased collision rate (see Section 5).

Page 127: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

126

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35

Time (s)

Mea

n A

ggre

gate

size

(m)

3.33 % w/v6.67 % w/v10.0 % w/v (Baseline)13.33 % w/v16.67 % w/v

Figure 4-13: Effect of suspension solid fraction on mean aggregate size. The flocculant dosing rate was maintained at 20 g (flocculant)/t (solid) as per the baseline condition.

0

50

100

150

200

250

300

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07

Absolute solid fraction (v/v)

Shea

r ra

te (s

-1)

0

0.001

0.002

0.003

0.004

Vis

cosi

ty (N

s m

-2)

Shear rateViscosity

Figure 4-14: Effect of feed suspension solid fraction on the measured fluid viscosity.

Page 128: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

127

4.3 Conclusions The aggregation/breakage kinetics of calcite flocculation have been studied

experimentally in turbulent pipe flow under a range of conditions, reflecting changes in

likely process variables: fluid shear, flocculant dosage, primary particle size, and solid

volume fraction. Reaction residence times were altered by varying the length of pipe

between the flocculant injection point and the aggregate sizing probe. The pressure

drop along the pipe was measured by manometer, allowing the calculation of the energy

dissipation, fluid viscosity, and the mean shear rate.

Efforts were taken to match the conditions in the pipe reactor to those expected in the

full-scale plant, although scale-up remains an issue. Flocculant/feed suspension mixing

is faster with small volumes, and it is likely to be considerably overestimated by

small-scale equipment. However, the major limitation is due to the tendency of solid to

settle out of small-scale low Reynolds number flows, restricting this study to a smaller

primary particle and higher mean shear rate than are expected in practice.

The initial aggregation rate was found to increase at higher shear rates, however the

final aggregate size was ultimately limited by aggregate breakage which is significantly

increased at a higher shear rate. The aggregate size is increased as a function of the

flocculant dosage and primary particle size, but reduced at higher solid fraction.

Page 129: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

128

5. Population balance model

5.1 Model development 5.1.1 Aggregation

The experimental data presented in Section 4 was used to develop a population balance

model for aggregation and breakage of calcite flocculated by high molecular weight

polymer flocculant. The population balance is based on the successful population

balance described by Hounslow et al. (1988) and Spicer and Pratsinis (1996a) (see also

Equation 2-45):

∑∑∑∞

=

=

=−−−

=−−

+−

Γ+−

β−β−β+β= −

ijjjj,iii

ijjj,ii

1i

1jjj,ii

21i1i,1i

2i

1j21

j1ij,1i1iji

NSNS

NNN2NNNN2dt

dN ij

5-1

where: Ni = Number of ith sized particles (m-3) t = Time (s) βij = Rate of aggregation of i and j sized particles (aggregation kernel) (m3s-1) Sj = Breakage rate (kernel) of jth sized particles (s-1) Γij = Breakage distribution function (number of i size particles produced from the breakage of a j sized particle) The widely accepted (Tables 2-1 & 2-2) turbulent collision kernel by Saffman and

Turner (1956) was used to describe the aggregation rate:

( )3jiij aaG294.1 +α=β 5-2

where: α = Capture efficiency [0,1] (taken as = M)

G = Average turbulent shear rate (s-1) ai = Radius of the ith particle (m)

Before flocculant addition, the primary particles are taken to be well dispersed as a

stable suspension, although in many systems there the primary particles may naturally

coagulate to some extent due to a low zeta potential. However, no large stable

aggregates are formed at this stage, despite their collision in the turbulent flow. In this

situation the particles are repelled by electrostatic repulsion (Gregory 1989, Kohler

1993, Hughes 2000) and the effective capture efficiency (α) is taken to be zero.

Alternatively, every collision could be considered to be successful (α = 1), but with the

breakage rate equal to the collision rate. Both methods were tried during model

Page 130: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

129

construction, but in this case the capture efficiency was taken to be zero before

flocculant addition, giving a simpler and more stable model, and a better fit with the

experimental data.

On the addition of flocculant, the aggregation rate exceeds the breakage rate, producing

the increase in the aggregate size shown in Figure 5-1. In this case the capture

efficiency must be greater than zero, however, if the capture efficiency is set to unity

immediately on flocculant addition, Equation 5-2 considerably overestimates the

aggregation rate compared to the experimental data. This is a common finding

(Table 2-2) and the capture efficiency is frequently taken to be in the range 0.05-0.5. In

this case, if the capture efficiency was set to a constant value after flocculant addition, it

had to be relatively small (~ 0.06) to give a reasonable fit to the data, suggesting that

particle collision (Equation 5-2) was not the rate limiting step. Alternatively, the

capture efficiency was taken to increase rapidly after flocculant addition

(Equations 5-3 - 5-6), reflecting the mixing/adsorption time required for the polymer to

reach the particle surface.

The mixing and adsorption behaviour of polymer flocculants is complex and poorly

described, although there is general agreement that adsorption is rapid under normal

industrial conditions (Gregory 1993, Bagster 1993, Mühle 1993, Hogg 1999a, 1999b),

and that the bulk mixing of the flocculant and feed streams is rate limiting. This effect

is likely to be highly scale-dependent, but its importance is evidenced by the complexity

of commercial feedwells, containing various design elements (Dahlstrom & Fitch 1985,

Perry & Green 1997, Svarovsky 2000) to produce efficient flocculant mixing.

Mixing is usually described in terms of a mixing index (Nagata 1975, Etchells & Short

1988, Godfrey & Amirtharajah 1991):

o

1Mσσ

−= 5-3

and assuming: α = M where:

σ = Concentration variation (σo at t = 0), given by:

Page 131: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

130

( )∫ −=L

0

22 dx)m(C)x(Cσ 5-4

and: M = Mixing index [0,1]

C(x) = Concentration at point x (or m = mean). However, experimental data for Equation 5-4 is generally unavailable, and alternatively

Etchells and Short (1988) suggest for turbulent pipe flow:

D/Lfk

o

1e−=σσ

5-5

where: k1 = Constant, Etchells and Short (1988) suggest = 0.5 (Fitted parameter # 1 := 0.3431) f = Pipe friction factor (Equation 4-4) L = Pipe length (m) D = Pipe diameter (m)

Combining Equations 5-3 and 5-5 gives:

D/Lfk1e1M −−= 5-6 Equation 5-6 rapidly approaches unity (by ~ 100 pipe diameters; Figure 5-2)

independent of the flocculant dosage, although this is unlikely in practice and

inconsistent with Equations 2-17 and 2-19. However, in this case the population

balance was made dosage dependent via the breakage kernel (Section 5.1.2,

Equation 5-16) reflecting an increase in aggregate strength with flocculant dosage

(Smith & Kitchener 1978, Ray & Hogg 1987). Both approaches were tried during

model construction, but making the breakage kernel (rather than the capture efficiency)

dosage dependent gave a better fit with the experimental data.

Equation 5-6 is a reasonably crude approximation of the overall mixing/adsorption

process and further work is required to better describe these effects, perhaps by a

combination of CFD and electrical impedance tomography. When the flocculant

mixing/adsorption process is better understood it should be possible to further refine the

capture efficiency term by including the effects of flocculant surface coverage and/or

the hydrodynamic effect.

Page 132: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

131

It would have been preferable to begin the experimental and model with the flocculant

already well mixed, however this is not possible with polymer flocculants. Studies of

aggregation using salt coagulants can be performed (Spicer & Pratsinis 1996a) by

adding coagulant to a stirred tank with a high impeller speed to mix the coagulant and

disperse the primary particles. The stirrer may then be turned down to give the

appropriate shear rate, and aggregation allowed to begin. However, this approach is not

possible with fragile polymer flocculants, and any initial rapid stirring results in

irreversible polymer degradation.

Aggregation kernels for turbulent flow (Tables 2-1 & 2-2) are usually written in terms

of the mean shear rate (G):

21

fG ⎟⎟⎠

⎞⎜⎜⎝

⎛μρ∈

= 5-7

and the dissipation rate (∈) is given by Equation 4-4, which rearranges to:

DfV2 3

∈= 5-8

and, for turbulent pipe flow in smooth pipes, the friction factor (f) is given by the Blasius equation:

41

Re

0791.0f = 5-9

where the pipe Reynolds number is:

μρ

= fDVRe 5-10

The slurry density (ρf) is readily calculated from experimental data given the solid

fraction (φ) and the solid density (ρs):

( )φ−ρ+φρ=ρ 1wsf 5-11 leaving the viscosity (μ) the only unknown in Equation 5-7. The increase in fluid

viscosity with solid fraction was described by Einstein (1908), using the first terms of a

series expansion in the form:

Page 133: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

132

( )...K1 Eos +φ+μ=μ 5-12 where: μs = Viscosity of suspension (N s m-2) μo = Fluid viscosity (no solid) (N s m-2) KE = Constant = 2.5 φ = Solid fraction [0,1]

Einstein originally proposed that KE = 1, but later, famously, revised it to KE = 2.5.

Equation 5-12 is only correct for very dilute solutions, and various alternative equations

(Table 2-5) have been proposed for higher solid fractions by adding additional terms.

For suspensions of high solid fraction the viscosity is typically given as a function of

the maximum solid fraction in the form (Govier & Aziz 1972, Liu & Masliyah 1996):

k

mos 1

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

−μ=μ 5-13

where: φm = Maximum solid fraction. k ≈ 2

μs = Suspension viscosity (N s m-2) μo = Viscosity of water (1.02 × 10-3 N s m-2 at 20 oC)

The maximum solid volume fraction is typically taken as 0.6 - 0.7 (Fleer & Scheutjens

1993, Strenge 1993, Schramm 1996, Bustos et al. 1999). At this point the particles are

taken to be in close contact and form a continuous network that resists shear, causing

the viscosity to rise exponentially.

The effective solid fraction will be higher than the actual solid fraction because the

aggregates are porous, incorporating fluid into the structure and increasing the enclosed

volume, and hence the suspension viscosity (Mills et al. 1991, Potanin & Uriev 1991).

The effective volume fraction (φeff) is described by fractal geometry (Mills et al. 1991,

Jiang & Logan 1991, Oles 1992, Kusters et al. 1997, Flesch et al. 1999) (see

Appendix):

⎟⎟⎠

⎞⎜⎜⎝

ρρ

φ=φagg

sseff 5-14

where: φs = Actual suspension solid volume fraction [0,1]

Page 134: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

133

ρs = Density of the solid (calcite = 2710 kg m-3) ρagg = Effective aggregate density (kg m-3), given by:

3D

p

aggsagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ=ρ 5-15

where: dagg = Diameter of aggregate (m) dp = Primary particle diameter (m) Df = Mass-diameter fractal dimension (see Appendix) i.e. substituting and using mean sizes gives:

kD3

p

agg

m

effos

f

d

d1μμ

−−

⎟⎟⎟

⎜⎜⎜

⎟⎟

⎜⎜

φφ

−= 5-16

In this case the maximum packing fraction (φm = 0.65) and exponent (k = 2) are

effectively model parameters estimated from the literature. During model construction

φm was initially treated as a fitted parameter [0.6,0.7], however the model was

surprisingly insensitive to φm over that range, and was subsequently fixed at an

intermediate value of 0.65, removing a degree of freedom. The fractal dimension (2.4)

was estimated from the hindered settling data (see Section 6).

5.1.2 Aggregate breakage

Aggregate breakage is included by the final two terms in the population balance

(Equation 5-1). In this case breakage is taken to be binary, producing two equal sized

daughter fragments, a common modelling approach (Randolph & Larson 1988, Chen

et al. 1990, Chatzi & Kiparissides 1995, Spicer & Pratsinis 1996, Flesch et al. 1999)

although more fragments covering a range of sizes (Valentas & Amundson 1966,

Valentas et al. 1966, Coulaloglou & Tavlarides 1977, Lu and Spielman 1985, Cheng

and Redner 1988, Lick and Lick 1988, Burban et al. 1989, Spicer & Pratsinis 1996a,

Ducoste & Clark 1998, Serra & Casamitjana 1998b) have also been used. In the case of

the population balance used here, the method of discretisation is relatively coarse

(aggregate mass doubles every channel, Equation 5-22), in an effort to reduce the

number of size fractions and equations. Because of this, most of the particles in a

distributed breakage function (Spicer & Pratsinis 1996a) will naturally appear in the

next smallest size range, unless the distribution is very broad or the number of

Page 135: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

134

fragments is assumed to be ≫2. In the case of a broad distribution, problems may arise

unless the distribution is truncated so that the daughter fragments must be smaller than

the parent aggregate.

Although there is general agreement that Saffman and Turner’s aggregation kernel is

appropriate (provided that the particle is spherical, neutrally buoyant, and is smaller

than the Kolmogorov micro-scale; Tables 2-1 & 2-2), there is less agreement on the

form of the breakage kernel. A typical modelling approach is to use Saffman and

Turner’s aggregation kernel in conjunction with a capture efficiency term, and a

breakage kernel that gives an acceptable fit with the experimental data.

The breakage kernel is usually taken to be a function of the aggregate size and the mean

energy dissipation or shear rate. Most of the theoretical work on aggregate breakage

(Parker et al. 1972, Shamlou & Titchener-Hooker 1993) has assumed that the

aggregates will grow up to some limiting size, when the aggregate will break. Although

this approach is attractive in its mathematical simplicity, aggregate distributions

typically cover a range of sizes (often described by some distribution like log-normal)

and not a distinct upper size limit.

For population balance models the breakage rate is usually (Glasgow & Lueche 1980,

Leu & Ghosh 1988, Spicer & Pratsinis 1996a, Serra & Casamitjana 1998b) taken to be

some continuous function of the aggregate size, reflecting the distribution of aggregate

strengths, orientations and the instantaneous fluid fluctuations in turbulent flow. Most

workers (Table 2-6) have found that the breakage rate is directly proportional to the

aggregate size, although higher exponents are also suggested (Glasgow & Lueche 1980,

Shamlou & Titchener-Hooker 1993). Pandya and Spielman (1982) found

experimentally that aggregate breakage was directly proportional to the aggregate size.

The breakage rate is usually also taken to be a function of the energy dissipation or

shear rate (Table 2-6). A higher shear rate typically provides rapid initial mixing and

fast initial aggregation (Figure 5-1), but eventually results in a smaller aggregate size

due to increased breakage. This is a typical experimental result (Curtis & Hocking

1970, Keys & Hogg 1978, Oles 1992, Chin et al. 1998) and is usually incorporated into

the population balance by assuming:

Page 136: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

135

x

i GS ∝ 5-17 where x is a higher exponent than in the aggregation kernel. That is, an increased shear

rate increases both the aggregation and breakage rates, but the breakage term is

dominant. Aggregation is usually taken as being directly proportional to G (e.g.

Equation 5-2) and x is typically reported to be 1.5-2 (Spicer & Pratsinis 1996a, Serra &

Casamitjana 1998b, Chung et al. 1998).

Equation 5-17 suggests (note Equation 5-7) that the breakage rate is decreased as a

function of viscosity, although some workers have suggested the more intuitive

response that the breakage will increase with the viscosity (Glasgow & Lueche 1980,

Kramer & Clark 1999). In most cases the form of Equation 5-17 has been successful,

because most studies have focused on coagulation in water treatment plants or river

estuaries which are characterised by a low solid fraction where the viscosity is

essentially unchanged from pure water (Williams & Simons 1992, Stumm & Morgan

1996, Spicer and Pratsinis 1996a, Serra & Casamitjana 1998b, Flesch et al. 1999,

Manning & Dyer 1999).

However, feed streams to thickener units used in mineral processing have a

considerably higher solid volume fraction, typically in the percent range (Pearse 1977,

Dahlstrom & Fitch 1985, Perry & Green 1997). In this case the breakage rate was

initially considered to be ∝ ∈yμz where μ, the viscosity, is given by Equation 5-10.

During model construction, parameter estimation repeatedly gave y = 0.7 ± 0.1, and

z = 1 ± 0.05. The value of ~ 0.7 is equivalent to x ~ 1.4 in Equation 5-17 (note

Equation 5-7) similar to other studies, and z was set to unity, removing a degree of

freedom. The final estimate of y was 0.677, perhaps suggesting Si ∝ ∈2/3.

Flocculant dosage dependency was also introduced into the breakage kernel to account

for the formation of larger aggregates at higher dosages (Figure 5-11). The dosage was

expressed as a surface coverage (θf) on a mass/area basis calculated from the

experimental addition rate and measured particle surface area (see Section 4):

Page 137: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

136

( )Θ−=θ 1MAm

s

ff 5-18

and the breakage kernel as:

paggf

i,aggk

2i dd

dkS

3

μ∈= 5-19A

pagg dd 0 ≤= 5-19B

where: θf = Effective flocculant surface coverage (kg m-2) k2 = Fitted parameter # 2 := 38.1 k3 = Fitted parameter # 3 := 0.677 M = Mixing index given by Equation 5-6 mf = Mass of flocculant (kg) As = Surface area of solid (m2) Θ = Flocculant degradation [0,1](Equation 5-20)

Studies of coagulation typically show an initial increase in the aggregate size after

coagulant addition, followed by the attainment of a stable steady-state size (Oles 1992,

Spicer & Pratsinis 1996a, Serra & Casamitjana 1998b, Flesch et al. 1999). However,

polymer flocculants do not give a steady-state size except if flocculation occurs

primarily through surface charge neutralisation. Aggregation by polymer bridging

typically results in a decreased aggregate size on extended shearing due to flocculant

degradation through chain scission or re-arrangement (Keys & Hogg 1978, Sikora &

Stratton 1981, Leu & Ghosh 1988, Williams et al. 1992, Bagster 1993, Hogg 2000).

Polymer flocculant degradation could be incorporated into the model with a decreasing

capture efficiency (Heath et al. 1999), or by an increased breakage rate via a weaker

aggregate. The latter approach was taken here, keeping the terms together in the

breakage kernel:

⎟⎟⎟

⎜⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

⎛φφ

Θ−μφ∈=Θ 3

1

3

s

effs

k4 1Mk

dtd 5-20

with initial condition: Θ = 0 at t = 0. and: k4 = Fitted parameter # 4 := 1224 The rate of polymer degradation is taken to be a function of the breakage rate, although

additional solid fraction dependence was required to achieve an acceptable fit with the

data. The dimensionless term (φeff/φs)1/3 is introduced to prevent the polymer becoming

Page 138: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

137

completely degraded. The term can be replaced by ≈ 2 (allowing 50 % degradation)

giving only a slight increase in the residual. However, the term used has the additional

advantage that it collapses to 1 (complete degradation) under unfavourable conditions

(e.g. extreme shear), allowing the aggregate size distribution to eventually return to the

primary particle size.

The physical interpretation of this effect is not clear, and could be taken as evidence of

flocculation by surface charge neutralisation. However in this case an anionic polymer

was used, and the calcite surface is also expected to be negatively charged (Geffroy

et al. 2000). Alternatively it could be taken as evidence that the flocculant binding

together the core of aggregates is not significantly degraded, because the aggregates are

not broken down to that extent. The polymer degradation would then be expected to be

a function of the aggregate surface area, volume, packing efficiency etc. The term used

is entirely empirical, although it is in a similar form to that used by Strenge (1993) to

describe the mean distance between particles in the aggregate.

5.1.3 Simplification for coding/simulation

Equations 5-6 and 5-16 effectively switch off the population balance when no flocculant

has been added. This is necessary to ensure that the primary particles do not aggregate

(or break), and the mean size is maintained when no flocculant is present. However,

although gPROMS (commercial simulation package described later) will accept step

functions (e.g. Equation 5-19) in the ordinary simulation mode, it will not perform

parameter estimations on codes with case functions. Hence Equation 5-19 is replaced

by the continuous approximation:

( )⎟⎠⎞

⎜⎝⎛ −

θ

μ∈= −

5pagg

3d/d

f

i,aggk

2i e1

dKS 5-21

where: dp = Volume weighted mean primary particle diameter (m)

In this case dp was fixed at 30 × 10-6 m, which is larger than any of the experimental

primary mean sizes, and below any of the mean aggregate sizes in the breakage

dominant region.

Page 139: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

138

If equations describing aggregate porosity (e.g. Equations 5-11, 5-12 & 5-20) are

inserted directly into the population balance (Equation 5-1), the model’s complexity

increases significantly and the convergence time becomes far longer. In order to

simplify the population balance, the fractal geometry describing aggregate porosity and

size can be calculated externally, bringing them before the integral.

The Hounslow/Spicer population balance conserves mass based on the assumption:

2V

Vi

1i

=+

5-22

where: Vi = Volume of the particles in the ith size range (m3) and with the initial size range set to some suitable value below the smallest primary

particle.

Equation 5-22 is only suitable for non-porous particles, e.g. for droplet coalescence in

solvent extraction or clouds. However, Equation 5-22 does provide a basis for mass

conservation via the assumption that Vi can be substituted by Vm,i, the mass equivalent

volume giving (see Appendix):

fD

3

p

i,mpi,agg d

ddd ⎟

⎟⎠

⎞⎜⎜⎝

⎛= 5-23

where: dagg,i = Diameter of the ith sized aggregate (m) dp = Diameter of the primary particle (m) dm,i = Mass effective diameter of the ith sized aggregate (m) In this case 35 size channels were used, covering a size range 0.2-3500 μm.

The population balance model described above contains 4 fitted parameters (Equations

5-6, 5-19 & 5-20) that are estimated from the experimental size data using a

conventional sum-of-squares minimisation:

( )2t,,d,,G

Exp,aggPB,aggk,k,k,k

∑p4321

d-dminθφ

=ψ 5-24

where: kn = Fitted parameters G = Spatially averaged shear rate (s-1) φ = Solid fraction (m3m-3)

θ = Flocculant surface coverage (kg m-2) dp = Primary particle size (m)

Page 140: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

139

dagg,PB = Modelled (PB = population balance) mean aggregate size (m) dagg,Exp = Experimental (FBRM) mean aggregate size (m) t = Time (s)

Simulations and parameter estimation were performed using gPROMS, a commercial

UNIX based dynamic simulation package, running on a SUN Enterprise 3000

mainframe. Individual simulations took in the order of 10 seconds to converge, while

parameter estimation using the 17 available data sets took several hours depending on

the level of accuracy required.

The fitted parameters are used to fit the model to the experimental data, with the

parameters adjusting various aspects of the model behaviour. The first parameter (k1)

alters the initial flocculant/suspension mixing rate in the pipe, changing the capture

efficiency and hence the initial rate of aggregation. The fitted value of 0.343 compares

favourably with the literature value of 0.5 suggested by Etchells and Short (1988).

However in its current form the mixing equation (Equation 5-6) only describes mixing

in turbulent pipe flow, and requires modification for other systems. Ultimately,

flocculant/suspension mixing and adsorption in thickener feedwells is likely to be

described by CFD, perhaps with validation by electrical impedance tomography.

The second model parameter (k2) effectively scales the breakage rate, and is likely to be

a complex function of the various factors affecting the aggregate strength. For example:

the flocculant/particle surface bonding chemistry, the flocculant molecular weight, the

presence of coagulating salts (if any), or the primary particle packing structure within

the aggregate.

The third fitted parameter (k3) essentially alters the effect of the energy dissipation rate

on the final aggregate size, producing a smaller aggregate at higher shear rates as

described above. The final parameter (k4) sets the rate of polymer degradation at

extended residence times. Due to the complex interrelated nature of the population

balance equations the parameters are somewhat interactive, with any deficiencies in the

model effectively accommodated elsewhere in the model. However, the model was

found to be robust and stable, converging to the same parameter values from a range of

initial estimates.

Page 141: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

140

5.2 Comparison with experimental data 5.2.1 Effect of flow regime (shear rate)

Figure 5-1 compares the modelled and experimental mean aggregate size data for

calcite under various flow conditions (see Section 4). A combination of pipe sizes (25.4

or 38.1 mm ID) and flow rates gave different fluid shear rates. As has been discussed

previously (Section 4), a higher mean fluid shear results in a rapid initial aggregation

rate, but ultimately results in a smaller mean aggregate size due to increased breakage.

The population balance model (lines) follows the same trends due to the interplay

between various factors. At a higher shear rate the flocculant mixing (Equation 5-6,

Figure 5-2) and particle collision (Equation 2-12) rates are higher, resulting in a higher

initial aggregation rate. However, a higher energy dissipation rate increases the

breakage (Equation 5-19) and polymer degradation (Equation 5-20) rates, resulting in a

smaller average aggregate size.

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 10 20 30 40 50 60 70 80

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Model, 38.1 mm ID, 0.343 m sֿ¹Experimental, 38.1 mm ID, 0.343 m sֿ¹, G = 52.6 sֿ¹, (settling obs. in pipe)Model, 38.1 mm ID, 0.554 m sֿ¹Experimental, 38.1 mm ID, 0.554 m sֿ¹, G = 175.2 sֿ¹Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)Experimental, 25.4 mm ID, 0.461 m sֿ¹, G = 174.1 sֿ¹ (Baseline)Model, 25.4 mm ID, 0.781 m sֿ¹Experimental, 25.4 mm ID, 0.781 m sֿ¹, G = 488.1 sֿ¹Model, 25.4 mm ID, 1.294 m sֿ¹Experimental, 25.4 mm ID, 1.294 m sֿ¹, G = 935.5 sֿ¹

Figure 5-1: Modelled (lines) and experimental (dots) change in mean aggregate size under different pipe flow conditions. At the lowest shear rate (open circles) solid settling was observed in the pipe, and this data should be interpreted with caution. Other conditions as per baseline.

Figure 5-2 shows the modelled flocculant/suspension mixing index (M, Equation 5-6)

as a function of time and the flow regime, allowing an estimation of the capture

efficiency (α), which is also taken to approach 1 when the flocculant becomes well

mixed. The actual capture efficiency may be lower than 1, although the real value is

unknown and any discrepancy is effectively compensated by the breakage kernel.

Page 142: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

141

0

0.2

0.4

0.6

0.8

1

1.2

0 10 20 30 40 50 60 70

Time (s)

Mix

ing

Inde

x ( α

=M)

Model, 25.4 mm ID, 1.294 m sֿ¹Model, 25.4 mm ID, 0.781 m sֿ¹Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)Model, 38.1 mm ID, 0.554 m sֿ¹Model, 38.1 mm ID, 0.343 m sֿ¹

Figure 5-2: Variation in modelled mixing index with time under different flow conditions.

Figures 5-1 and 5-2 suggest that the initial aggregation rate is limited by

flocculant/suspension mixing, rather than the particle collision rate (Equation 2-12).

This is supported by experimental evidence where the mixing condition was changed,

either by flocculant dilution or by changing the flocculant injection velocity by

changing the injection nipple size. Figure 5-3 shows the increase in the aggregate size

with flocculant dilution with a fixed length of short pipe (compare Figure 5-1, 38.1 mm

ID, 0.55 m s-1). Using a short length of pipe and gentle shear rate allowed the particle

size measurement to be taken where the aggregate size was still increasing rapidly and

was likely to be influenced by the mixing condition. The overall dosage was

maintained at 20 g t-1, but a dilute flocculant stream produced better mixing due to a

higher flocculant stream flow, and lower viscosity.

Increasing the flocculant injection velocity by changing the nipple diameter (not shown)

gave a similar result, with a smaller nipple increasing the flocculant stream momentum,

the turbulence at the injection point, producing an increased aggregate size. In all cases

the flocculant stream was minor (0.25-2 %) compared to the main slurry flow in the

pipe reactor.

Page 143: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

142

0.0E+00

2.0E-05

4.0E-05

6.0E-05

8.0E-05

1.0E-04

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09Flocculant concentration (% w/v)

Figure 5-3: Effect of flocculant stream dilution on the measured aggregate size for a fixed (short) length of pipe, giving a mean residence time of 3 s, see Figure 5-1. 38.1 mm ID pipe, 0.34 m s-1 mean flow velocity.

Figure 5-4 shows the variation in the modelled fluid viscosity under the conditions

described by Figure 5-1. The fluid viscosity is taken to be a function of the effective

solid fraction (Equation 5-16), including aggregate porosity using fractal geometry.

Since aggregates become increasingly porous as they increase in size, the total enclosed

volume of the aggregates also increases, leading to a higher effective solid fraction

(Equation 5-14) and hence fluid viscosity (Williams et al. 1992, Barnes & Holbrook

1993). This effectively prevents gelation, where the aggregates might grow to the point

where they fill the container, because as the effective solid fraction approaches the

maximum (φm), the viscosity tends towards infinity (Equation 5-13). This leads to an

increased pressure drop (as measured by manometer), and an increased energy

dissipation rate (the flow velocity is fixed via a positive displacement pump), and hence

an increased breakage rate (Equation 5-19).

Since the fluid viscosity is dependent on the aggregate size, various other fluid

parameters also change and are calculated within the population balance. Figures 5-5,

5-7 & 5-8 show changes to the mean shear rate, pipe Reynolds' number and the friction

factor respectively. The use of manometers allows experimental estimates of the same

fluid properties, with the predicted and experimental average shear rates compared in

Figure 5-6. These fluid parameters are described using standard pipe flow equations

(Equations 5-7 – 5-10) incorporating the effect of solid fraction on the suspension

Page 144: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

143

viscosity (Equations 5-13 – 5-16), achieving an acceptable fit to the data (Figure 5-6)

without the need for any additional fitted parameters.

0

0.001

0.002

0.003

0.004

0 10 20 30 40 50 60 70

Time (s)

Vis

cosi

ty (N

s m

-2)

Model, 38.1 mm ID, 0.343 m sֿ¹

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 25.4 mm ID, 0.781 m sֿ¹

Model, 25.4 mm ID, 1.294 m sֿ¹

Figure 5-4: Variation in the modelled fluid viscosity under different flow regimes. Other conditions as per baseline.

0

200

400

600

800

1000

1200

0 10 20 30 40 50 60 70

Time (s)

Mea

n tu

rbul

ent s

hear

rat

e (s

-1) Model, 25.4 mm ID, 1.294 m sֿ¹

Model, 25.4 mm ID, 0.781 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 38.1 mm ID, 0.343 m sֿ¹

Figure 5-5: Variation in modelled spatially averaged shear rate under various flow regimes.

Page 145: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

144

10

100

1000

10 100 1000

Measured mean shear rate (s-1)

Mod

elle

d m

ean

shea

r ra

te (s

-1)

dp = 2.36 µm

dp = 3.47 µm

dp = 15.1 µm

Dosage = 5 g tֿ¹

Dosage = 10 g tֿ¹

38.1 mm ID, 0.554 m sֿ¹

Baseline

25.4 mm ID, 0.781 m sֿ¹

25.4 mm ID, 1.294 m sֿ¹

3.33 % w/v

6.67 % w/v

13.33 % w/v

16.67 % w/v

10 g tֿ¹, 38.1 mm ID, 0.343 m sֿ¹

10 µm, 7 % w/v

40 g tֿ¹, 25.4 mm ID, 0.781 m sֿ¹

Figure 5-6: Comparison between modelled (Equation 5-7 – 5-16) and experimental (Equation 4-1 – 4-7) spatially averaged shear rates.

0

5000

10000

15000

20000

25000

30000

35000

0 10 20 30 40 50 60 70

Time (s)

Pipe

Rey

nold

s Num

ber

(Re) Model, 25.4 mm ID, 1.294 m sֿ¹

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 38.1 mm ID, 0.343 m sֿ¹

Figure 5-7: Variation in the modelled pipe Reynolds’ number under different flow regimes.

Page 146: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

145

0

0.002

0.004

0.006

0.008

0.01

0.012

0 10 20 30 40 50 60 70

Time (s)

Pipe

fric

tion

fact

or (f

)

Model, 38.1 mm ID, 0.343 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 25.4 mm ID, 0.781 m sֿ¹

Model, 25.4 mm ID, 1.294 m sֿ¹

Figure 5-8: Variation in the modelled pipe Fanning friction factor under different flow regimes.

Figure 5-9 shows the population balance correctly conserved mass through time

according to:

∑ ρ=φ

i s

iit

mN 5-25

Where: ρs = Density of the solid (2710 kg m-3) mi = Mass of ith aggregate (kg) Ni = Number of ith aggregates (m-3)

9.99

10

10.01

0 10 20 30 40 50 60 70

Time (s)

Solid

vol

ume

frac

tion

(% w

/v)

Model, 38.1 mm ID, 0.343 m sֿ¹

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 25.4 mm ID, 0.781 m sֿ¹

Model, 25.4 mm ID, 1.294 m sֿ¹

Figure 5-9: Variation in the modelled absolute mass fraction under different flow regimes. Conservation of mass.

Page 147: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

146

Aggregation with polymer flocculant typically results in partially irreversible breakage,

manifesting as a gentle reduction in the aggregate size after an initial peak as the

polymer becomes degraded (Bagster 1993). Figure 5-10 shows the extent of flocculant

degradation predicted by the model under various flow conditions. Due to the

incorporation of the (φeff/φs)1/3 term in Equation 5-20, the degradation tends towards a

value of less than 1 (where 1 would represent complete degradation). This improves the

fit with the experimental data, and is rationalised on the basis that the repeated

aggregation/breakage process will not affect flocculant binding the core of aggregates,

because the aggregates are not broken down to that extent. It could also be interpreted

as evidence supporting the electrostatic patch model of flocculation (Kasper 1971,

Gregory 1973), since polymer chain scission is unlikely to affect the flocculant’s

charge. However, if the flow conditions are very aggressive, aggregation will only

proceed to a limited extent and φeff → φs, therefore the flocculant could be completely

degraded. Such a situation is clearly undesirable industrially, but may occur around

feed entry points or baffles with high local shear rates.

0

0.1

0.2

0.3

0.4

0.5

0.6

0 10 20 30 40 50 60 70

Time (s)

Poly

mer

deg

rada

tion

inde

x Θ

Model, 25.4 mm ID, 1.294 m sֿ¹

Model, 25.4 mm ID, 0.781 m sֿ¹

Model, 25.4 mm ID, 0.461 m sֿ¹ (Baseline)

Model, 38.1 mm ID, 0.554 m sֿ¹

Model, 38.1 mm ID, 0.343 m sֿ¹

Figure 5-10: Variation in the modelled polymer degradation index under different flow regimes.

Page 148: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

147

5.2.2 Effect of flocculant dosage

Figure 5-11 shows the predicted and experimental mean aggregate sizes as a function of

time and flocculant dosage. The dosage does not affect the aggregation term in the

population balance (except for a slight change in the initial mixing due to a slight

change in the flocculant stream momentum) and the capture efficiency is taken to

approach 1 as the flocculant/suspension mixing becomes complete, regardless of the

actual flocculant dosage. However, flocculant dosage is incorporated into the breakage

kernel of the population balance (Equation 5-19), where a higher dosage leads to a

stronger aggregate, a decreased breakage rate, and hence a larger aggregates. The

flocculant dosage is known experimentally on a mass basis, but is incorporated into the

population balance as a surface coverage, i.e. kg(floc)/m2(solid) (Equation 5-18). The

surface area of the primary particles was measured by laser diffraction (Malvern

Mastersizer) and BET, and hence the effective surface coverage also depends on the

primary particle size (Figure 5-12).

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Model, dosage = 20 g tֿ¹, (Baseline)Experimental, dosage = 20 g tֿ¹, (Baseline)Model, dosage = 10 g tֿ¹Experimental, dosage = 10 g tֿ¹Model, dosage = 5 g tֿ¹Experimental, dosage = 5 g tֿ¹

Figure 5-11: Modelled (lines) and experimental (dots) change in mean aggregate size with change in flocculant dosage. Other conditions as per baseline.

Higher dosages were also tried (see Section 4), but resulted in stratification/settling in

the pipe reactor, leading to uncertainty in the solid fraction and residence time. Both

the modelled and experimental results gave larger aggregates at higher dosages,

although the data match was poor (not shown). At extremely high flocculant dosages

there is data to suggest (Healy 1961, Fleer & Scheutjens 1993, Svarovsky 2000) that the

aggregate size can be reduced by steric re-stabilisation as the particle surface becomes

Page 149: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

148

completely covered, although it is unlikely that such a high dosage would be reached in

industrial applications (Hogg 2000).

5.2.3 Effect of primary particle size

The primary particle size distribution is also likely to be a process variable, depending

on the feed (milling, crystallisation, ore etc). This effect was incorporated into the

population balance model by assuming that the aggregate breakage rate is a function

(Equation 5-18) of the effective flocculant surface coverage. That is, a larger primary

particle, with a smaller effective surface area (per unit mass) has a higher coverage, a

higher strength, and hence results in larger aggregates (Figure 5-12).

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

2.5E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Model, dp = 15.1 µmExperimental, dp = 15.1 µmModel, dp = 5.59 µm, (Baseline)Experimental, dp = 5.59 µm, (Baseline)Model, dp = 3.47 µmExperimental, dp = 3.47 µmModel, dp = 2.36 µmExperimental, dp = 2.36 µm

Figure 5-12: Modelled (lines) and experimental (dots) change in mean aggregate size using differently sized primary particles. Other conditions as per baseline.

Figure 5-12 shows that using the effective flocculant dosage gives the correct model

response with respect to primary particle size, although the modelled size

underestimates the experimental data for small primary particles. An alternative

particle surface area measurement (BET) was also tried, giving similar results, but no

improvement in the fit. Although the discrepancy is clearly noticeable to the eye, the

residual error is comparatively minor and did not justify the addition of a further fitted

parameter.

Alternatively, the aggregate strength could be considered to be a function of the packing

efficiency within the aggregate (Gregory 1989, Mühle 1993), which would change with

Page 150: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

149

primary particle size according to Equation 5-23. Attempts were made along these lines

during model construction, but ultimately did not fit the experimental data as well as

relating the aggregate strength to the flocculant surface coverage. The latter also gave a

simpler model.

The incorporation of fractal geometry into the population balance to account for

aggregate porosity complicates the modelling of different primary particle sizes. Fractal

geometry describes the increase in aggregate porosity with aggregate size. However,

since the primary particles are distributed (known via Malvern Mastersizer) and the

population balance does not distinguish between aggregates and primary particles, the

smaller primary particles will be undersized by the population balance. Conversely, the

size of large primary particles will be overestimated, as they cannot be distinguished (by

the population balance) from small aggregates of the same mass.

This issue was corrected by changing the number of primary particles in each channel

of the population balance so that the correct number of collision and breakage events

occurred. The distribution was still based on the Malvern volume averaged data, and

the total number of primary particles normalised to give the correct solid fraction

(10 % w/v for most runs). The particle porosity was taken to be zero at the mean

primary particle size (taken as 6.59 × 10-6 m), however in situations when the mean

primary particle size was different (i.e. Figure 5-12), the normalisation step resulted in

the total solid fraction being different from the actual 10 % w/v. Altering the calculated

total mass fraction to give the correct aggregation/breakage rate was seen as preferable

to the reverse. Since the effective viscosity was based on the effective solid volume

fraction (which is given correctly), rather than the absolute solid fraction, the fluid

parameters (shear rate, Reynolds number, friction factor etc) are also unaffected by the

above correction.

The above issue could also be tackled more rigorously using a two-dimensional

population balance, where the distribution of primary particles within each aggregate

was accounted for explicitly. However, this would considerably complicate the model,

requiring the use of 1225 (352) model equations rather than the 35 used here, almost

certainly making the model too unwieldy for use as a sub-model within a CFD code.

Page 151: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

150

5.2.4 Effect of solid fraction

The solid fraction is also a process variable and can be altered in some processes by

overflow or underflow recycle (Perry & Green 1997). Figure 5-13 shows a comparison

between modelled and experimental mean aggregate sizes as a function of time and

solid fraction. The reduction in aggregate size with solid fraction was somewhat

unexpected, given that aggregation (Equation 5-2) is usually taken as second order with

respect to particle number, whereas breakage is taken as first order (Equation 5-1 &

5-19). However, by making the breakage rate a function of viscosity (Figure 5-14) and

the energy dissipation rate (which also increases with the viscosity;

Equations 5-13 - 5-16) the breakage rate was increased at high solid fraction, producing

the desired reduction in the aggregate size (Figure 5-13). In addition, an increased solid

fraction gives a reduced shear rate (Figure 5-15) reducing the aggregation rate. A

reduction in aggregate size with solid fraction has been observed elsewhere (Lick &

Lick 1988, Williams et al. 1992).

At very low solid fraction (< 3 % w/v), the viscosity approaches that of pure water, and

the model predicts a reduction in aggregate size as the solid fraction is further decreased

due to the difference in the order of the aggregation and breakage kernels. The increase

in aggregate size with increased solid fraction has been observed previously in studies

at very low solid fraction (Thomas 1964, De Boer et al. 1989, Kobayashi et al. 1999).

Lower solid fractions were not investigated here due to the difficulty of measuring the

hindered settling velocity (see Section 6) at low solid fraction where the mudline is no

longer distinct, and the settling rate is rapid.

Page 152: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

151

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Model, 3.33 % w/v Experimental, 3.33 % w/vModel, 6.67 % w/v Experimental, 6.67 % w/vModel, 10.0 % w/v, (Baseline) Experimental, 10.0 % w/v, (Baseline)Model, 13.33 % w/v Experimental, 13.33 % w/vModel, 16.67 % w/v Experimental, 16.67 % w/v

Figure 5-13: Modelled (lines) and experimental (dots) change in mean aggregate size using different solid fractions. Other conditions as per baseline.

0

0.001

0.002

0.003

0.004

0.005

0 5 10 15 20 25 30 35 40

Time (s)

Vis

cosi

ty (N

s m

-2)

Model, 16.67 % w/vModel, 13.33 % w/vModel, 10.0 % w/v, (Baseline)Model, 6.67 % w/vModel, 3.33 % w/v

Figure 5-14: Variation in modelled fluid viscosity through time with various solid fractions.

Page 153: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

152

0

50

100

150

200

250

300

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n tu

rbul

ent s

hear

rat

e (s

-1)

Model, 3.33 % w/vModel, 6.67 % w/vModel, 10.0 % w/v, (Baseline)Model, 13.33 % w/vModel, 16.67 % w/v

Figure 5-15: Variation in modelled spatially averaged shear rate through time with various solid fractions.

Figure 5-16 again shows the population balance conserved mass correctly. Although

conservation of mass does not guarantee that the population balance is working

correctly, it is simple to check and picks up gross coding and numerical errors that

frequently result in the model gaining or losing mass.

0

5

10

15

20

0 5 10 15 20 25 30 35 40

Time (s)

Solid

vol

ume

frac

tion

(% w

/v) Model, 16.67 % w/v

Model, 13.33 % w/vModel, 10.0 % w/v, (Baseline)Model, 6.67 % w/vModel, 3.33 % w/v

Figure 5-16: Variation in the modelled absolute mass fraction with different solid fractions. Mass is conserved.

Page 154: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

153

5.2.5 Additional runs to check model predictivity

Figures 5-1, 5-11, 5-12 and 5-13 show results from a sparse matrix of experimental pipe

reactor runs (see Section 4, Figure 4-4) performed using the baseline as the central

point, and varying the process variables (fluid shear, flocculant dosage, solid fraction,

primary particle size) independently away from that the centre. However, additional

experimental runs were performed in the gaps of the matrix, by changing two process

variables simultaneously. These were not included in the parameter estimation for the

population balance, but were used to assess the predictivity of the model (Figure 5-17).

The predictions appear reasonable, except for the latter stages of the 15.1 μm,

6.67 % w/v run. This appears to be due to the effect of the primary particle size, which

shows some deviation from the experimental values (see Figure 5-12).

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

2.5E-04

3.0E-04

0 10 20 30 40 50 60 70Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

Model, dp = 15.1 µm, 6.67 % w/vExperimental, dp = 15.1 µm, 6.67 % w/vModel, dosage = 40 g tֿ¹, 25.4 mm ID, 0.781 m sֿ¹Experimental, dosage = 40 g tֿ¹, 25.4 mm ID, 0.781 m sֿ¹Model, dosage = 10 g tֿ¹, 38.1 mm ID, 0.343 m sֿ¹Experimental, dosage = 10 g tֿ¹, 38.1 mm ID, 0.343 m sֿ¹

Figure 5-17: Modelled (lines) and experimental (dots) change in mean aggregate size with change under various additional conditions. Other conditions as per baseline.

Page 155: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

154

5.2.6 Effect of changing fractal dimension

In all the above figures a fractal dimension of 2.4 was assumed, as per the estimation

from the experimental settling data (Section 6). This allowed a dramatic simplification

of the population balance, by reading the effective aggregate sizes into the population

balance as an array. However, an alternative model was also coded, where fractal

geometry was incorporated directly into the population balance’s differential equations.

The convergence time was increased by an order of magnitude, and simply changing the

array values to suit the fractal dimension would probably have been a simpler modelling

path in retrospect.

Figure 5-18 shows the effect of changing the fractal dimension on the predicted mean

aggregate size. A higher fractal dimension (lower porosity) results in an initially

reduced aggregation rate due to a reduced aggregate capture radius. However, the

lower porosity gives a lower effective solid fraction, reducing the viscosity and the

breakage rate, hence the increased aggregate size later in the aggregation process.

Unfortunately, there are no experimental data available to confirm the trend in Figure 5-

18, and it is not clear how the fractal dimension could be changed in isolation.

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n A

ggre

gate

size

(m)

Df = 2.6Df = 2.5Df = 2.4Df = 2.3Df = 2.2

Figure 5-18: Modelled variation in the mean aggregate size with fractal dimension (i.e. aggregate porosity). Other conditions are as per baseline.

Page 156: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

155

5.3 Conclusions A population balance model has been developed to describe the kinetics of

aggregation/breakage of calcite particles flocculated with high molecular weight

flocculant in turbulent flow. In addition to the effect of fluid shear normally

incorporated into population balance models of aggregation, the model accounts for the

important full-scale process variables, namely: flocculant dosage, primary particle size

and solid fraction, representing a significant advance on the population balance models

described previously in the literature.

The population balance is based on the successful population balance proposed by

Hounslow et al. (1988), later modified to incorporate aggregate breakage by Spicer and

Pratsinis (1996a). Saffman and Turner’s (1956) collision kernel is used in conjunction

with a capture efficiency (α) term based on the degree of flocculant/suspension mixing.

This accounts for the stability of the suspension before flocculant addition, and prevents

the overestimation of the aggregation rate during the initial mixing period.

The breakage kernel is a function of the mean energy dissipation rate, suspension

viscosity and aggregate diameter. Flocculant dosage dependence is also incorporated

into the breakage term, with a higher dosage increasing the aggregate strength, leading

to a reduced breakage rate and ultimately a larger aggregate size. The partial non-

reversibility of polymer aggregate breakage is included with a decay term that accounts

for the flocculant degradation due to repeated aggregation/breakage.

The effects of aggregate porosity are incorporated by the use of fractal geometry,

increasing both the effective capture radii and the effective solid fraction. The increase

in the suspension viscosity with solid volume fraction is also incorporated, and results

in changes in the energy dissipation rate, fluid shear, pipe friction factor and Reynolds

number. The inclusion of these additional terms allow the population balance to

describe aggregation and breakage kinetics in suspensions of high solid fraction as

found in mineral processing thickeners, as opposed to the low solid fractions considered

by most workers studying estuarine and wastewater systems.

Page 157: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

156

6. Relating hindered settling rate with aggregate size

Previous sections have covered the effect of various process variables on the resulting

aggregate size. However, the primary objective of aggregation is to increase the settling

rate and hence efficiency of solid-liquid separation, with the actual aggregate size being

of secondary importance. This section investigates the relationship between the

aggregate size and the resulting hindered settling rate, ultimately allowing the settling

rate to be predicted from the population balance model.

The gravity sedimentation velocity of individual solid spheres settling in creeping flow

(Re → 0) was first described theoretically by Stokes (1851) (Happel & Brenner 1973,

Seville et al. 1997):

( )μ

ρ−ρ=

18gdU ls

2

6-1

where: U = Settling velocity (m s-1) d = Particle diameter (m) g = Gravity (9.8 m s-2) ρ = Density (s = solid, l = liquid) (kg m-3)

μ = Fluid viscosity (N s m-2)

In practice, mineral particles/aggregates are rarely spherical, and large aggregates may

be highly porous. Non-spherical particles generally settle slightly slower (~ 0.7-1.0)

than spheres (Pettyjohn & Christiansen 1948, Heiss & Coull 1952, Happel & Brenner

1973, Kousaka et al. 1981, Seville et al. 1997), although needle shaped particles settle

slightly faster if the major axis remains vertical. Porous aggregates have larger

hydrodynamic profiles than equivalent mass solid particles, considerably reducing their

settling velocity (Gregory 1997). However, highly porous aggregates may allow some

fluid flow through the structure, counteracting this effect somewhat (Veerapaneni &

Wiesner 1996).

Large, dense particles may settle too rapidly to be accurately described by Stokes’ law,

and by a particle Reynolds number above about 0.1 inertial effects become significant

(Rushton et al. 1996, Seville et al. 1997) and the drag coefficient increases faster than

Stokes’ law describes, reducing the settling velocity (Table 2-7). Both empirical and

theoretical relationships have been proposed for higher Reynolds number settling

Page 158: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

157

(Moudgil & Vasudevan 1989, Clift et al. 1978, Nguyen-Van et al. 1994, Perry & Green

1997), although in most cases aggregates settle sufficiently slowly to be described by

Stokes’ equation (Gregory 1997)

As the suspension solid fraction is increased there is a gradual, but poorly defined,

transition to hindered-settling. The hindered settling regime is characterised by a

distinct solid/liquid interface (mudline) that settles to leave a clear supernatant above

(Bhatty et al. 1982, Fitch 1975a, 1987, Tong et al. 1998). The settling velocity of the

mudline decreases as a function of the solid fraction, due to the decreasing permeability

of the settling layer and an increase in the upward velocity of the displaced fluid

(Govier & Aziz 1972, Pearse 1977, Perry & Green 1997, Bustos et al. 1999), as

described by Richardson and Zaki (1955) (Tables 2-8 & 2-9):

( )no 1UU φ−= 6-2 where:

U = Settling velocity (m s-1) Uo = Settling velocity at infinite dilution (m s-1) n = Exponent, usually taken as 4.65 (Table 2-9). φ = Solid volume fraction [0,1]

Aggregation increases the hindered-settling velocity by increasing the particle size,

forming channels that increase the sediment permeability. Settling behaviour in the

hindered settling regime is usually described with Kynch (1952) theory, but it does not

account for the effect of flocculation (Williams & Simons 1992, Svarovsky 2000)

because the settling velocity is usually taken as a unique function of the solid fraction.

As the solid loading is increased further, the particles will eventually form a continuous

network and settling will also be restrained by mechanical support from below

(Michaels & Bolger 1962, Pearse 1977, Chandler & Hogg 1987, Healy et al. 1994,

Perry & Green 1997). The mechanical strength of the network is a function of the

solid/packing fraction and the strength of inter-particle bonding. The sediment will

compress when the weight of sediment overburden exceeds the compressive yield stress

(Buscall & White 1987, Healy et al. 1994, Green et al. 1996, Bustos et al. 1999). The

excess overburden weight not supported mechanically is restrained hydrodynamically,

and is the force required to squeeze the fluid back up through the collapsing sediment

Page 159: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

158

(Fitch 1975b, Wakeman & Holdich 1984, Concha et al. 1996, Diplas & Papanicolaou

1997, Perez et al. 1998). In practice channel formation may occur spontaneously, or

can be encouraged by rakes with vertical pickets (Pearce 1977, Dahlstrom & Fitch

1985, Williams 1992, Svarovsky 2000).

The aim of this section is to develop a validated mathematical relationship between the

aggregate size and the initial hindered settling rate under a range of conditions likely in

a mineral processing thickener. The influence of fluid shear, flocculant dosage, feed

solid volume fraction, primary particle size and residence time are addressed.

6.1 Experimental Aggregation of calcite (Omya-carb, Commercial Minerals) with polymeric flocculant

(Nalco 9902) was performed in a linear horizontal pipe reactor, as described in

Section 4. Size measurements were taken using the Lasentec FBRM probe described in

Section 3, and hindered settling rates taken in a graduated 0.5 m settling column

(Figures 4-1 & 6-1).

The settling column was constructed from acrylic tube (38.1 mm ID), and the flow was

isolatable (note valves) and removable for extended settling measurements. The valve

arrangement allowed a portion of the flow to be captured in the settling column, without

disrupting the flow through the pipe reactor and past the sizing probe, allowing

continuous operation. The column was marked at regular intervals (1 & 5 cm) and lit

from behind by fluorescent tube, allowing the measurement of the hindered settling rate

by following the fall of the mudline through time using a stopwatch.

The pipe work and fittings were designed to keep the top of the settling column as close

as possible to the particle sizing probe, reducing the additional residence time. This

was the primary reason for mounting the pipe reactor above floor level (note

Figure 6-1). The fittings around the bend and valve were also constructed to produce a

smooth inner surface in an effort to reduce turbulence. Despite this, the time offset was

still considerable (see Section 6.2.1).

Page 160: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

159

Figure 6-1: Schematic valve arrangement to capture portion of flow in settling column.

6.2 Results and discussion 6.2.1 Effect of fluid shear

Figure 6-2 shows the development of the mean aggregate size and corresponding

hindered settling velocity as a function of the spatially averaged fluid shear rate and

residence time. As described previously (Sections 4 & 5), a high turbulent shear rate

initially increases the aggregation rate by increased mixing and particle collision.

However, aggregate breakage is dramatically increased at high shear rates, and larger

aggregates are ultimately formed at a lower shear rate. The hindered settling velocity

follows a similar trend, but there is a time offset due to the additional pipe-work and

fittings between the sizing probe and top of the settling column (Figure 6-1).

Linked ball valves

FBRM sizing probe

FBRM sizing probe

Linked ball valves

To waste

Pipe reactor

To waste

Pipe reactor

Settling column Settling

column

Page 161: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

160

-1.0E-04

-5.0E-05

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 10 20 30 40 50 60 70

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

0

5

10

15

20

Hin

dere

d se

ttlin

g ve

loci

ty (m

hr-1

)

38.1 mm ID, 0.343 m sֿ¹, (settling obs. in pipe)

38.1 mm ID, 0.554 m sֿ¹

25.4 mm ID, 0.461 m sֿ¹, (Baseline)

25.4 mm ID, 0.781 m sֿ¹

25.4 mm ID, 1.294 m sֿ¹

Figure 6-2: Effect of mean pipe shear rate on the development of the mean aggregate size and hindered settling velocity. Other conditions as per baseline.

The additional time offset was calculated by the velocity-head method (Perry & Green

1997), using the Bernoulli equation for incompressible fluids:

2

VK)zz(g2V

2VP 2

12

211

222 +−+

α−

α=

ρΔ 6-3

where: α = Velocity profile factor g = Gravity (9.8 m s-2) z = Elevation (m), unchanged in this case with horizontal pipe

ΔP = Pressure drop (N m-2) V = Mean fluid velocity (m s-1)

ρ = Fluid density (kg m-3) K = Velocity head, tabulated values for fittings, or for pipes:

DfL4K = 6-4

where: f = Fanning friction factor (= 0.079/Re0.25 for turbulent flow, smooth pipes) L = Pipe length (m), 0.27 m in this case D = Pipe diameter (m), 0.040 m in this case

Figure 6-1 shows that in this case several fittings are present between the particle sizing

probe and the settling column. The valve on the settling tube was assumed to be fully

open with K = 0.17 as suggested by Perry and Green (1997). The protruding sizing

probe was also assumed to have K = 0.17, and the tee given as K = 1.0 (ibid.). In

addition the expansion loss when using the smaller diameter pipe reactor (25.4 mm ID)

was given by the Borda-Carnot equation (ibid.):

Page 162: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

161

2

2

12

1

AA1

2VP

⎟⎟⎠

⎞⎜⎜⎝

⎛−=

ρΔ 6-5

where: A = Cross sectional area of the pipe (m2) The contribution from the various fittings and pipe were combined, and converted to an

equivalent time depending on the flow velocity and size of the main pipe. The

additional equivalent time between the FBRM sizing probe and the top of the settling

column varied from 0.44 s (25.4 mm pipe, 39.33 L min.-1) to 7.31 s (38.1 mm pipe,

14.16 L min.-1). In Figure 6-3 below (25.4 mm pipe, 14.01 L min.-1) the offset was

1.86 s. The additional time offset is unfortunate since the data is lost from low

residence times where the aggregate size increases rapidly. However, it does serve to

reinforce the observation that flocculation is rapid in well mixed suspensions of high

solid fraction.

6.2.2 Effect of flocculant dosage

Figure 6-3 shows the increase in the aggregate size and hindered settling velocity with

flocculant dosage. The increased size is attributed to a higher flocculant surface

coverage and increased particle bridging (Bagster 1993), leading to stronger aggregates

and a reduced breakage rate. However, the flocculant is gradually degraded through

time due to the repeated aggregation/breakage. This results in a reduced flocculant

activity and the gentle reduction in the aggregate size at extended residence times. If

the particles were coagulated, or if charge neutralisation (patch model) was the

dominant flocculation mechanism, a stable steady state aggregate size would be

produced (Bagster 1993) where the aggregation and breakage rates were balanced.

Higher flocculant dosages (40 & 80 g t-1) were also tried (not shown), increasing the

aggregate size and settling rate further, however the increased settling rate caused the

solid to settle prematurely in the pipe reactor (see Section 4).

Although the flow rate could have been increased, allowing the suspension of larger

aggregates formed at a higher flocculant dosage, this would also result in a higher shear

rate and increased aggregate breakage (Figure 6-2). Since flocculant is the major

operating cost in many units (Perry & Green 1997), and excessive flocculant addition

Page 163: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

162

can increase the sediment yield stress (Healy et al. 1994), reducing the fluid shear and

hence the flocculant requirement may lead to process improvements.

-1.5E-04

-1.0E-04

-5.0E-05

0.0E+00

5.0E-05

1.0E-04

1.5E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

0

2

4

6

8

10

12

14

16

Hin

dere

d se

ttlin

g ve

loci

ty (m

hr-1

)

Dosage = 20 g tֿ¹, (Baseline)Dosage = 10 g tֿ¹Dosage = 5 g tֿ¹

Figure 6-3: Effect of flocculant dosage on the development of the mean aggregate size and hindered settling velocity. Other conditions as per baseline.

6.2.3 Effect of suspension solid fraction

Figure 6-4 shows the effect of the solid fraction on the mean aggregate size and

corresponding hindered settling velocity. The decreased size at high solid fraction is

attributed (Sections 4 & 5) to an increase in fluid viscosity and energy dissipation rate,

leading to an increased breakage rate and smaller aggregates. The hindered settling

velocity is further reduced by greater hindrance at an increased solid fraction, as

described by Equation 6-2, although the effective thickener throughput is limited by the

solid settling flux given by:

sUφρ=ψ 6-6 where:

ψ = Settling flux (kg m-2s-1) U = Settling velocity (m s-1) φ = Solid volume fraction [0,1] ρs = Solid density (kg m-3)

Page 164: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

163

-1.5E-04

-1.0E-04

-5.0E-05

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

0

10

20

30

40

50

60

70

Hin

dere

d se

ttlin

g ve

loci

ty (m

hr-1

)

3.33 % w/v6.67 % w/v10.0 % w/v, (Baseline)13.33 % w/v16.67 % w/v

Figure 6-4: Effect of suspension solid fraction on the development of the mean aggregate size and hindered settling velocity. Other conditions as per baseline.

Page 165: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

164

6.2.4 Effect of primary particle size

Figure 6-5 shows the effect of primary particle size on the aggregate size and

subsequent hindered settling velocity. As described previously (Sections 4 & 5), a

larger primary particle has a lower surface area (per unit mass) and hence a higher

flocculant surface coverage. This leads to a stronger aggregate and an increased size

due to a reduction in the breakage rate. In addition, a larger primary particle also leads

to a lower aggregate porosity (Equation 6-8), further increasing the settling velocity.

-1.0E-04

-5.0E-05

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

2.5E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n ag

greg

ate

diam

eter

(m)

0

10

20

30

40

50

60

70

80

90

100

Hin

dere

d se

ttlin

g ve

loci

ty (m

hr-1

)

dp = 24.3 µm, (settling obs. in pipe)dp = 15.1 µmdp = 5.59 µm, (Baseline)dp = 3.47 µmdp = 2.36 µm

Figure 6-5: Effect of primary particle size (dp) on the development of the mean aggregate size and hindered settling velocity. Other conditions as per baseline.

6.2.5 Fractal geometry

Aggregate porosity is usually described by fractal geometry (Meakin 1988), where

although the total enclosed volume of the aggregate increases as a cube of its diameter,

its mass increases at some lower fractional (hence: fractal) power:

fD

p

agg

dd

m ⎟⎟⎠

⎞⎜⎜⎝

⎛∝ 6-7

where: m = Aggregate mass (kg) d = Diameter (agg = aggregate, p = primary particle) (m) Df = Mass-length fractal dimension The increase in porosity with aggregate size is usually rationalised on the basis of the

increasing voidage produced later in the aggregation process where large aggregates

result from aggregate-aggregate collision (Figure 2-7).

Page 166: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

165

Aggregate porosity has two major impacts on the hindered settling velocity. Firstly, it

dramatically reduces the aggregate density, reducing the driving force for the particle to

settle under gravity (Equation 6-1), giving (Gonzales & Hill 1998, Atteia 1998, Ellis &

Glasgow 1999, Manning & Dyer 1999):

( ) ( )3D

p

agglslagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ−ρ=ρ−ρ 6-8

where: ρ = Density(s = solid, l = liquid, agg = aggregate) (kg m-3)

Secondly, the effective suspension solid volume fraction is increased, increasing the

inter-particle hindrance as described by Richardson and Zaki (Equation 6-2) (Potanin &

Uriev 1991, Flesch et al. 1999):

fD3

p

aggseff d

d−

⎟⎟⎠

⎞⎜⎜⎝

⎛φ=φ 6-9

where: φ = Solid volume fraction (s = solid, eff = effective) [0,1] Combining equations 6-1, 2, 8 and 9, and substituting mean sizes gives:

( ) 65.4D3

p

aggs

3D

p

aggls

2agg

h

f

f

d

d1

18

d

dgd

U⎟⎟⎟

⎜⎜⎜

⎟⎟

⎜⎜

⎛φ−

μ

⎟⎟

⎜⎜

⎛ρ−ρ

=

6-10

The settling velocity may be further reduced by the effect of particle shape as the

particles deviate from the sphericity assumed by the above equations. However, the

effect of shape is relatively minor at low particle Reynolds number (∼ 0.7-1.1 times the

rate, Pettyjohn & Christiansen 1948, Heiss & Coull 1952, Happel & Brenner 1973,

Kousaka et al. 1981, Seville et al. 1997, Figure 2-17), and is partially offset by the

permeability of porous aggregates, allowing some fluid flow through the structure,

decreasing the drag slightly (Veerapaneni & Wiesner 1996, Gregory 1997). Hence, the

effects of particle shape, permeability, and Reynolds’ number have been ignored as

unnecessary complications at this stage, and Equation 6-10 incorporates the dominant

effects (suspension solid fraction and aggregate density) into Stokes’ Equation to

predict the hindered settling rate.

Page 167: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

166

In this case the fluid viscosity is taken to be a constant 0.00102 N s m-2 (water), and is

unaffected by the aggregation process. During hindered settling the particles are

essentially joined together as a continuous, permeable network, with the displaced water

passing through channels in the sediment. Previously (Sections 4 & 5) the overall

suspension viscosity was described as a function of the solid fraction and degree of

aggregation, however in that situation the flow is fundamentally different, with a high

degree of relative particle movement in the turbulent pipe flow.

Equation 6-10 was used to estimate the fractal dimension (Df) from the experimental

data via :

2

t,d,,,G Exp,h

106.Eq,hExp,h

Dp

f UUU

min ∑θφ

⎟⎟⎠

⎞⎜⎜⎝

⎛ −= 6-11

where: Uh,Exp = Experimental hindered settling rate (m s-1)

Uh,Eq.6-10 = Predicted hindered settling rate from Equation 6-10 (m s-1) In this case the fractal dimension (Df) was found to produce the best fit with a value of

2.42 (Figure 6-6), which is a typical value for aggregates formed by flocculation at high

shear rates (Table 2-3).

0.1

1

10

100

0.1 1 10 100

Experimental hindered settling velocity (m hr-1)

Pred

icte

d hi

nder

ed se

ttlin

g ve

loci

ty (m

hr-1

)

dp = 2.36 µm

dp = 3.47 µm

Baseline # 1

dp = 15.08 µm

dp = 24.3 µm

Dose = 5 g tֿ¹

Dose = 10 g tֿ¹

3.33 % w/v

6.67 % w/v

13.33 % w/v

16.67 % w/v

38.1 mm ID, 0.343 m sֿ¹

38.1 mm ID, 0.554 m sֿ¹

Baseline # 2

25.4 mm ID, 0.781 m sֿ¹

25.4 mm ID, 1.294 m sֿ¹

38.1 mm ID, 0.343 m sֿ¹, 10 g tֿ¹

dp = 15.0 µm, 6.67 % w/v

25.4 mm ID, 0.781 m sֿ¹, 40 g tֿ¹

Figure 6-6: Comparison between measured and predicted (Equation 6-10) hindered settling rates.

Aggregate settling data typically displays considerable scatter, however a sensitivity

analysis was performed to determine if the fractal dimension (Df) was a function of any

Page 168: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

167

of the experimental variables (shear rate G, flocculant dosage θf, primary particle size dp

and solid fraction φ). Both linear and non-linear sensitivities were tried as per

Equations 6-12 and 6-13, the linear according to:

⎟⎟⎠

⎞⎜⎜⎝

⎛φ

φ−φ+⎟

⎟⎠

⎞⎜⎜⎝

⎛ −+⎟⎟

⎞⎜⎜⎝

⎛θ

θ−θ+⎟⎟

⎞⎜⎜⎝

⎛ −+=

BL

BLExp

BL,p

BL,pExp,p

BL

BLExp

BL

BLExp*ff D

ddd

CBG

GGADD 6-12

with: G = Mean shear rate (s-1)

θ = Flocculant dosage (kg m-2) dp = Primary particle size (m)

φ = Solid fraction [0,1] Exp = Experimental condition BL = Experimental condition @ baseline Df

* = 2.43 A = 0.0244 B = 0.0842 C = -0.0060

D = -0.0786

and the non-linear:

z

BL

Expy

BL,p

Exp,px

BL

Expw

BL

Exp*ff d

dGG

DD ⎟⎟⎠

⎞⎜⎜⎝

φ

φ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

θ

θ⎟⎟⎠

⎞⎜⎜⎝

⎛= 6-13

with: Df

* = 2.44 W = 0.01949 X = 0.0337 Y = 0.00314 Z = -0.0337 Figure 6-7 shows the fit using the linear method (Equation 6-12), with the residual

reduced from 19.4 (Figure 6-6) to 11.9. The non-linear method (not shown, Equation

6-13) reduced the residual fractionally further, to 10.5. The improvement of Figure 6-7

over Figure 6-6 is minor, considering the number of degrees of freedom increased from

1 to 5, indicating that Equation 6-10 holds reasonably well within the experimental

scatter.

Page 169: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

168

0.1

1

10

100

0.1 1 10 100

Experimental hindered settling velocity (m hr-1)

Pred

iced

hin

dere

d se

ttlin

g ve

loci

ty (m

hr-1

)

dp = 2.36 µm

dp = 3.47 µm

Baseline # 1

dp = 15.08 µm

dp = 24.3 µm

Dose = 5 g tֿ¹

Dose = 10 g tֿ¹

3.33 % w/v

6.67 % w/v

13.33 % w/v

16.67 % w/v

38.1 mm ID, 0.343 m sֿ¹

38.1 mm ID, 0.554 m sֿ¹

Baseline # 2

25.4 mm ID, 0.781 m sֿ¹

25.4 mm ID, 1.294 m sֿ¹

38.1 mm ID, 0.343 m sֿ¹, 10 g tֿ¹

dp = 15.0 µm, 6.67 % w/v

25.4 mm ID, 0.781 m sֿ¹, 40 g tֿ¹

Figure 6-7: Comparison between measured and predicted (Equation 6-12) hindered settling rates.

The sensitivity analysis indicates some increase in the fractal dimension (less porous)

with shear and flocculant dosage, but a decrease with solid fraction. However, caution

may be required interpreting these observations, since none of the effects are very

strong (i.e. Df ∼ 2.4 ± 0.1) and do not change the overall observation that the initial

hindered settling rate is increased with flocculant dosage (Figure 6-3) and primary

particle size (Figure 6-5), but decreased with solid fraction (Figure 6-4) and shear rate

(Figure 6-2).

Figures 6-2, 3, 4 and 5 show a slight reduction in aggregate size on extended shearing,

typical of aggregates formed by polymer flocculants (Keys & Hogg 1978, Pelton 1981,

Williams et al. 1992, Bagster 1993). This effect could be interpreted as either aggregate

compaction, or partially irreversible breakage (or both) (Mills et al. 1991, Oles 1992,

Ellis & Glasgow 1999). However, aggregate breakage appears to be the dominant

mechanism for this system, because in all cases the hindered settling rates also drop

with aggregate size on extended shearing. If compaction was the dominant size

reduction mechanism, the hindered settling velocity would be expected to rise due to the

increased density (Equation 6-8) and reduced effective solid fraction (Equation 6-9).

Page 170: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

169

6.3 Conclusions The initial hindered settling velocities of flocculated calcite suspensions have been

studied under various process conditions (fluid shear, flocculant dosage, primary

particle size, solid fraction) and related to the on-line aggregate size measurements. As

would be expected a larger aggregate size leads to a higher hindered settling velocity,

however the effects of inter-particle hindrance must also be considered. Richardson and

Zaki’s (1955) relationship describing the hindered settling velocity of un-flocculated

suspensions has been adapted to flocculated suspensions by incorporating fractal

geometry to describe aggregate porosity. This has enabled the estimation of the fractal

dimension (2.4) from the experimental data.

Page 171: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

170

7. Dynamic optimisation and extrapolation to full-scale 7.1 Predicting settling behaviour –Dynamic optimisation Computer modelling may lead to process improvements through an increased

understanding of the system, however, in this case the primary objective is to allow the

prediction of optimal process conditions. These are typically determined by a model

optimisation; in this case performed within the software (gPROMS). However, care is

required in choosing the correct manipulated and objective variables. For example,

attempting to optimise the mean aggregate size as a function of flocculant dosage (note

Figure 6-3) would be meaningless, since the optimisation would rapidly converge to the

maximum allowable dosage.

An alternative optimisation might use the data from Figure 6-2 to predict the flow

conditions (i.e. shear rate, perhaps tapered) to give the largest aggregate in the shortest

possible time. However, aggregation at high solid fractions is fundamentally different

from, say, a slow chemical reaction, and sufficient time is probably already available for

complete aggregation. In this situation the optimisation would probably be better

performed to yield the largest possible aggregate, and the process changed (e.g. via a

change in the flocculant dosing point) to suit.

Another optimisation strategy would be to alter the feed solid fraction (e.g. by overflow

or underflow recycle). Figure 7-1 shows the effect of solid fraction on the predicted

average aggregate size through time. Figure 7-2 shows the variation in the maximum

mean aggregate size with solid fraction. A dynamic optimisation gave the peak at

4.32 % w/v. However, the primary objective of a thickener is to separate solid from

liquid, and the actual aggregate size is essentially irrelevant (accepting that larger

aggregates typically lead to a higher settling rate) and the objective function should be

written in terms of the solid settling rate, rather than the aggregate size. Figure 7-3

shows that the maximum hindered settling velocity is predicted to occur at a lower

3.32 % w/v for this system. The maximum hindered settling velocity occurs at a lower

solid concentration because the hindered settling rate is also a function of the solid

fraction (Equation 6-2).

Page 172: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

171

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 5 10 15 20 25 30 35 40

Time (s)

Mea

n A

ggre

gate

size

(m)

1.00 % w/v1.93 % w/v3.32 % w/v4.32 % w/v5.18 % w/v6.67 % w/v10 % w/v, (Baseline)

Figure 7-1: Modelled mean aggregate size as a function of time and solid fraction.

0.0E+00

5.0E-05

1.0E-04

1.5E-04

2.0E-04

0 10 20 30 40 50 60

Solid concentration (% w/v)

Peak

pre

dict

ed m

ean

aggr

egat

e si

ze (m

)

Peak @ 4.32 %

Figure 7-2: Modelled maximum aggregate size as a function of the solid fraction. The maximum aggregate size occurs somewhere between 5-10 seconds after flocculant addition, see Figure 7-1.

Page 173: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

172

0.0E+00

5.0E-04

1.0E-03

1.5E-03

2.0E-03

2.5E-03

3.0E-03

0 10 20 30 40 50 60

Solid concentration (% w/v)

Peak

pre

dict

ed h

inde

red

sett

ling

velo

city

(m s-1

)

Peak @ 3.32 %

Figure 7-3: Modelled maximum hindered settling velocity as a function of the solid fraction.

While Figure 7-3 indicates that the highest initial hindered settling velocity occurs at

3.32 % w/v for this system, the total solid throughput is of primary interest for a

thickener (with boundary values for acceptable overflow clarity and underflow solid),

hence the solid mass settling flux rate (kg m-2s-1 or t m-2hr-1) is calculated as the product

of the settling rate and the solid mass fraction (Equation 6-6) (Figures 7-4 & 7-5) with

the optimal value of 5.18 % w/v, occurring after 9.79 s (Figure 7-4).

If the optimisation is static rather dynamic, the results change fractionally. For

example, if the residence time in the pipe reactor is limited to 3 s, then the predicted

optimal solid fraction becomes 6.4 % w/v, because a higher solid fraction leads to a

faster initial aggregation rate via an increased collision rate. Alternatively, if the mean

residence time is taken to be 30 s, the optimal feed solid decreases fractionally to 4.8 %

w/v, due to a decrease in the polymer degradation at lower solid fraction.

Page 174: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

173

0

0.02

0.04

0.06

0.08

0.1

0.12

0 5 10 15 20 25 30 35 40

Time (s)

Solid

flux

rat

e (k

g m

-2s-1

)

4.80 % w/v5.18 % w/v6.40 % w/v

Figure 7-4: Effect of residence time on maximum initial settling flux.

0.0E+00

2.0E-02

4.0E-02

6.0E-02

8.0E-02

1.0E-01

1.2E-01

0 10 20 30 40 50 60

Solid concentration (% w/v)

Peak

pre

dict

ed fl

ux r

ate

(kg

s-1m

-2) Peak @ 5.18 %

Throughput limiting solid fraction

Figure 7-5: Modelled maximum flux rate as a function of the solid fraction. The maximum flux rate occurs between 5-10 seconds, see Figure 7-4.

The optimisation described above maximises the initial settling flux, since only the

initial hindered settling velocity was collected during the experimental sessions.

However, a thickener is unlikely to be limited by the initial settling flux, and Figure 7-5

includes the required settling flux (operating line) to give a dynamic form of the

Yoshioka/Hasset combined flux curve (see Figure 2-21). The line is drawn tangentially

to the flux curve such that it also passes through the desired underflow solid

concentration (taken as 50 % w/v here). The y-axis intercept gives the upper limiting

feed flux for the thickener (i.e. about 6 × 10-2 kg m-2s-1 in this case). The limiting solid

Page 175: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

174

concentration is about 28 % in this case, a Kynch settling layer that would occur lower

in hindered settling region.

Although Figure 7-5 has the conventional form, the flux curve was produced by running

the population balance simulation with a range of initial solid fractions (as per the Coe

and Clevenger experimental method). The combined flux curve is usually constructed

from the analysis of a single batch-settling test (Section 2.7.2, Kynch 1952, Talmage &

Fitch 1955), avoiding the effect of solid fraction on the aggregation process.

The combined flux curve makes various assumptions based on Kynch theory

(Section 2.7.2), significantly that the thickener is limited by the hindered settling rate.

If, in fact, the compaction/bed region is rate limiting, then some alternative model is

required (Svarovsky 2000).

7.2 Extrapolation to Industrial scale Fluid dynamics are notoriously difficult to scale accurately, and this work is no

exception. Industrial flows are characterised by large geometries and hence high

Reynolds number. The size of the macro-scale turbulent eddies gives them a high

tangential velocity, enabling the suspension of large, rapidly settling particles, but at a

comparatively low fluid shear rate. Wastewater treatment plants typically aim for a

mean shear rate of ∼ 60 s-1 (Amirtharajah et al. 1991).

However, it is difficult to replicate these conditions on a small scale. The reduced

geometry leads to a lower Reynolds number and the tendency for the solid to settle out

in the experimental apparatus. The Reynolds number can be increased by increasing the

flow velocity, but that would also increase the shear rate, causing irreversible aggregate

breakage (Figure 6-2). Alternatively, solid settling can be suppressed by using less

dense, and/or smaller primary particles.

Figure 7-6 shows the development of the mean aggregate size predicted by the

population balance, with appropriate conditions changed to better reflect the industrial

scale. Primary particle size distributions vary considerably depending on the process

(Pearse 1977, Perry & Green 1997), varying from μm to mm scale, with the product of

Page 176: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

175

grinding circuits typically reasonably coarse (∼ 10-1000 μm) compared to crystallisers

(∼ 0.1-100 μm). The primary particle distribution was altered to give a mean primary

particle size of 80 × 10-6 m (80 μm), and the flow velocity reduced to give a (assumed

still turbulent) shear rate of about 40-50 s-1.

The increased primary particle size also lead to an increased effective dosage via a

reduction in the solid surface area, but the absolute dosage was kept at 20 g t-1.

Figure 7-6 shows that under these conditions the final aggregate size is in the range

1000 × 10-6 m (1000 μm); approximately correct for full-scale.

0.0E+00

2.0E-04

4.0E-04

6.0E-04

8.0E-04

1.0E-03

1.2E-03

0 10 20 30 40 50 60 70 80 90 100

Time (s)

Figure 7-6: Modelled change in mean aggregate size under simulated plant conditions – see text.

It should be noted that the conditions assumed in the population balance used to

generate Figure 7-6 are not actually achievable experimentally in the pipe reactor. The

coarse solid would rapidly settle out of the flow, which would have become laminar by

such a low flow rate.

Figure 7-7 shows the resulting settling rate and flux rates, which also appear to be

plausible for the full-scale plant, accepting that the flux rate is the initial settling flux,

not the limiting flux (see discussion around Figure 7-5). It is also interesting to note

that the model predicts very little polymer degradation at a low shear rate, with the

settling rate effectively constant after the first 10 s.

Page 177: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

176

0

10

20

30

40

50

60

70

80

90

100

0 10 20 30 40 50 60 70 80 90 100

Time (s)

Hin

dere

d se

ttlin

g ra

te (m

hr-1

)

0

2

4

6

8

10

12

14

Solid

flux

rat

e (t

hr-1

m-2

)

Hindered settling rateSolid flux rate

Figure 7-7: Modelled hindered settling rate and solid flux rates under simulated plant conditions – see text & Figure 7-6.

7.3 Conclusions This section has briefly investigated how the population balance and associated settling

model can be used to optimise full-scale thickener operation, in particular by altering

the solid fraction to give optimal aggregation to produce the maximum settling flux.

The optimisations were also dynamic, as the values used represented the maximum

values achieved during the aggregation process (around 5-10 seconds in this case).

Ultimately, the model will be used as a sub-model within the full CFD model of a

thickener, allowing the scope of the optimisation to be expanded considerably. In that

case the feedwell design (feed entry velocity, flocculant dosing point(s), baffle size and

position etc) will be able to be optimised to give the best shear profile and residence

time. The optimal feed solid fraction can also be determined, giving the highest settling

flux and hence unit throughput.

Page 178: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

177

8. Conclusions and future work

8.1 Conclusions The aggregation/breakage kinetics of calcite flocculation in turbulent pipe flow have

been modelled by population balance. The model draws heavily on successful work

previously described in the literature, but extends it by combining various aspects into a

single model that describes both the aggregate size and settling rate as a function of the

key process variables. The model is robust and stable, and although it adequately

describes complex aggregation and settling behaviour, it remains simple enough to form

a sub-model within a larger CFD simulation of a thickener/clarifier.

The population balance model is based on the descriptions by Hounslow et al. (1988)

and Spicer and Pratsinis (1996a), with the collision rate described by Saffman and

Turner’s (1956) turbulent shear kernel. However, Saffman and Turners kernel

(Equation 2-7) considerably overestimates the initial aggregation rate immediately after

flocculant addition, and a capture efficiency (α) term is introduced to account for the

discrepancy. If the capture efficiency was treated simply as a fitted parameter, it had to

be relatively small (α ∼ 0.06) to give a reasonable fit to the experimental data,

suggesting that particle/particle collision is not initially rate limiting.

Experimental evidence (Figure 5-3) shows that changing the initial

flocculant/suspension mixing (e.g. by flocculant stream dilution and/or velocity)

significantly changed the initial aggregation rate, suggesting that the initial capture

efficiency was a function of the macro-scale flocculant/suspension mixing. In this case,

the capture efficiency (Equation 5-6) is taken to be zero before flocculant addition

(stable suspension), and rapidly (Figure 5-2) approach unity as the

flocculant/suspension become well mixed. Equation 5-6 is a relatively crude

approximation of the overall flocculant mixing, collision and adsorption process. In this

case, no account is taken for polymer/particle collision or adsorption, as they are taken

to be rapid (Gregory 1993, Bagster 1993, Mühle 1993, Hogg 1999) compared to the

bulk macro-scale flocculant/suspension mixing. Macro-scale mixing is likely to be

highly scale-dependent, and much slower on the full scale. Industrial-scale mixing is

difficult to replicate on a laboratory scale, although emerging techniques like

computational fluid dynamics (CFD) (Lainé et al. 1999, Ducoste & Clark 1999, Farrow

Page 179: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

178

et al. 2000), or electrical impedance tomography (Brown et al. 1985, Webster 1990,

Salkeld 1991, Dickin et al. 1992, Williams & Simons 1992) are likely to increase the

understanding of this area in the future.

Equation 5-6 suggests that the capture efficiency approaches unity as the flocculant

becomes well mixed. This is unlikely to be the case in practice, and the capture

efficiency is likely to be a function of the hydrodynamic interaction (Equation 2-17) and

the flocculant dosage (Equation 2-19). However, in this case the model was made

dosage dependent via the breakage kernel (Equation 5-19), where a higher surface

coverage gave a stronger aggregate, less breakage, and hence a larger final aggregate

size. The effective active flocculant surface coverage was taken to decrease through

time due to the repeated aggregation/breakage process. The flocculant is degraded due

to polymer chain scission or rearrangement, and is shown experimentally (Figures 4-5,

4-9, 4-11 & 4-13) as a gentle reduction in the aggregate size at extended residence

times. Flocculant degradation is clearly undesirable industrially, but may occur around

feed or flocculant entry points, baffles etc. Although sufficient fluid shear is required to

provide effective mixing, excessive shear results in irreversible aggregate breakage.

In addition to fluid shear (G) and flocculant dosage (θf), the effects of the suspension

solid volume fraction (φ) and primary particle size (dp) were also considered. These

may also be process variables depending on the system (feed, milling, crystallisation

etc), and the solid fraction can be altered in some units via overflow or underflow

recycle. The final aggregate size was found to increase with primary particle size, and

although a full description is likely to be a complex function of the primary particle

packing and bonding within the aggregate structure, in this case a reasonable

relationship was found based on the effective flocculant coverage (Equation 5-18).

That is, a larger primary particle has a lower surface area (per unit mass) and therefore a

higher flocculant surface coverage, increased strength, and hence a larger final

aggregate size.

The effect of solid fraction has received little attention in the literature due to the

previous difficulty of on-line aggregate sizing in concentrated suspensions. Also, most

workers have considered coagulation in river estuary systems or water treatment

clarifiers, which are characterised by a low solid fraction. At low solid fraction (< 1 %),

Page 180: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

179

there is evidence (Thomas 1964, de Boer et al. 1989, Kobayashi et al. 1999) to suggest

that the aggregate size increases with solid fraction, probably reflecting an increased

particle collision rate. This is consistent with the aggregation and breakage kernels,

because aggregation (Equation 2-7) is taken to be second order with respect to particle

number, whereas breakage (Equation 5-19) is first order. That is, doubling the number

of primary particles doubles the breakage rate, but quadruples the aggregation rate,

leading to a higher degree of aggregation.

However, at the high solid fractions (3.33-16.7 % w/v) studied here, the final aggregate

size is reduced with solid fraction. This is attributed to the increased suspension

viscosity at high solid fraction (Equation 5-16), leading to a higher energy dissipation

rate (flow rate fixed by positive displacement pump) and increased breakage. Breakage

kernels (Table 2-6) are typically written in terms of G (as per aggregation kernels),

suggesting that the breakage rate would decrease at higher viscosity (Equation 5-7).

However, most workers have previously considered systems with a solid fraction that is

so low that the viscosity is effectively unchanged from water. In this case, during

model construction the breakage kernel was initially taken as ∝ ∈yμz where y and z

were fitted parameters. Parameter z repeatedly converged to 1 ± 0.05 and was

subsequently set = 1, removing a degree of freedom from the model. Parameter y was

typically in the range 0.7 ± 0.1 (final estimate = 0.677, perhaps Si ∝ ∈2/3).

The four model parameters were estimated from experimental data collected in a pipe

reactor. A range of pipe diameters, lengths and suspension flow rates gave a range of

mean shear rates (G) and residence times. The shear rate and suspension viscosity were

calculated from the pressure drop along the pipe reactor, as measured by a manometer

bank. The manometer showed an increase in fluid viscosity (larger ΔP) as a function of

the aggregate size, and also served to confirm the lower limiting flow rate when the

solid settled out in the pipe reactor. The relatively high mean shear rate (∼ 100-1000 s-1

here) in small-scale turbulent pipe flow is an incentive to reduce the flow rate, but is

ultimately limited by the transition to laminar flow, with poor mixing and solid

stratification/settling in the pipe.

A sparse matrix (Figure 4-4, Table 4-1) of experimental runs was performed by

changing the process variables (G, φ, θf and dp) independently away from a common

Page 181: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

180

baseline. Some of the data were of limited value due to settling, but the remaining data

were used for model parameter estimation. Several additional runs were performed

(Figure 5-17) in the gaps of the experimental matrix by changing two process variables

simultaneously. These data were not used for parameter estimation, but were used to

confirm the predictivity of the model.

In addition to the on-line, in-pipe particle sizing (FBRM), hindered settling

measurements were taken using a vertical settling column (Figure 6-1). This allowed a

relationship to be found between the mean aggregate size and the hindered settling

velocity (Equation 6-10), ultimately allowing the population balance to predict the

initial settling flux. The relationship is based on Richardson and Zaki’s (1955)

extension of Stokes’ (1851) law for particulate settling in viscous flow. A variety of

other factors also influence the hindered settling velocity (aggregate shape and

permeability, fluid inertial effects at higher particle Reynolds number), but the

dominant effect is due to the increased volume (and reduced density) of porous

aggregates. Aggregate porosity was described with fractal geometry, allowing the

fractal dimension to be estimated (Df = 2.4) from the experimental data (Equations 6-10

& 6-11).

Although this work has provided some insight into the size of aggregates ultimately

produced by the competing effects of aggregation and breakage, the initial macro-scale

mixing of the flocculant/suspension is described in a simplified form appropriate for

only turbulent pipe flow and requires further investigation for other geometries like

feedwells. The practical importance of macro-scale mixing is evidenced by various

industrial design practices to encourage efficient flocculant mixing (e.g. flocculant

stream dilution, multi-point addition, baffles, in-line mixers etc).

There is also a need to consider the down-stream effects of changes to the flocculation

process. Although the initial hindered settling velocity has been investigated here, the

subsequent compaction of the sediment will be important in many applications. For

example, increasing the flocculant dosage to increase the throughput may not be viable

if it produces a voluminous sediment with a high yield stress that limits the final

underflow solid concentration. Issues may also arise further down-stream, and the

Page 182: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

181

filterability and pumping requirement of the underflow may also be compromised (or

improved) by changes to the flocculation conditions.

8.2 Suggested future research 1. Code the population balance as a sub-model into a CFD model describing a

mineral processing thickener/clarifier. This was the overall aim of the work

presented here, and should allow process improvements (throughput, clarity,

perhaps also underflow solid) through improved flocculation. The crude

flocculant mixing/adsorption described will require replacing by a 3-dimensional

description provided by the CFD model.

2. Expand the aggregation kinetics experimental using the pipe reactor to other feed

materials, flocculant types, liquor temperature/chemistry etc. Fit the population

balance model to these new data sets, and build up a matrix of parameter values of

different systems. Ultimately the parameters could be replaced by functions

accounting for the differences between the various systems.

3. Further investigate the relationship between the aggregation state and the

subsequent sedimentation, in particular yield stress and sediment dewaterability,

this would greatly increase the ability to use the population balance model to

improve the unit performance. Similarly, the link between aggregation and

overflow clarity should be investigated, allowing the optimisation of units limited

by overflow clarity.

4. Various other techniques could be used to further characterise the

aggregation/breakage process, in particular the use of CFD to investigate

fluid/particle interactions on a small scale, e.g. the hydrodynamic force that acts

between colliding particles, or the action of fluid shear to break aggregates.

Micromechanical devises might be used to further understand the forces of

attraction/repulsion between particles, and the influence of flocculants on these

forces.

Page 183: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

182

9. Notation A = Area (m2) A = Hamaker constant Ab = Thickener area from batch experiment (m2) As = Surface area of solid (m2) ai = Radius of the ith particle (m) BL = Experimental condition at baseline C = Floc strength coefficient C(x) = Concentration at point x (or m = mean) CA = A/36πμGa3 Cd = Coefficient of drag (dimensionless) Ci,A = Geometric average length of ith channel (μm) = (Ci,u × Ci,l)1/2 Ci,l = Length of the lower boundary of the ith chord channel (μm) Ci,u = Length of the upper boundary of the ith chord channel (μm) Cm = Mean concentration through vessel CN = Drag coefficient in Newton’s equation D = Units of water per unit solid D∞ = D at infinite time D = Pipe diameter (m) Df = Mass-length fractal dimension (1 [thin rods] ≤ Df ≤ 3 [spheres etc]) d = Separation distance d = Particle diameter (m) dp = Diameter of the primary particle (m) dagg = Diameter of aggregate (m) dagg,i = Diameter of the ith sized aggregate (m) dagg,Exp = Experimental (FBRM) mean aggregate size (m) dagg,PB = Modelled (population balance) mean aggregate size (m) dm = Mass equivalent diameter (m) dm,i = Mass effective diameter of the ith sized aggregate (m) Exp = Experimental condition F = Force (N or kg m s-2) F = Solid feed rate (m3s-1) Fb = Force of buoyancy (N or kg m s-2) Fd = Force on particle due to fluid drag (N, or kg m s-2) Fg = force of gravity (N or kg m s-2) f = Fanning friction factor (dimensionless) G = Mean turbulent shear rate (s-1) g = Acceleration due to gravity (9.8 m s-2) H = Expected full scale compression zone depth (m) H = Mudline height (m) H = Height of suspension (Ho at t = 0) HK = Height of Kynch isoconcentration line. K = Boltzmann’s constant (1.38 × 10-23 J K-1) K = Velocity head KB = Floc breakage coefficient KE = Einstein’s constant (2.5) K = Constant k = Constant k1 = Fitted parameter # 1 = 0.3431 k2 = Fitted parameter # 2 = 38.1 k3 = Fitted parameter # 3 = 0.677 k4 = Fitted parameter # 4 = 1224 L = Length (m) L = Characteristic aggregate length (m) L = Pipe length (m) Li = Length of the ith channel

Page 184: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

183

M = Mixing index [0,1] m = Exponent m = Mass (kg) mf = Mass of flocculant (kg) mi = Mass of ith aggregate (kg) ml,i = Mass of liquid in the pores of aggregate, size i (kg) mp = Mass of primary particle (kg) ms = Mass of solid in aggregate (kg) N = Number of particles Ni = Number of ith sized particles (m-3) Ni = Number of i sized chord counts observed (s-1) Ni,n = Number of n-weighted chord counts in the ith channel (μmn s-1) Ni,o = Raw un-weighted counts in the ith channel (s-1) Np

* = Number of pth sized particles predicted (s-1) Nt = Total observed particle counts (s-1) Nt

* = Corrected total particle counts (s-1) n = Exponent P = Pressure (N m-2) P = Fitted parameter(s) Py(φ) = Compressive yield strength of the sediment (N m-2) Re = Reynolds number (dimensionless) Si = Breakage rate of the ith sized particle (s-1) Sv = specific surface area of solid per unit volume of particles (m2m-3 = m-1) T = Absolute temperature (K) t = Time (s) tc = Time required to reach required sediment concentration (days) tm = Metric Tonnes (1000 kg) U = Velocity (m s-1) Uf = Feed settling velocity (m s-1) Uh = Hindered settling velocity (m s-1) Uh,Eq.6-10 = Predicted hindered settling rate from Equation 6-10 (m s-1) Uh,Exp = Experimental hindered settling rate (m s-1) Ul = Settling velocity of limiting concentration (m s-1) UN = Settling velocity given by Newton’s equation (m s-1) Uo = Free settling rate of the particle (m s-1) Ur = Settling velocity required to prevent solid reaching the overflow (m s-1) Us = Settling velocity given by Stokes’ equation (m s-1) Uu = Settling velocity due to underflow removal (m s-1) u = Velocity component in the x direction (m s-1) V = Mean flow velocity (m s-1) VB = Velocity of the bulk pipe flow (m s-1) VF = Velocity of the flocculant stream (m s-1) Vi = Volume of the particles in the ith size range (m3) Vo = Initial volume of the suspension (m3) Vsed = Equilibrium volume of the sediment (m3) Vt = Interaction energy VT = Total suspension volume (m3) vx = Instantaneous velocity of component x (m s-1) v = Velocity component in the y direction (m s-1) v = Volume of particle (m3) vi = Volume of ith particle

xv = Time averaged velocity (m s-1) 'xv = X-component velocity fluctuation (m s-1)

Wp,i = Width of pth particle giving chords in the ith channel (μm) w = Velocity component in the z direction (m s-1) waccel. = Relative particle velocity of particles due to acceleration (inertial effect) wshear = Relative particle velocity of particles due to fluid shear. z = Height of compression zone (m)

Page 185: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

184

z = Valency z = Elevation (m) z∞ = Height of compression zone at infinite time (m) zc = Height of compression zone at critical time (m) α = Capture efficiency [0,1] α = Velocity profile factor β = Ratio of apparent to actual volume fraction βij = Rate of collision between i and j sized particles (aggregation kernel) (m3s-1) ∈ = Energy dissipation rate per unit mass (J s-1kg-1, m2s-3) Φ = Power input per unit volume (kg m-1s-3) θ = Flocculant surface coverage (kg m-2) σ = Concentration variation σo = Concentration variation at time = 0 η = Dynamic viscosity (N s m-2) η = Kolmogoroff micro-scale of length (m) κ = Deybe-Huckel parameter κ = Permeability (m2) κ(φ) = Dynamic compressibility π = Pi (∼3.14) ε = Porosity, i.e. liquid fraction (1-φ) (m3m-3) Ω = Relative drag of an aggregate compared to solid sphere of same diameter τi = Particle relaxation time (s) τk = Kolmogoroff micro-scale of time (s) Γjk = Breakage distribution function ΔP = Pressure drop along pipe (N m-3) μs = Suspension viscosity (N s m-2) μs = Viscosity of suspension (kg m-1s-1, N s m-2) μo = Viscosity of water (1.02 × 10-3 N s m-2) φ = Solid volume fraction (1-ε) (m3m-3) φo = Initial solid fraction (m3m-3) φu = Underflow solid fraction (m3m-3) φm = Maximum solid fraction, (taken as 0.65) φeff = Effective solid fraction φl = Limiting solid fraction (m3m-3) φs = Actual suspension solid volume fraction [0,1] φ∞ = Solid fraction after infinite time (m3m-3) φcrit = Critical solid fraction that limits thickener throughput (m3m-3) φf = Feed solid fraction (m3m-3) φg = Solid fraction where gelation (compression) begins (m3m-3) φ(ε) = Function of ε φ(z) = Volume fraction at position z ρ = Density (kg m-3) ρs = Density of the solid (calcite = 2710 kg m-3) ρeff = Effective aggregate density (kg m-3) ρl = Density of the liquid (water = 1000 kg m-3) ρsl = Density of slurry (kg m-3) Θ = Flocculant degradation index [0,1] λ = Characteristic wavelength (m) θf = Effective flocculant surface coverage (kg m-2) γ1 = Dimensionless function of surface charge γ = Laminar shear rate (s-1) ν = Kinematic viscosity (m2s-1) ψ = Sphericity, ratio of surface area of equivalent sphere to actual surface area ψ = Settling flux (kg m-2s-1) ψm = mass settling flux (kg h-1m-2)

Page 186: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

185

10. References Abel, J.S., Stangle, G.C. and Schilling, C.H. 1994, ‘Sedimentation in flocculating colloidal suspensions’, J. Mater. Res., Vol. 9, No. 2, pp. 451-461. Abrahamson, J. 1975, ‘Collision rates of small particles in a vigorously turbulent fluid’, Chem. Eng. Sci., Vol. 30, pp. 1371-1379. Adachi, Y. and Kamiko, M. 1994, ‘Sedimentation of a polystyrene latex floc’, Powder Technology, Vol. 78, pp. 129-135. Adachi, Y. and Ooi, S. 1990, ‘Geometrical structure of a floc’, Journal of Colloid Interface Science, Vol. 135, pp. 374-384. Adachi, Y. and Tanaka, Y. 1997, ‘Settling velocity of an aluminium-kaolinite floc’, Wat. Res., Vol. 31, No. 3, pp. 449-453. Adachi, Y. and Kamiko, M. 1994, ‘Sedimentation of a polystyrene latex floc’, Powder Technology, Vol. 78, pp. 129-135. Adachi, Y., Stuart, M.A.C. and Fokkink, R. 1994, ‘Dynamic aspects of bridging flocculation studied using standardized mixing’, Journal of Colloid Interface Science, Vol. 167, pp. 346-351. Adachi, Y., Stuart, M.A.C. and Fokkink, R. 1994, ‘Kinetics of turbulent coagulation studied by means of end-over-end rotation’, Journal of Colloid Interface Science, Vol. 165, pp. 310-317. Akers, R.J., Rushton, A.G. and Stenhouse, J.I.T. 1987, ‘Floc breakage: The dynamic response of the particle size distribution in a flocculated suspension to a step change in turbulent energy dissipation’, Chem. Eng. Sci., Vol. 42, No. 4, pp. 787-798. Alder, P.M. 1981, ‘Streamlines in and around porous particles’, Journal of Colloid Interface Science, Vol. 81, pp. 531-535. Alfano, J.C., Carter, P.W. and Whitten, J.E. 1999, ‘Use of scanning laser microscopy to investigate microparticle flocculation performance’, Journal of Pulp and Paper Research, Vol. 25, No. 6, pp. 189-195. Alfano, J.C., Carter, P.W. and Whitten, J.E. 2000, ‘Polyelectrolyte-induced aggregation of microcrystalline cellulose: reversibility and shear effects’, Journal of Colloid and Interface Science, Vol. 223, pp. 244-254. Allain, C., Cloitre, M. and Parisse, F. 1996, ‘Settling by cluster deposition in aggregating colloidal suspensions’, Journal of Colloid and Interface Science, Vol. 178, pp. 411-416. Allen, H.S. 1900, ‘The motion of a sphere in a viscous fluid’, London, Edinburgh and Dublin Philosophical Magazine, and Journal of Science, Vol. 50, pp. 323-338, 519-534. Allen, T. 1990, Particle Size Measurements, Chapman and Hall. Amal, R., Coury, J.R., Walsh, W.P. and Waite, T.D. 1990a, ‘Structure and kinetics of aggregating haematite’, Colloid and Surf., Vol. 46, pp. 1-19. Amal, R., Raper, J.A. and Waite, T.D. 1990b, ‘Fractal structure of hemetite aggregates’, Journal of Colloid Interface Science, Vol. 46, pp. 1-19.

Page 187: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

186

Amirtharajah, A., Clark, M.M. and Trussell, R.R. 1991, Mixing in Coagulation and Flocculation, American Water Works Association Research Foundation, USA. Anderson, J.E. 1996, ‘Evaluation of the Lasentec Laser microscope for monitoring refining’, Papermakers Conference, pp. 225-235. Anon. 1989, Slurry Pumping Handbook, Warman. Argaman, Y. and Kaufman, W.J. 1968, ‘Turbulence in orthokinetic flocculation’, SERL Report No. 68-5, Sanitary Engineering Research Laboratory, University of California, Berkeley. Argaman, Y. and Kaufman, W.J. 1970, ‘Turbulence and flocculation’, Journal of the Sanitary Engineering Division, ASCE, Vol. 96 (SA2), pp. 223-241. Atteia, O. 1998, ‘Evolution of size distributions of natural particles during aggregation: modelling versus field results’, Colloids and Surfaces A - Physicochemical and Engineering Aspects, Vol. 139, No. 2, pp. 171-188. Aubert, C and Cannell, DS 1986, ‘Restructuring of colloidal silica aggregates’, Phys. Rev. Lett., Vol. 56, pp. 738-741. Bagster, D.F. 1993, ‘Aggregate behaviour in stirred vessels’, in Shamlou, P.A. (Ed.), Processing of solid-liquid separations, Butterworth-heinemann LTD, UK, pp. 26-58. Bajpai, A.K. and Bajpai, S.K. 1995a, ‘Dynamic measurement of adsorption of hydrolysed polyacrylamide (HPAM) onto hematite’, Colloid Polym. Sci., Vol. 273, pp. 1028-1032. Bajpai, A.K. and Bajpai, S.K. 1995b, ‘Kinetics of polyacrylamide adsorption at the iron oxide-solution interface’, Colloids Surf. A: Physicochem. Eng. Aspects, Vol. 101, pp. 21-28. Bale, A.J and Morris, A.W. 1987, ‘In situ measurement of particle size in estuarine waters’, Estuar. Coast. Shelf Sci., Vol. 24, pp. 253-263. Bale, A.J. 1996, ‘In situ laser optical particle sizing’, J. Sea Research, Vol. 36, pp. 31-6. Barker, G.C. and Grimson, M.J. 1990, ‘Theory of sedimentation in colloidal suspensions’, Colloids and Surfaces, Vol. 43, pp. 55-66. Barnes, H.A. and Holbrook, S.A. 1993, ‘High concentration suspensions: preparation and properties’, in Shamlou, P.A. (Ed.): Processing of Solid-Liquid Separations, Butterworth-Heinemann LTD, UK. Barrett, P. and Glennon, B. 1999, ‘In-line FBRM monitoring of particle size in dilute agitated suspensions’, Part. Part. Syst. Charact., Vol. 16, pp. 207-211. Barth, H.G. and Sun, S-T. 1991, ‘Particle size analysis’, Analytical Chemistry, Vol 63, pp. 1R-10R. Batterham, R.J., Hall, J.S. and Barton, G. 1981, ‘Pelletizing kinetics and simulation of full scale balling circuits’, 3rd International Symposium on Agglomeration, pp. A 136-A151. Becker. R., 1999, Personal Communication. Bemer, G.G. and Zuiderweg, F.J. 1980, ‘Growth regimes in the spherical agglomeration process’, Fine Part. Process., Proc. Int. Symp., Vol. 2, pp. 1524-1546. Berlin, A.A. and Kislenko, V.N. 1995, ‘Kinetic models of suspension flocculation by polymers’, Colloids and Surfaces A - Physicodhemical and Engineering Aspects, Vol. 104, pp. 67-72. Berlin, A.A., Kislenko, V.N. and Solomentseva, I.M. 1998, ‘Mathematical modelling of the suspension flocculation by polyelectrolytes’, Colloid Journal, Vol. 60, No. 5, pp. 592-597.

Page 188: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

187

Berlin, A.A., Solomentseva, I.M. and Kislenko, V.N. 1997, ‘Suspension flocculation by polyelectrolytes: experimental verification of a developed mathematical model’, Journal of Colloid and Interface Science, Vol. 191, No. 2, pp. 273-276. Bergström, L., Schilling, C.H. and Aksay, I.A. 1992, ‘Consoliodation behaviour of flocculated alumina suspensions’, J. Amer. Ceram. Soc., Vol. 75, No. 12, pp. 3305-3314. Bhatty, J.I., Davies, L. and Dollimore, D. 1982, ‘The use of hindered settling data to evaluate particle size or floc size, and the effect of particle-liquid association on such sizes’, Surface Technology, Vol. 15, pp. 323-344. Bird, R.B., Stewart, W.E. and Lightfoot, E.N. 1960, Transport Phenomena, John Wiley and Sons, USA. Blanco, A., Negro, C., Hooimeijer, A. and Tigero, J. 1996, ‘Polymer optimization in paper mills by means of a particle size analyser: an alternative to zeta potential measurements’, Appita, Vol. 49, No. 2, pp. 113-116. Blasius, H. 1913, ‘Das Ähnlichkeitsgesetz bei Reibungsvorgängen in Flüssigkeiten’, Forschn. Gebiete Ingenierw., No. 131. Blunt, M. 1989, ‘Hydrodynamic force distribution on a fractal cluster’, Phys. Rev. Chem., Vol. 39, pp. 5801-5806. Bongartz, D., Schneider, A. and Hesse, A. 1999, ‘Direct measurement of calcium oxalate nucleation with a laser probe’, Urol. Res., Vol. 27, pp. 135-140. Brakalov, L.B. 1987, ‘A connection between the orthokinetic coagulation capture efficiency of aggregates and their maximum size’, Chem. Eng. Sci., Vol. 42, pp. 2373-2383. Brinkman, H.C. 1948, ‘On the permeability of media consisting of closely packed porous particles’, Appl. Sci. Res. Vol. A1, pp. 81-86. Brown, B.H., Barber, D.C. and Saegar, A.D. 1985, ‘Applied potential tomography: possible clinical applications’, Clin. Phy. Physiol. Meas., Vol. 6, Pt II, 109-121. Brunk, P.B., Kock, D.L. and Lion, L.W. 1998, ‘Turbulent coagulation of particles’, J. Fluid Mechanics, Vol. 364, pp. 81-113. Burban, P., Lick, W. and Lick, J. 1989, ‘The flocculation of fine-grained sediments in estuarine waters’, J. Geophys. Res., Vol. 94, pp. 8323-8330. Burger, R. and Concha, F. 1998, ‘Mathematical model and numerical simulation of the settling of flocculated suspensions’, International Journal of Multiphase Flow, Vol. 24, pp. 1005-1023. Burgers, J.M. 1941, ‘Hydrodynamics – On the influence of the concentration of a suspension upon the sedimentation velocity’, Proc. Koninkl. Adad. Wetenschap., Amsterdam, Vol. 42, pp. 1045-1051. Burgers, J.M. 1942, ‘Hydrodynamics – On the influence of the concentration of a suspension upon the sedimentation velocity’, Proc. Koninkl. Adad. Wetenschap., Amsterdam, Vol. 45, pp. 9-16. Buscall, R. and White, L.R. 1987, ‘The consolidation of concentrated suspensions’, Chem. Soc. Faraday Trans. I, Vol. 83, pp. 873-891. Bustos, M.C., Concha, F., Burger, R. and Tory, E.M. 1999, Sedimentation and Thickening, Kuwer Academic Publishers, London. Camp, T.R. and Stein, P.G. 1943, ‘Velocity gradients and internal work in fluid motion’, J. Boston Soc. Civ. Eng., Vol. 20, pp. 219 -237.

Page 189: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

188

Caron-Charles, M. and Gozlan, J.P. 1996, ‘Improvement of the floc resistance to a centrifugal shear field by polymer adjunction’, Chemical Engineering Science, Vol. 51, pp. 4649-4659. Casson, L.W. and Lawler, D.F. 1990, ‘Flocculation in turbulent flow: measurement and modelling of particle size distributions’, J. AWWA, Vol. 82, pp. 54-68. Chandler, S. and Hogg, R. 1987, ‘Sedimentaion and consolidation in destabalised suspensions’, Flocculation in Biotechnology and Separation Systems, Elsevier, Amsterdam, pp. 279-296. Chatzi, E.G. and Kiparissides, C. 1995, ‘Steady-state drop-size in high holdup fraction dispersion systems’, AIChE Journal, Vol. 41, pp. 1640-1652. Chellam, S. and Wiesner, M.R. 1993, ‘Fluid mechanics and fractal aggregates’, Wat. Res., Vol. 27, pp. 1493-1496. Chen, C-J. and Jaw, S-Y. 1998, Fundamentals of Turbulence Modeling, Taylor & Francis, USA. Chen, D. and Doi, M. 1989, ‘Simulation of aggregating colloids in shear flow II’, J. Chem. Phys., Vol. 94, No. 4, pp. 2656-2663. Chen, W., Fisher, R.R. and Berg, J.C. 1990, ‘Simulation of particle size distribution in an aggregation-breakup process’, Chemical Engineering Science, Vol. 45, pp. 3003-3006. Chen, W-J. 1998, ‘Effects of surface charge and shear during orthokinetic flocculation and on the adsorption and sedimentation of kaolin suspensions in polyelectrolyte solutions’, Separation Science and Technology, Vol. 33, No. 4, pp. 569-590. Cheng, Z. and Redner, S. 1988, ‘Scaling theory of Fragmentation’, Physical Review Letters, Vol. 60, No. 24, pp. 2450-4253. Chin, C.J., Yiacoumi, S. and Tsouris, C. 1998, ‘Shear-induced flocculation of colloidal particles in stirred tanks’, Journal of Colloid and Interface Science, Vol. 206, no. 2, pp. 532-545. Chiu, W-J. and Don, T-M. 1989, ‘A study on viscosity of suspensions’, Journal of Applied Polymer Science, Vol. 37, pp. 2973-2986. Chung, C.B., Park, S.H., Han, I.S., Seo, I.S. and Yang, B.T. 1998, ‘Modelling of ABS latex coagulation processes’, AIChE Journal, Vol. 44, No. 6, pp. 1256-1265. Clark, M.M. 1985, ‘Critique of Camp and Stein’s RMS velocity gradient’, J. Env. Eng., Vol. 111, pp. 741-754. Clark, N.N and Turton, R. 1988, ‘Chord length distributions related to bubble size distributions in multiphase flow’, Int. J. Multiphase Flow, Vol. 14, No. 4, pp. 413-424. Cleasby, J.L. 1984, ‘Is velocity gradient a valid turbulent flocculation parameter?’, J. Env. Eng., Vol. 110, pp. 875-987. Clift, R., Grace, J.R. and Weber, M.E. 1978, Bubbles, Drops and Particles, Academic Press, New York. Coe, H.S. and Clevenger, G.H. 1916, ‘Methods for determining the capacities of slime settling tanks’, Trans. Amer. Inst. Min. Engng., Vol. 60, pp. 356-384. Coghill, P.J., Millen, M.J., Rainey, S. and Sowerby, B.D. 1997, ‘On-line particle size analysis of fine mineral slurries’, Sixth Mill Operator’s Conference, AusIMM, Madang, 7-9 October. Cohen, R.D. and Weisner, M.R. 1990, ‘Equilibrium structure and the kinetics of fractal aggregation’, J. Surface Sci. Technol., Vol. 6, pp. 25-46.

Page 190: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

189

Concha, F. and Almendra, E.R. 1979, ‘Settling velocities of particlulate systems, 1. Settling velocities of individual spherical particles’, International Journal of Mineral Processing, Vol. 5, pp. 349-367. Concha, F., Bustos, M.C. and Barrientos, A. 1996, ‘Phenomenological theory of sedimentation’, Chapter 3 in: Sedimentation of Small Particles in a Viscous Fluid, Ed: Tory, E.M., Computational Mechanics Publications, Boston, USA. Coulaloglou, C.A. and Tavlarides, L.L. 1977, ‘Description of Interaction processes in agitated liquid-liquid dispersions’, Chem. Eng. Sci., Vol. 32, pp. 1289-1297. Cowie, B. and Schmoll, J. 1994, ‘Investigation of a novel facility design concept for heavy oil North Sea development’, May 5 1994, Society for Petroleum Engineering, “Off-shore Technical Conference”. Cresswell, P.J., Harig, F.E., Johnston, R.R.M., Leigh, G.M. and Thurlby, J.A. 1994, ‘Modelling alumina trihydrate precipitation - prediction of laboratory kinetic and size data’, 6th AusIMM Extractive Metallurgy Conference, Brisbane, pp. 325-332. Curtis, A.S.G. and Hocking, L.M. 1970, ‘Collision efficiency of equal spherical particles in a shear flow’, Trans. Faraday Soc., Vol. 66, pp. 1381-1390. Dahlstrom, D.A. and Fitch, E.B. 1985, ‘Thickening’, Chapter 9.2 in: SME Mineral Processing Handbook, Ed: Weiss, N.L. American Institute of Mining, Metallurgical, and Petroleum Engineers, Inc., Kingsport Press, USA. Dahlstrom, D.A. and Emmetr, J.R. 1985, ‘Solid-liquid separations’, Chapter 13-4 in: SME Mineral Processing Handbook, Ed: Weiss, N.L. American Institute of Mining, Metallurgical, and Petroleum Engineers, Inc., Kingsport Press, USA. Daniels, R. and Barta, A. 1992, ‘Preparation, characterization and stability assessment of oil-in-water emulsions with hydroxypropylmethyl cellulose as emulsifier’, October 20-21, Lasentec user meeting. Danielson, T.J., Muzzio, F.J. and Ottino, J.M. 1991, ‘Aggregation and structure formation in chaotic and regular flows’, Phys. Rev. Lett., Vol. 66, pp. 3128-3131. Darcy, H.P.G. 1856, Les Fontaines Publiques de la Ville de Dijon, Victor Dalmont, Paris. Daugherty, R.L., Franzini, J.B. and Finnemore, E.J. 1989, Fluid Mechanics with Engineering Applications, McGraw-Hill, Singapore. Davis, R.H. and Gecol, H. 1994, ‘Hindered settling function with no empirical parameters for polydisperse suspensions’, AIChE Journal, Vol. 40, No. 3, pp. 570-575. Davies, J.T. 1972, Turbulence Phenomena, Academic Press, London. Daymo, E.A., Hylton, T.D. and May, T.H. 1999, ‘Acceptance testing of the Lasentec Rocussed Beam Reflectance Measurment (FBRM) monitor for slurry transfer applications at Hanford and Oak Ridge’, Nuclear Waste Instrumentation Engineering, Proceedings SPIE, Vol. 3536, pp. 82-92. de Boer, G.B.J., Hoedemakers, G.F.M. and Thoenes, D. 1989a, ‘Coagulation in turbulent flow: part I’, Chem. Eng. Res. Des., Vol. 67, pp. 301-307. de Boer, G.B.J. Hoedemakers, G.F.M. and Thoenes, D. 1989b, ‘Coagulation in turbulent flow: part II’, Chem. Eng. Res. Des., Vol. 67, pp. 308-315. de Kretser, R.G., Usher, S.P. Lester, D. and Scales, P.J. 2000, ‘Compressional de-watering of suspensions – from fundamentals to application’, CHEMECA 2000, Perth Western Australia, July, p. 425.

Page 191: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

190

Delichatsios, M.A. 1980, ‘Particle coagulation in steady turbulent flows: applications to smoke aging’, Journal of Colloid Interface Science, Vol. 78, pp. 163-174. Delichatsios, M.A. and Probstein, R.F. 1975, ‘Coagulation in turbulent flow: theory and experiment’, Journal of Colloid Interface Science, Vol. 51, pp. 394-405. Delichatsios, M.A. and Probstein, R.F. 1976, ‘The effect of coalescence on the average drop size in liquid-liquid dispersions, Ind. Eng. Chem. Fund., Vol. 15, pp. 134-138. Deryagin, B.V. and Landau, L.D. 1941, ‘Theory of the stability of strongly charged lyophobic sols and the adhesion of strongly charged particles in solutions of electrolytes’, Acta Physicochimica URSS, Vol. 14, pp. 633-662. Di Felice, R. 1993, ‘Solid suspension in liquid fluidised beds’, in Shamlou, P.A. (Ed.): Processing of solid-liquid separations, Butterworth-Heinemann LTD, UK. Dickin, F.J., Hoyles, B.S., Hunt, A., Huang, S.M., Ilyas, O., Lenn, C., Waterfall, R.C., Williams, R.A., Xie, C.G. and Beck, M.S. 1992, ‘Tomographic imaging of industrial-process equipment – Techniques and applications’, IEE Proceedings-G Circuits Devices and Systems, Vol. 139, No. 1, pp. 72-82. Diehl, S. 1997, ‘Dynamic and steady-state behavior of continuous sedimentation’, SIAM J. Appl. Math., Vol 57, No. 4, pp. 991-1018. Dijkstra, J.J.F.M., de Nie, L.H., and Pot, A. 1996, ‘In-line measurement of crystal size. Campaign experience with the Lasentec M200 system’, Int. Sugar J., Vol. 98, Nol. 1174, pp. 521-523. Diplas, P. and Papanicolaou, A.N. 1997, ‘Batch analysis of slurries in zone settling regime’, Journal of Environmental Engineering, Vol. 123, No. 7, pp. 659-667. Dircan, C. 1981, ‘The structure and growth of aggregates in flocculation’, MS Thesis, The Pennsylvania State University. Ditter, W., Eisenlauer, J. and Horn. 1982, ‘Laser optical methods for dynamic flocculation testing’ in Tadros (Ed.): Effect of polymers on Dispersion Processes, Academic Press, London, 1982. Dobias, B. 1993, Coagulation and Flocculation: Theory and Applications, Marcel Dekker, USA. Dollimore, D. and Horridge, T.A. 1971, ‘The floc size of kaolinite flocculated by polyacrylamide’, Trans. J. Brit. Ceram. Soc., Vol. 70, No. 5, pp. 191-194. Dollimore, D. and McBride, G.B. 1968, ‘Alternative methods of calculating particle size from hindered settling measurements’, J. Appl. Chem., Vol. 18, pp. 136-140. Ducoste, J.J. and Clark, M.M. 1998, ‘The influence of tank size and impeller geometry on turbulent flocculation: II. Model’, Environmental Engineering Science, Vol. 15, No. 3, pp. 225-235. Ducoste, J.J. and Clark, M.M. 1999, ‘Turbulence in flocculators: comparison of measurements and CFD simulations’, AIChE Journal, Vol. 45, No. 2, pp. 432-436. Durand, R. 1953, ‘Basic relationships of the transportation of solids in pipes – Experimental research’, Proceedings – Minnesota International Hydraulics Convention, pp. 89-103. Durand, R. and Condolios, E. 1952, ‘The hydraulic transport of coal and solid materials in pipes’, Proceedings of a Colloquium on the Hydraulic Transport of Coal, pp. 39-52. Dyer, K.R. and Manning, A.J. 1999, ‘Observation of the size, settling velocity and effective density of flocs, and their fractal dimensions’, Journal of Sea Research, Vol. 41, pp. 87-95.

Page 192: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

191

Einstein, A 1908, ‘Eine neue bestimmung der molekül-dimensionen’, Annalew der Physik, Vol. 19, pp. 289-306. Eirich, F.R. 1960, Rheology Theory and Applications, Academic Press, London. Eisenlauer, J. and Horn, D. 1984, ‘Influence of shear and salt on flocculation in laminar tube flow’, in: Solid-Liquid Separation (Ed. J. Gregory), Ellis Horwood, Chichester, pp. 183-195. El-Hamouz, A.M. and Stewart, A.C. 1996, ‘On-line measurement of oil-water dispersion using a Par-Tec M300 laser backscatter instrument’, SPE Annual Technical Conference, Denver Colorado USA, pp. 785-796. Elimelech, M., Gregory, J., Jia, X. and Williams, R.A. 1995, Particle Deposition and Aggregation Measurement, Modelling and Simulation. Butterworth-Heinemann. Ellis, C.E. and Glasgow, L.A. 1999, ‘Deformability and breakage of flocs’, Advances in Environmental Research, Vol. 3, No. 1, pp. 15-27. Etchells, A.W. and Short, D.G.R. 1988, ‘Pipeline mixing – a user’s view part 1 – turbulent blending’, 6th European Conference on Mixing, Pavia, Italy, 24-26 May, pp. 539-544. Fair, G.M. and Gemmell, R.S. 1964, ‘A mathematical model of coagulation’, Journal of Colloid Interface Science, Vol. 19, pp. 360-372. Fakatselis, T.E. 1994, ‘Swenson continuous DTB crystallizers: real time control for improved product quality and process efficiency’, AIChE 1994 Annual Meeting. Famularo, J. and Happel, J. 1965, ‘Sedimentation of dilute suspensions in creeping flow’, AIChE Journal, Vol. 11, pp. 981-988. Farinato, R.S. and Dubin, P.L. (Eds.) 1999, Colloid-Polymer Interactions, From Fundamentals to Practice, John-Wiley & Sons, UK. Farrell, R.J. and Tsai, Y. 1995, ‘Nonlinear controller for batch crystallization: development and experimental demonstration’, AIChE Journal, Vol. 41, No. 10, pp. 2318-2321. Farrow, J.B., Fawell, P.D., Johnston, R.R.M., Nguyen, T.B., Rudman, M., Simic, K. and Swift, J.D. 2000, ‘Recent developments in technologies and methodologies for improving thickener performance’, Chemical Engineering Journal, Vol. 80, pp. 149-155. Farrow, J.B., Gong, W-Q., and Warren, L.J. 1985, ‘Floc size and density by the hindered settling method’, CHEMECA ’85, pp. 211-216. Farrow, J.B., Horsley, R.R. and Meagher, L. 1989, ‘The effect of alkali and alkaline earth cations on the rheology of concentrated quartz suspensions’, Journal of Rheology, Vol. 33, No. 8, pp. 1213-1230. Farrow, J.B. and Swift, J.D. 1996, ‘A new procedure for assessing the performance of flocculants’, International Journal of Mineral Processing, Vol. 46, pp. 263-275. Farrow, J.B. and Warren, L.J. 1989a, ‘A new technique for characterising flocculated suspensions’, Dewatering Technology and Practice, Brisbane Australia, pp. 61-64. Farrow, J.B. and Warren, L.J. 1989b, ‘The measurment of floc density—floc size distributions’, Flocculation and Dewatering, Engineering Foundation, New York, pp. 153-165. Farrow, J.B. and Warren, L.J. 1993, ‘Measurement of the size of aggregates in suspension’, Chapter 9 in: Coagulation and Flocculation: Theory and Applications, Ed: Dobias, B., Marcell Dekker, New York. pp. 391-426.

Page 193: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

192

Fawell, P., Richmond, W. Jones, L. and Collisson, M. 1997, ‘Focussed beam reflectance measurement in the study of mineral suspensions’, Chem. in Aust., 64, No. 2, pp. 4-6. Feke, D.L. and Showalter, W.R. 1983, ‘The effect of Brownian diffusion of shear-induced coagulation of colloidal dispersions’, Journal of Fluid Mechanics, Vol. 133, pp, 17-36. Fitch, B. 1975a, ‘Current theory and thickener design (1975) – Part 1, Introduction – zone settling theories’, Filtration and Separation, July 1975, pp. 355-359. Fitch, B. 1975b, ‘Current theory and thickener design (1975) – Part 2, Compression theory’, Filtration and Separation, September/October 1975, pp. 480-553. Fitch, B. 1975c, ‘Current theory and thickener design (1975) – Part 3, Design procedures’, Filtration and Separation, November/December 1975, pp. 636-638. Fitch, B. 1990, ‘A two-dimensional model for the free-settling regime in continuous thickening’, AIChE Journal, Vol. 36, No. 10, pp. 1545-1554. Fitch, E.B. 1987, ‘Gravity sedimentation operations’, Chapter 25 in: Encyclopedia of Chemical Processing and Design, Eds: McKetta, J.J. and Cunningham, W.A., Marcel Dekker. Fleer, G.J. and Scheutjens, J.M.H.M. 1993, ‘Modelling polymer adsorption, steric stabilization and flocculation’, in: Dobias, B. (Ed.), Coagulation and Flocculation: Theory and Applications, Marcel Dekker, USA, pp. 209-263. Flesch, J.C., Spicer, P.T. and Pratsinis, S.E. 1999, ‘Laminar and turbulent shear-induced flocculation of fractal aggregates’, AIChE Journal, Vol. 45, No. 5, pp. 1114-1124. Font, R. 1988, ‘Compression zone effect in batch sedimentation’, AIChE Journal, Vol. 34, No. 2, pp. 229-238. Font, R. 1990, ‘Calculation of the compression zone height in continuous thickeners’, AIChE Journal, Vol. 36, No. 1, pp. 3-12. Font, R. 1991, ‘Analysis of the batch sedimentation test’, Chemical Engineering Science, Vol. 46, No. 10, pp. 2473-2482. Friedlander, S.K., 1977, Smoke, Dust, and Haze, Wiley, New York. Galvin, K.P. and Waters, A.G. 1987, ‘The effect of sedimentation feed flux on the solid flux curve’, Powder Technology, Vol. 53, pp. 113-120. Gardner, K.H. and Theis, T.L. 1996, ‘A unified kinetic model for particle aggregation’, Journal of Colloid and Interface Science, Vol. 180, pp. 162-173. Geffroy, C., Persello, J., Foissy, A., Lixon, P., Tournilhac, F. and Cabane, B. 2000, ‘Molar mass selectivity in the adsorption of polyacrylates on calcite’, Colloids and Surfaces, A: Physicochemical and Engineering Aspects, Vol. 162, pp. 107-121. Gillespie, T. 1983, ‘The effect of aggregation and particle size distribution on the viscosity of Newtonian suspensions’, Journal of Colloid and Interface Science, Vol. 94, No. 1, pp. 166-173. Glasgow, L.A. and Kim, Y.H. 1986, ‘Characterization of agitation intensity in flocculation processes’, J. Environ. Eng., Vol. 112, pp. 1158-1163. Glasgow, L.A. and Liu, S.X. 1995, ‘Effects of dosing regimen and agitation profile upon floc characteristics’, Chem. Eng. Commun., Vol. 132, pp. 223-237.

Page 194: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

193

Glasgow, L.A. and Lueche, R.H. 1980, ‘Mechanisms of deaggregation for clay-polymer flocs in turbulent systems’, Ind. Eng. Chem. Fundam., Vol. 19, pp. 148-156. Godfrey, J.C. and Amirtharajah, A. 1991, ‘Mixing in liquids’, in: Amirtharajah, A. et al. (Ed.) Mixing in Coagulation and Flocculation, American Water Works Association Research Foundation, USA, pp. 35-79. Gonzalez, E.A. and Hill, P.S. 1998, ‘A method for estimating the flocculation time of monodispersed sediment suspensions’, Deep-Sea Research, Vol. 45, pp. 1931-1954. Gould, B.W. 1985, ‘Velocity Shear Gradients in Pipes and Jets’, Effluent and Water Treatment J., Vol. 25, No. 4, pp. 127-128. Govier, G.W. and Aziz, K. 1972, The Flow of Complex Mixtures in Pipes, Van Nostrand Reinhold Company. Melbourne. Green, M.D., Eberl, M. and Landman, K.A. 1996, ‘Compressive yield stress of flocculated suspensions: deturmination via experiment’, AIChE Journal, Vol. 42, No. 8, pp. 2308-2318. Greenberg, S.A., Jarnutowski, R. and Chang, T.N. 1965, ‘The behaviour of polysilicic acid, II. The rheology of silica suspensions’, Journal of Colloid and Interface Science, Vol. 20, pp. 20-43. Gregory, J. 1973, ‘Rates of flocculation of latex particles by cationic polymers’, Journal of Colloid and Interface Science, Vol. 42, pp. 448-456. Gregory, J. 1981, ‘Flocculation in laminar flow tube’, Chem. Eng. Sci., Vol. 36, pp. 1789-1794. Gregory, J. 1987a, ‘Flocculation by polymers and polyelectrolytes’, Solid/Liquid Dispersions, Academic press, pp. 163-181. Gregory, J. 1987b, ‘Kinetic aspects of polymer adsorption and flocculation’, Flocculation in Biotechnology and Separation Systems, Y.A. Attia (Ed.), Elsevier, Amsterdam, pp. 31-44. Gregory, J. 1989, ‘Fundamentals of flocculation’, CRC Critical Reviews in Environmental Control, Vol. 19, pp. 185-230. Gregory, J. 1993, ‘Stability and flocculation of suspensions’, in Shamlou, P.A. (Ed.), Processing of Solid-Liquid Separations, Butterworth-Heinemann LTD, UK, pp. 59-92. Gregory, J. 1997a, ‘The role of floc density in solid-liquid separation’, Filtration and Separation, Vol. 35, No. 4, pp. 367-371. Gregory, J. 1997b, ‘The density of particle aggregates’, Wat. Sci. & Tech, Vol. 36, No. 4, pp. 1-13. Gregory, J. and Gubai, L. 1991, ‘The effects of dosing and mixing conditions on polymer flocculation of concentrated suspensions’, Chem. Eng. Comm, Vol. 108, pp. 3-21. Gruy, F. and Saint-Raymond, H. 1997, ‘Turbulent coagulation efficiency’, Journal of Colloid Interface Science, Vol. 185, pp. 281-284. Guibai, L. and Gregory, J. 1991, ‘Flocculation and sedimentation of high-turgidity waters’, Water Research, Vol. 25, pp. 1137-1143. Ham, R.K. and Christman, R.F. 1969, ‘Agglomerate size changes in coagulation’, J. Sanit. Eng. Div. ASCE, Vol. 95, pp. 481-502. Han, M.Y. and Lawler, D.F. 1992, ‘The (relative) insignificance of G in flocculation’, Journal of the American Water Works Association, Vol. 84, No. 10, pp. 79-91.

Page 195: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

194

Hanseler, J. and McKean, W. 1988, ‘Laser technology offers new way to measure furnish components’, Pulp and Paper Canada, Vol. 89, No. 9, pp. 25-32. Hansen, S. and Ottino, J.M. 1996, ‘Aggregation and cluster size evolution in non-homogeneous flows’, Journal of Colloid Interface Science, Vol. 179, pp. 89-103. Happel, J and Brenner, H. 1973, ‘The motion of a rigid particle of arbitrary shape in unbounded fluid’, Chapter 5 in: Low Reynolds Number Hydrodynamics, Noordhoff International Publishing, Leyden. Haw, M.D., Poon, W.C. and Pusey, P.N. 1997, ‘Structure and arrangement of clusters in cluster aggregation’, Physical Review, Vol. 56, No. 2, pp. 1981-1933. Healy, T.W. 1961, ‘Flocculation-dispersion behavior of quartz in the presence of a polyacrylamide flocculant’, Journal of Colloid Science, Vol. 16, pp. 609-617. Healy, T.W. and La Mer, V.K. 1964, ‘The energetics of flocculation and redespersion by polymers’, Journal of Colloid Science, Vol. 19, pp. 323-331. Healy, T.W., Boger, E.V., White, L.R., Leong, Y.K. and Scales, P.J. 1994, ‘Thickening and clarification; how much do we really know about dewatering?’, AusIMM Extractive Metallurgy Conference, Brisbane 3-6 July, pp. 105-109. Heath, A.R., Fawell, P.D., Bahri, P.A. and Swift, J.D., 2002, ‘Estimating average particle size by focussed beam reflectance measurement (FBRM)’, Particle and Particle Systems Characterisation., Vol. 19, pp. 84-95. Heath, A.R., Fawell, P.D., Bahri, P.A. Swift, J.D. 1999, ‘Population balance modelling of flocculation processes’, CHEMECA, Perth, September 1999. Hecker, R., Kirwan, L., Fawell, P.D., Farrow, J.B., Swift, J.D. and Jefferson, A. 1999, ‘Focused beam reflectance measurement for the continuous assessment of flocculant performance’, in: Polymers in Mineral Processing, J.S. Laskowksi (Ed.), Canadian Instiute of Mining, Metallurg, Petroleum, Montreal, Quebec, pp. 91-105. Heffels, C., Polke, R., Radle, M. Sachweh, B., Schafer, M. and Scholz, N. 1998, ‘Control of particulate processes by optical measurement techniques’, Part. Part. Syst. Charact., Vol. 15, pp. 211-218. Heiss, J.F. and Coull, J. 1952, ‘The effect of orientation and shape on the settling velocity of non-isometric particles in a viscous medium’, Chemical Engineering Process, Vol. 48, No. 3, pp. 133-140. Herbst, J.A. and Sepulveda, J.L. 1985, ‘Characterisation of test samples’, in: SME Mineral Processing Handbook, Ed: Weiss, N.L. Society of Mining Engineers, New York, USA, section 30. Hiemenz, P.C. 1986, Principles of Colloid and Surface Chemistry, Marcel Dekker, USA. Higashitani, K., Yamauchi, K, Matsuno, Y. and Hosokawa, G. 1983, ‘Turbulent coagulation of particles in a viscous fluid’, J. Chem. Eng. Japan, Vol. 16, pp. 299-304. Hildred, K.L., Townson, P.S., Hutson, G.V. and Williams, R.A. 2000, ‘Characterisation of particulates in the BNFL enhanced actinide removal plant’, Powder Technology, Vol. 108, pp. 164-172. Hill, P.S. 1996, ‘Sectional and discrete representations of floc breakage in agitated suspensions’, Deep-Sea Research Part 1 - Oceanographic Research Papers, Vol. 43, No. 5, pp. 679-702. Hobbel, F., Davies, R., Rennie, F.W., Allen, T., Butler, L.E., Waters, E.R., Smith, J.T. and Sylvester, R.W. 1991, ‘Modern methods of on-line size analysis for particulate process streams’, Part. Part. Syst. Charact., Vol. 8, pp. 29-34.

Page 196: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

195

Hogg, R. 1984, ‘Collision efficiency factors for polymer flocculation’, Journal of Colloid Interface Science, Vol. 102, pp. 232-236. Hogg, R. 1999a, ‘Polymer adsorption and flocculation’, in Laskowski, J.S. (Ed.): Polymers in Mineral Processing, Symposium organised by: The Canadian Institute of Mining, Metallurgy and Petroleum. Quebec, Canada August 1999, Met. Soc., pp. 3-17. Hogg, R. 1999b, ‘The role of polymer adsorption kinetics on flocculation’, Colloids and Surfaces, Vol. 146, pp. 253-263. Hogg, R. 2000, ‘Flocculation and dewatering’, Int. J. Miner. Process., Vol. 58, pp. 223-236. Hogg, R. and Bunnaul, P. 1992, ‘Sediment compressibility in thickening of flocculated suspensions’, Minerals and Metallurgical Processing, Vol. 9, pp. 184-188. Hogg, R., Klimpel, R.C. and Ray, D.T. 1987, ‘Agglomerate structure in flocculated suspensions and its effect on sedimentaion and dewartering’, Minerals and Metallurgical Processing, Vol. 4, pp. 108-114. Hokanson, J. and Preikschat, E. 1992, ‘In-line particle size measurement for improved process control in the mineral processing industry’, Light Metals, 1991, pp. 39-42. Honaker, R.Q., Luttrell, G.H. and Yoon, R.H. 1994, ‘In-situ measurement of coagula size using an on-line particle size analyzer’, SME Annual Meeting, New Mexico. Horne, D.S. 1987, ‘Determination of the fractal dimension using turbidimetric techniques’, Faraday Discuss. Chem. Soc., Vol. 83, pp. 259-270. Horsley, R.R., Jones, R.L. and Snow, R.J. 1984, ‘The influence of inorganic impurities on the rheological behaviour of silica based slurries’, IX International Congress on Rheology, Mexico, pp. 617-622. Horwatt, S.W., Feke, D.L. and Manas-Zloczower, I. 1992, ‘The influence of structural heterogeneities on the cohesivity and breakup of agglomerates in simple shear flow’, Powder Technology, Vol. 72, pp. 113-119. Hounslow, M.J., Ryall, R.L. and Marshall, V.R. 1988, ‘A discretized population balance for nucleation, growth and aggregation’, AIChE Journal, Vol. 34, No. 11, pp. 1821-1832. Hsu, J.-P., Lin, D.-P. and Tseng, S. 1995, ‘The sticking probability of colloidal particles in polymer-induced flocculation’, Colloid Polym. Sci., Vol. 273, pp. 271-278. Huang, H. 1994, ‘Fractal properties of flocs formed by fluid shear and differential settling’, Phys. Fluids, Vol. 6, No. 10, pp. 3229-3234. Huang, P.Y. and Helums, J.D. 1993, ‘Aggregation and disaggregation kinetics of human blood-platelets: part III. The disaggregation under shear-stress of platelet aggregates’, Biophysical Journal, Vol. 65, No. 1, pp. 354-361. Hughes, M.A. 2000, ‘Coagulation and Flocculation’, in Svarovsky, L. (Ed.), Solid-Liquid Separation, 4th edition, Butterworth-Heinmann, UK. Ilievski, D. 1996, ‘Can gibbsite precipitation be modelled from first principles?’, Fourth Int. Alumina Quality Workshop (Darwin, NT) Proc. pp. 263-271. Ilievski, D. and Hounslow, M.J. 1995, ‘Agglomeration during precipitation. 2. Mechanism deduction from tracer data’, AIChE Journal, Vol. 41, pp. 525-535. Ilievski, D. and White, E.T. 1994, ‘Agglomeration during pPrecipitation: Agglomeration mechanism identification for Al(OH)3 crystals in stirred caustic aluminate solutions’, Chem. Eng. Sci., Vol. 49, pp. 3227-3239.

Page 197: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

196

Ilievski, D. and White, E.T. 1995, ‘Agglomeration during precipitation. 1. Tracer crystals for Al(OH)3 Precipitation’, AIChE Journal, Vol. 41, pp. 518-524. Ilievski, D. and White, E.T. 1995, ‘Modelling Bayer precipitation with agglomeration’, Light Metals, pp. 55-61. Ilievski, D., White, E.T. and Hounslow, M.J. 1993, ‘Agglomeration Mechanism Identification Case Study: Al(OH)3 Agglomeration during precipitation from seeded supersaturated caustic aluminate solutions’, Sixth Int. Symp. on Agglomeration (Nagoya, Japan) Proc., pp. 93-98. Ives, K.J. 2000, ‘Orthokinetic flocculation’, in Svarovsky, L. (Ed.), Solid-Liquid Separation, 4th edition, Butterworth-Heinmann, UK. Ives, K.J. and Bhole, A.G. 1973, ‘Theory of flocculation for continuous flow system’, J. Environ. Eng. Div. ASCE, Vol. 99, pp. 17-34. Ives, K.J. and Bhole, A.G. 1975, ‘Study of flowthrough Couette flocculators – I. Design for uniform and tapered flocculation’, Water Research, Vol. 9, pp. 1085-1092. Ives, K.J. and Bhole, A.G. 1977, ‘Study of flowthrough Couette flocculators – II. Laboratory studies of flocculation kinetics’, Water Research, Vol. 11, pp. 209-215. Jakeman, A.J. and Anderssen, R.S. 1975, ‘Abel type integral equations in stereology, I General discussion’, Journal of Microscopy, Vol. 105, pt. 2, pp. 121-133. Jiang, Q. and Logan, B.E. 1991, ‘Fractal dimensions of aggregates determined from steady-state size distributions’, Environ. Sci. Technol., Vol. 25, No. 12, pp. 2031-2038. Johnson, B.K., Szeto, C., Davidson, O. and Andrews, A. 1997, ‘Optimization of pharmaceutical batch crystallization for filtration and scale-up’, AIChE Annual Meeting, Los Angeles, paper 16a. Johnson, C.P., Li, X. and Logan, B.E. 1996, ‘Settling velocities of fractal aggregates’, Env. Sci. Tech., Vol. 30, pp. 1911-1918. Johnson, S.B., Franks, G.V., Scales, P.J., Boger, D.V. and Healy, T.W. 2000, ‘Surface chemistry-rheology relationships in concentrated mineral suspensions’, Int. J. Miner. Process., Vol. 58, pp. 267-304. Johnston, R.R.M. and Cresswell, P.J. 1996, ‘Modelling alumina precipitation: dynamic solution of the population balance equation’, Fourth International Alumina Quality Workshop, pp. 281-290. Jullien, R. and Meakin, P. 1989, ‘Simple models for the restructuring of three dimensional ballistic aggregates’, Journal of Colloid Interface Science, Vol. 127, pp. 265-272. Kanungo, S.B. and De, P.K. 1970, ‘Physico-chemical studies of precipitated magnesium hydroxide: part III – estimation of floc characteristics from hindered settling rates of dilute suspensions’, Indian Journal of Technology, Vol. 8, pp. 180-184. Kao, S.V., Nielsen, L.E. and Hill, C.T. 1975, ‘Rheology of concentrated suspensions of spheres, 1. Effect of the liquid-solid interface’, Journal of Colloid and Interface Science, Vol. 53, No. 3, pp. 358-366. Kao, T.Y. and Wood, P.J. 1974, Trans. Soc. Min. Eng. of AIME, Vol. 255, pp 39. Kasper, D.R. 1971, Theoretical and experimental investigations of the flocculation of charged particles in aqueous solutions by polyelectrolytes of opposite charge, PhD thesis, California Institute of Technology.

Page 198: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

197

Keys, R.O. and Hogg, R. 1978, ‘Mixing problems in polymer flocculation’, Water 1978, AIChE Symposium Series, pp. 63-72. Kim, Y.H. and Glasgow, L.A. 1987, ‘Simulation of aggregate growth and breakage in stirred tanks’, Industrial & Engineering Chemistry Research, Vol. 26, pp. 1604-1609. Klimpel, R.C., Dirican, C. and Hogg, R. 1986, ‘Measurment of agglomerate density in flocculated fine particle suspensions’, Powder Technology, Vol. 4, pp. 45-49. Klute, R. and Amirtharajah, A. 1991, ‘Particle destabilization and flocculation reactions in turbulent pipe flow’, in Amirtharajah, A. et al. (Eds.), Mixing in Coagulation and Flocculation, American Water Works Association Research Foundation, USA, pp. 217-255. Kobayashi, M., Adachi, Y. and Ooi, S. 1999, ‘Breakup of fractal flocs in a turbulent flow’, Langmuir, Vol. 15, pp. 4351-4356. Koh, P.T.L. 1984, ‘Compartmental modelling of a stirred tank for flocculation requiring a minimum critical shear rate’, Chemical Engineering Science, Vol. 39, No. 12, pp. 1759-1764. Koh, P.T.L., Andrews, J.R.G. and Uhlherr, P.T.H. 1984, ‘Flocculation in stirred tanks’, Chemical Engineering Science, Vol. 39, No. 6, pp. 975-985. Koh, P.T.L., Andrews, J.R.G. and Uhlherr, P.T.H. 1987, ‘Modelling shear-flocculation by population balances’, Chemical Engineering Science, Vol. 42, No. 2, pp. 353-362. Kohler, H.H. 1993, ‘Thermodynamics of Adsorption from solution’, in Dobias, B. (Ed.), Coagulation and Flocculation: Theory and Applications, Marcel Dekker, USA. Kolb, M., Botet, R. and Jullien, R. 1983, ‘Scaling of Kinetically Growing Clusters’, Physical Review Letters, Vol. 51, No. 13, pp. 1123-1122. Kolmogoroff, A.N. 1941, ‘Dissipation of energy in locally isotropic turbulence’, Doklady Acadamy of Science, USSR, Vol. 32, pp. 16-18. Kos, P. 1980, ‘Review of sedimentation and thickening’, Chaper 80 in: Fine Particle Processing, Ed: Somasundaran, P., American Institute of Mining, Metallurgical, and Petroleum Engineers, Inc., New York, USA. Kostoglou, M. and Karabelas, A.J. 1994, ‘Evaluation of zero order methods for simulating particle coagulation’, Journal of Colloid and Interface Science, Vol. 163, pp. 420-431. Kousaka, Y, Okuyama, K. and Payatakes, A.C. 1981, ‘Physical meaning and evaluation of dynamic shape factor of aggregate particles’, Journal of Colloid and Interface Science, Vol. 84, No. 1, pp. 91-99. Kramer, T.A. and Clark, M.M. 1999, ‘Incorporation of aggregate breakup in the simulation of orthokinetic coagulation’, Journal of Colloid and Interface Science, Vol. 216, pp. 116-126. Kranenburg, C. 1999, ‘Effects of floc strenght on viscosity and deposition of cohesive sediment suspensions’, Continental Shelf Research, Vol. 19, pp. 1665-1680. Kruis, F.E. and Kusters, K.A. 1997, ‘The collision rate of particles in turbulent flow’, Chem. Eng. Comm., Vol. 158, pp. 201-230. Krutzer, L.L.M., Van Diemen, A.J.G. and Stein, H.N. 1995, ‘The influence of the type of flow on the orthokinetic coagulation rate’, Journal of Colloid Interface Science, Vol. 171, pp. 429-438. Kruyt, H.R. (Ed) 1952, Colloid Science, Vol. 1, Elsevier Publishing Company, London.

Page 199: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

198

Kuboi, R. Komasawa, I. and Otake, T. 1972, ‘Collision and coalescence of dispersed drops in turbulent liquid flow’, Journal of Chemical Engineering of Japan, Vol. 5, No. 4, pp. 97-98. Kumar, S. and Ramkrishna, D. 1996, ‘On the solution of population balance equations by discretisation - I - A fixed pivot technique, Chemical Engineering Science, Vol. 51, No. 8, pp. 1311-1132. Kumar, S. and Ramkrishna, D. 1997, ‘On the solution of population balance equations by discretisation - III - Nucleation, growth and aggregation of particles’, Chemical Engineering Science, Vol. 52, No. 24, pp. 4659-4679. Kusters, K.A. 1991, ‘The influence of turbulence on aggregation of small particles in agitated vessels’, Ph.D. Thesis, Eindhoven University of Technology. The Netherlands. Kusters, K.A., Wijers, J.G. and Thoenes, D. 1997, ‘Aggregation kinetics of small particles in agitated vessels’, Chem. Eng. Sci., Vol. 52, pp. 107-121. Kusuda, T., Koga, K., Yorozu, H. and Awaya, Y. 1981, ‘Density and settling velocity of flocs’, Memoirs of the Faculty of Engineering, Kyushu University, Vol. 41, No. 3, pp. 269-280. Kynch, G.J. 1952, ‘A theory of sedimentation’, Trans. Faraday Soc., Vol. 48, pp. 166-176. La Mer, V.K. and Healy, T.W. 1963, ‘Adsorption-flocculation reactions of macromolecules at the solid-liquid interface’, Rev. Pure Appl. Chem., Vol. 13, pp. 112-133. Lainé, S., Phan, L., Pellarin, P. and Robert, P. 1999, ‘Operating diagnostics on a flocculator-settling tank using fluent CFD software’, Wat. Sci. Tech., Vol. 39, No. 4, pp. 155-162. Landman, K.A., White, L.R. and Eberl, M. 1995, ‘Pressure filtration of flocculated suspensions’, AIChE Journal, Vol. 41, pp. 1867-1700. Law, D.J., Bale, A.J. and Jones, S.E. 1997, ‘Adaptation of focused beam reflectance measurement to in-situ particle sizing in estuaries and coastal waters’, Marine Geology, Vol. 140, pp. 47-59. Lawler, D.F. 1993, ‘Physical aspects of flocculation: from microscale to macroscale’, Wat. Sci. Tech., Vol. 27, pp. 165-180. Lawler, D.F. 1997, ‘Particle size distribution in treatment processes: theory and practice’, Wat. Sci. Tech. Vol. 36. No. 4, pp. 15-23. Lee, C.W. and Brodkey, R.S. 1987, ‘A visual study of pulp floc dispersion mechanisms’, AIChE Journal, Vol. 33, No. 2, pp. 297-302. Leu, R. and Ghosh, M.M. 1988, ‘Polyelectrolyte characteristics and flocculation’, J. AWWA, Vol. 80, pp. 159-167. Levich, V.G. 1962, Physicochemical Hydrodynamics. Prentice Hall, pp. 213-219. Lick, W. and Lick, J. 1988, ‘Aggregation and disaggregation of fine-grained lake sediments, J. Great Lakes Res., Vol. 14, No. 4, pp. 514-523. Lin, M.Y., Lindsay, H.M., Weitz, D.A., Ball, R.C., Klein, R., and Meakin, P. 1989, ‘Universality of fractal aggregates as probed by light scattering’, Proc. R. Soc. London, Vol. 423, pp. 71-87. Lin, M.Y., Klein, R., Lindsay, H.M., Weitz, D.A., Ball, R. C. and Meakin, P. 1990, ‘The structure of fractal colloidal aggregates of finite extent’, Journal of Colloid and Interface Science, Vol. 137, No. 1, pp. 263-280.

Page 200: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

199

Lindquest, G.M. and Stratton, R.A. 1976 ‘The role of polyelectrolyte charge density and molecular weight on the adsorption and flocculation of colloidal silica with polyethylenimine’, Journal of Colloid and Interface Science, Vol. 55, pp. 45-59. Litster, D.J., Smit, D.J. and Hounslow, M.J. 1995, Adjustable discretized population balance for growth and aggregation’, AIChE Journal, Vol. 41, No. 3, pp. 591-603. Liu, S. and Masliyah, J.H. 1996, ‘Flow of suspensions in pipelines’, in: Suspensions, Fundamentals and Applications to the Petroleum Industry, in Schramm, L.L. (Ed.): Library of Congress Cataloguing-in-publication Data, USA. Liu, W. and Clark, N.N. 1995, ‘Relationships between distributions of chord lengths and distributions of bubble sizes including their statistical parameters’, Int. J. Multiphase Flow, Vol. 21, pp. 1073-89. Liu, W., Clark, N.N. and Karamavruc, A.I. 1998, ‘Relationship between bubble size distributions and chord-length distribution in heterogeneously bubbling systems’, Chemical Engineering Science, Vol. 52, No. 6, pp. 1267-1276. Lu, C.F. and Spielman, L.A. 1985, ‘Kinetics of floc breakage and aggregation in agitated liquid suspensions’, Journal of Colloid and Interface Science, Vol. 103, No. 1, pp. 95-105. Lu, S., Ding, Y. and Guo, J. 1998, ‘Kinetics of fine particle aggregation in turbulence’, Advances in Colloid and Interface Science, Vol. 78, pp. 197-235. Mabire, F., Audebert, R. and Quivoron, C. 1984, ‘Flocculation properties of some water-soluble cationc copolymers toward silica suspensions: A semiquantitative interpretation of the role of molecular weight and cationicity through a patchwork model’, Journal of Colloid Interface Science, Vol. 97, No. 1, pp. 120-136. Manning, A.J. and Dyer, K.R. 1999, ‘A laboratory examination of the floc characteristics with regard to turbulent shearing’, Marine Geology, Vol. 160, pp. 147-170. Marchal, P., David, R., Klein, J.P. and Villermaux, J. 1988, ‘Crystallization and precipitation engineering - I. An efficient method for solving population balance in crystallization with agglomeration’, Chemical Engineering Science, Vol 43, No. 1, pp. 59-67. Matsumo, K. and Mori, Y. 1975, ‘Settling velocity of floc – new measurment method of floc density’, Journal of Chemical Engineering of Japan, Vol. 8, No. 2, pp. 143-147. Matsuo, T. and Unno, H. 1981, ‘Forces acting on floc and strength of floc’, J. Envir. Eng. Div., ASCE, Vol. 107, pp. 527-545. Maude, A.D. and Whitmore, R.L. 1958, ‘A generalized theory of sedimentation’, Br. J. Appl. Phys., Vol. 9, pp. 477-482. McNown, J.S. and Lin, P-N. 1952, ‘Sediment concentration and fall velocity’, Midwestern Conference on Fluid Dynamics, Ohio State University, pp. 401-411. Meakin, P. 1983, ‘Formation of fractal clusters and networks by irreversible diffusion-limited aggregation’, Physical Review Letters, Vol. 51, No. 13, pp. 1119-1122. Meakin, P. 1985, ‘The effects of random bond breaking on diffusion limited cluster-cluster aggregation’, J. Chem. Phys., Vol. 83, No. 3, pp. 3645-3649. Meakin, P. 1988, ‘Fractal aggregates’, Adv. Colloid Interface Sci., Vol. 28, pp. 249-331. Meakin, P. and Jullien, R. 1988, ‘The effects of restructuring on the geometry of clusters formed by diffusion-limited, ballistic, and reaction-limited cluster-cluster aggregation’, J. Chem. Phys., Vol. 89, No. 1, pp. 246-250.

Page 201: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

200

Michaels, A.S. and Bolger, J.C. 1962, ‘Settling rates and sediment volumes of flocculated kaolin suspensions’, I&EC Fundamentals, Vol. 1, No. 1, pp. 24-33. Mills, P.D.A., Goodwin, J.W. and Grover, B.W. 1991, ‘Shear field modification of strongly flocculated suspensions - aggregate morphology’, Colloid Polymer Sci., Vol. 269, pp. 949-963. Mishler, R.T. 1912, ‘Settling slimes at the Tigre Mill’, Engng. Min. J., Vol. 94, pp. 643-646. Mishra, P.N., Severson, D.E. and Owens, T.C. 1970, ‘Rheological study of concentrated silica suspensions’, Chemical Engineering Science, Vol. 25, pp. 653-663. Monnier, O., Fevotte, G., Hoff, C. and Klein, J.P. 1997, ‘Model identification of batch cooling crystallizations through calorimetry and image analysis’, Chemical Engineering Science, Vol. 52, No. 7, pp. 1125-1139. Monnier, O., Klein, J. P., Hoff, C. and Ratsimba, B. 1996, ‘Particle size determination by laser reflection: methodology and problems’, Part. Part. Syst. Charact., Vol. 13, pp. 10-17. Moudgil, B.M. and Vasudevan, T.V. 1989, ‘Evaluation of floc properties for dewatering fine particle suspensions’, Minerals and Metallurgical Processing, Vol. 6, No. 3, pp. 142-145. Moussa, T. and Tiu, C. 1994, ‘Factors affecting polymer degradation in turbulent pipe flow’, Chem. Eng. Sci., Vol. 49, pp. 1681-1692. Mühle, K. 1993, ‘Floc stability in laminar and turbulent flow’, in Dobias, B. (Ed.), Coagulation and Flocculation: Theory and Applications, Marcel Dekker, USA, pp. 355-282. Muralidar, R. and Ramkrishna, D. 1985, ‘An inverse problem in agglomeration kinetics’, Journal of Colloid and Interface Science, Vol. 112, No. 2, pp. 348-361. Murphy, D.D., Bakker, M.G. and Spears, D.R. 1994, ‘New Insights into the Mechanism of Shear-Flocculation of Silica with Fatty Acids’, Minerals and Metallurgical Processing, Vol. 11, pp. 26-30. Murphy, D.D., Spears, D.R. and Bakker, M.G. 1995, ‘The role of micellar structures in shear flocculation. Part 1: Flocculation of quartz with alkylamine surfactants’, Colloids Surf. A: Physicochem. Eng. Aspects, Vol. 96, pp. 143-154. Nagata, S. 1975, Mixing Principles and Applications, John Wiley & Sons. Nasr-El-Din, H.A. 1996, ‘Flow of suspensions in pipelines’, in Schramm, L.L. (Ed.): Suspensions, Fundamentals and Applications in the Petroleum Industry, Library of Congress Cataloging-in-Publication Data, USA, pp. 177-226. Nauman, E.B. and Clark, M.M. 1991, ‘Residence time distribution’, in Amirtharajah, A. et al. (Eds.), Mixing in Coagulation and Flocculation, American Water Works Association Research Foundation, USA, pp. 127-169. Neimark, A.V., Koylu, U.O. and Rosner, D.E., 1996, ‘Extended characterisation of combustion-generated aggregates: Self-affinity and Lacunarities’, Journal of Colloid and Interface Science, Vol 180, pp. 590-597. Newitt, D.M., Richardson, J.F., Abbott, M. and Turtle, R.B. 1955, ‘Hydraulic conveying of solids in horizontal pipes’, Trans. Instn. Chem. Engrs., Vol. 33, pp. 93-113. Newton, I. 1687, Principa Mathematica, Book II, Proposition XXXVIII.

Page 202: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

201

Nguyen-Van, A., Schulze, H.J. and Kmet, S. 1994, ‘A simple algorithm for the calculation of the terminal velocity of a single solid sphere in water’, International Journal of Mineral Processing, Vol. 41, pp. 305-310. Nicmanis, M. and Hounslow, M.J. 1998, ‘Finite-element methods for steady-state population balance equations’, AIChE Journal, Vol. 44, No. 10, pp. 2258-2272. ∅degaard, H. 1979, ‘Orthkinetic flocculation of phosphate precipitates in a multicompartment reactor with non-ideal flow’, Prog. Wat. Tech., Suppl. 1, pp. 61-68. Oles, V. 1992, ‘Shear-induced aggregation and breakup of polystyrene latex particles’, Journal of Colloid Interface Science, Vol. 154, pp. 351-358. Orokasar, A.R. and Turian, R.M. 1980, ‘The critical velocity in pipeline flow of slurries’, AIChE Journal, Vol. 26, No. 4, pp. 550-558. Owen, A.T., Fawell, P.D., Swift, J.D. and Farrow, J.B. 2002, ‘The impact of polyacrylamide solution age on flocculation performance’, submitted to: International Journal of Mineral Processing. Vol. 67, pp. 123-144. Pandya, J.D. and Spielman, L.A. 1982, ‘Floc breakage in agitated suspensions: theory and data processing strategy’, Journal of Colloid Interface Science, Vol. 90, pp. 517-531. Parker, D.S., Kaufman, W.J. and Jenkins, D. 1972, ‘Floc breakup in turbulent flocculation processes’, Sanitary Engineering Division, Vol. 8702, pp. 79-99. Paterson, R.W. and Abernathy, F.H. 1970, ‘Turbulent flow drag reduction and degradation with dilute polymer solutions’, Journal of Fluid Mechanics, Vol. 43, pp. 689-710. Patil, D.P. and Andrews, J.R.G. 1998, ‘An analytical solution to continuous population balance model describing floc coalescence and breakage - a special case’, Chemical Engineering Science, Vol. 53, No. 3, pp. 599-601. Patil, D.P., Andrews, J.R.G. and Uhlherr, P.T.H. 1997, ‘A lumped discrete population balance model for shear flocculation - model development’, Int. J. Miner. Process., Vol. 50, pp. 289-306. Pearse, M.J. 1977, Gravity Thickening Theories: A Review, Warren Spring Laboratory, Dept. of Industry, UK. Pearson, H.J., Valioulis, I.A., and List, E.J. 1984, ‘Monte Carlo simulation of coagulation in discrete particle-size distributions, Part 1. Brownian motion and fluid shearing’, Journal of Fluid Mechanics, Vol. 143, pp. 367-385. Pelton, R.H. 1981, ‘A model for flocculation in turbulent flow’, Colloids Surf. Vol. 2, pp. 277-291. Peng, S.J. and Williams, R.A. 1994, ‘Direct measurement of floc breakage in flowing suspensions’, Journal of Colloid Interface Science, Vol. 166, pp. 321-332. Peng, S.J. and Williams, R.A. 1993, ‘Control and optimisation of mineral flocculation and transport using on-line particle size analysis’, Minerals Engineering, Vol. 6, pp. 133-153. Perez, M., Font, R. and Pastor, C. 1998, ‘A mathematical model to simulate batch sedimentation with compression behavior’, Comptuers Chem. Engng., Vol. 22, No. 11, pp. 1531-1541. Perry, H.R. and Green, D.W. 1997, Perry’s Chemical Engineers Handbook, Seventh Ed., McGraw-Hill. Pettyjohn, E.S. and Christiansen, E.B. 1948, ‘Effect of particle shape on free-settling rates of isometric particles’, Chemical Engineering Progress, Vol. 44, No. 2, pp. 157-172.

Page 203: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

202

Phillips, J.M. and Walling, D.E. 1995A, ‘Measurement in situ of the effective particle-size characteristics of fluvial suspended sediment by means of a field-portable laser backscatter probe: some preliminary results’, Mar. Freshwater Res., Vol. 46, pp. 349-357. Phillips, J.M. and Walling, D.E. 1998, ‘Calibration of a Par-Tec 200 laser back-scatter probe for in situ sizing of fluvial suspended sediment’, Hydrological Processes., Vol. 12, pp. 221-231. Phillips, J.M. and Walling, D.E. 1999, ‘The particle size characteristics of fine-grained channel deposits in the river Exe basin, Devon, UK’, Hydrological Processes, Vol. 13, pp. 1-9. Phillips, J.M. 1996, ‘The effective particle size characteristics of fluvial suspended sediment’, Ph.D. Thesis, Exeter University, UK. Phillips, J.M. and Walling, D.E. 1995B, ‘An assessment of the effects of sample collection, storage and resuspension on the representativeness of measurements of the effective particle size distribution of fluvial suspended sediment’, Wat. Res., Vol. 29, pp. 2498-2508. Poilov, V.Z., Kosvintsev, K., Gramlich, J, Kodura, J. and Karvot, R. 1997, ‘Formation of Magnesium chloride particles in bulk polythermal crystallization’, Russian Journal of Applied Chemistry, Vol. 70, No. 7, pp. 1036-1040. Potanin, A.A. 1991, ‘On the mechanism of aggregation in the shear flow of suspensions’, Journal of Colloid Interface Science, Vol. 145, pp. 140-157. Potanin, A.A. 1993, ‘On the computer simulation of the deformation and breakup of colloidal aggregates in shear flow’, Journal of Colloid Interface Science, Vol. 157, pp. 399-410. Potanin, A.A. and Uriev, N.B. 1991, ‘Microrheological models of aggregated suspensions in shear flow’, Journal of Colloid Interface Science, Vol. 142, pp. 385-395. Purchas, D.B. 1981, Solid/Liquid Separation Technology, Uplands Press, England. Randolph, A.D. and Larson, M.A. 1988, Theory of Particulate Processes, 2nd ed., Academic Press, N.Y. Ray, D.T. and Hogg, R. 1987, ‘Agglomerate breakage in polymer-flocculated suspensions’, Journal of Colloid Interface Science, Vol. 116, pp. 256-268. Reid, K.J. and Stachowicz, M.S. 1990, ‘Evaluation of the Par-tec 200 particle size analyser’, Control '90 - Mineral and Metallurgical Processing, R.K. Rajamani and J.A. Herbst (Eds.), Soc. Min., Metall., Expl.: Littleton, Colorado, pp. 143-149. Reynolds, P.A. and Jones, T.E.R. 1989, ‘An experimental study of the settling velocities of single particles in non-Newtonian fluids’, International Journal of Mineral Processing, Vol. 25, pp. 47-77. Rho, T., Park, J., Kim, C., Yoon, H.-K. and Suh, H.-S. 1996, ‘Degradation of polyacrylamide in dilute solution’, Polym. Degrad. Stability, Vol. 51, pp. 287-293. Richardson, J.F. and Zaki, W.N. 1955, ‘Sedimentation and fluidisation’, Trans. Instn. Chem. Engrs., Vol. 32, pp. 35-53. Ritchie, R. 1965, Certain Aspects of Flocculation as Applied to Sewage Purification, Ph.D Dissertation, London University. Roberts, E.J. 1949, ‘Thickening – art or science?’, Trans. Amer. Inst. Min. Mettal. Engng., Vol. 184, pp. 61-64. Ruehrwein, R.A. and Ward, D.W. 1952, ‘Mechanism of clay aggregation by polyelectrolytes’, Soil Sci., Vol. 73, pp. 485.

Page 204: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

203

Rushton, A., Ward, A.S. and Holdich, R.G. 1996, Solid-Liquid Filtration and Separation Technology, VCH, Germany. Russel, W.B., Saville, D.A. and Schowalter, W.R. 1989, Colloidal Dispersions, Cambridge University Press, Cambridge. Saffman, P. and Turner, J. 1956, ‘On the collision of drops in turbulent clouds’, Journal of Fluid Mechanics, Vol. 1, pp. 16-30. Salkeld, J.A., Hunt, A., Thorn, R., Dickin, F.J., Williams, R.A., Conway, W.F. and Beck, M.S. 1991, ‘Tomographic imaging of phase boundaries in multi-component processes’, in Williams, R.A. and de Jaeger, N.C. (Eds.): Advances in Measurement and Control of Colloidal Processes, Butterworth-Heinemann, Oxford, pp. 95-106. Schramm, L.L. (Ed.), 1996, Suspensions, Fundamentals and Applications in the Petroleum Industry, American Chemical Society, USA. Sengupta, D.K., Kan J., Al Taweel, A.M. and Hamza, H.A. 1997, ‘Dependence of separation properties on flocculation dynamics of kaolinite flocculation’, Int. J. Miner. Process., Vol. 49, pp. 73-85. Serra, T. and Casamitjana, X. 1998a, ‘Effect of the shear and volume fraction on the aggregation and breakup of particles’, AIChE Journal, Vol. 44, No. 8, pp. 1724-1730. Serra, T. and Casamitjana, X. 1998b, ‘Modelling the aggregation and break-up of fractal aggregates in a shear flow’, Applied Scientific Research, Vol. 59, No. 2-3, pp. 255-268. Serra, T. and Casamitjana, X. 1998c, ‘Structure of the aggregates during the process of aggregation and break up under shear flow’, Journal of Colloid and Interface Science, Vol. 206, No. 2, pp. 505-511. Seville, J.P.K., Tüzün, U. and Clift, R. 1997, Processing Particulate Solids, Blackie Academic & Professional, Melbourne. Shah, S.N. and Lord, D.L. 1991, ‘Critical velocity correlations for slurry transport with non-Newtonian fluids’, AIChE Journal, Vol. 37, No. 6, pp. 683-870. Shamlou, P.A. (Ed.) 1993, Processing of Solid-Liquid Separations, Butterworth-Heinemann LTD, UK. Shamlou, P.A. and Titchener-Hooker, N. 1993, ‘Aggregate behaviour in stirred vessels’, in Shamlou, P.A. (Ed.): Processing of Solid-Liquid Separations, Butterworth-Heinemann LTD, UK. Sharma, S.K., Stanley, D.A. and Harris, J. 1994, ‘Determining the effect of physicochemical parameters on floc size using a population balance model’, Minerals and Metallurgical Processing, Vol. 11, pp. 168-173. Shaw, D.J. 1993, Introduction to Colloid & Interface Chemistry, fourth edition, Butterworth and Heinemann. Sikora, M.D. and Stratton, R.A. 1981, ‘The shear stability of flocculated colloids’, Tappi, Vol. 64, No. 11, pp. 97-101. Simmons, M.J.H., Langston, P.A. and Burbidge, A.S. 1999, ‘Particle and droplet size analysis from chord distributions’, Powder Technology, Vol. 102, pp. 75-83. Simons, S.J.R. 1996, ‘Modelling of agglomerating systems: From spheres to fractals’, Powder Technology, Vol. 87, No.1, pp. 29-41. Smit, D.J., Hounslow, M.J. and Paterson, W.R. 1994, Aggregation and Gelation - I. Analytical solutions for CST and batch operation, Chem. Eng. Sci., Vol. 49, pp. 1025-1035.

Page 205: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

204

Smit, D.J., Hounslow, M.J. and Paterson, W.R. 1995, Aggregation and gelation - III. Numerical classification of kernels and case studies of aggregation and growth, Chem. Eng. Sci., Vol. 50, No. 5, pp. 849-862. Smit, D.J., Paterson, W.R. and Hounslow, M.J. 1994, Aggregation and gelation - II. Mixing effects in continuous flow vessels’, Chem. Eng. Sci., Vol. 49, pp. 3147-3167. Smith, D.K.W. and Kitchener, J.A. 1978, ‘The strength of aggregates formed in flocculation, Chem. Eng. Sci., Vol. 33, pp. 1631-1636. Smoluchowski, M. 1916, ‘Drei Vortrage uber diffusion brownsche bewegung und koagulation vol kolloidteilchen’, Physick Zeitschrift, Vol. 17, pp. 557. Smoluchowski, M. 1917, ‘Versuch einer mathematischen theorie der koagulationskinetic kolloider lösungen’, Z. Phys. Chem., Vol. 92, pp. 129-168. Sonntag, R.C. 1993, ‘Coagulation kinetics’, in: Dobias, B. (Ed.), Coagulation and flocculation: Theory and applications, Marcel Dekker, USA. Sonntag, R.C. and Russel, W.B. 1986, ‘Structure and breakup of flocs subjected to fluid stresses I. Shear experiments’, Journal of Colloid and Interface Science, Vol. 113, pp. 399-413. Sparks, R.G. and Dobbs, C.L. 1993, ‘The use of laser backscatter instrumentation for the on-line measurement of the particle size distribution of emulsions’, Part. Part. Syst. Charact, Vol. 10, pp. 279-289. Spears, D.R. and Stanley, D.A. 1994, ‘A study of shear-flocculation of silica’, Minerals and Metallurgical Processing, Vol. 11, pp. 5-11. Spells, K.E. 1955, ‘Correlations for use in transport of aqueous suspensions of fine solids through pipes’, Trans. Instn. Chem. Engrs., Vol. 33, pp. 80-84. Spicer, P.T. 1997, Shear-induced aggregation-fragmentation: mixing and aggregate morphology effects, Ph.D. thesis, Department of Chemical Engineering, University of Cincinnati, USA. Spicer, P.T. and Pratsinis, S.E. 1996a, ‘Coagulation and fragmentation: Universal steady-state particle-size distribution’, AIChE Journal, Vol. 42, No. 6, pp. 1612-1620. Spicer, P.T. and Pratsinis, S.E. 1996b, ‘Shear-induced flocculation - the evolution of floc structure and the shape of the size distribution at steady state’, Water Research, Vol. 30, No. 5, pp. 1049-1056. Spicer, P.T., Keller, W. and Pratsinis, S.E. 1996, ‘The effect of impeller type on floc size and structure during shear-induced flocculation’, Journal of Colloid and Interface Science, Vol. 184, pp. 112-122. Spicer, P.T., Pratsinis, S.E., Raper, J., Amal, R., Bushell, G. and Meesters, G. 1998, ‘Effect of shear schedule on particle size, density, and structure during flocculation in stirred tanks’, Powder Technology, Vol. 97, pp. 26-34. Spicer, P.T., Pratsinis, S.E., Trennepohl, M.D. and Meesters, G.H.M. 1996, ‘Coagulation and fragmentation: the variation of shear rate and the time lag for attainment of steady state’, Industrial & Engineering Chemistry Research, Vol. 35, No. 9, pp. 3074-3080. Spielman, L.A. 1978, ‘Hydrodynamic aspects of flocculation’ in The Scientific Basis of Flocculation, Ives, K.J. Ed. Sijthoff and Noordhoff, The Netherlands, pp. 207-233. Stanley-Wood, N.G. and Lines, R.W. 1992, Royal Society of Chemistry, Cambridge, England. Steinour, H.H. 1944a, ‘Rate of sedimentaion, nonflocculated suspensions of uniform spheres’, Industrial and Engineering Chemistry, Vol. 36, No. 7, pp. 618-624.

Page 206: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

205

Steinour, H.H. 1944b, ‘Rate of sedimentaion, suspensions of uniform-size angular particles’, Industrial and Engineering Chemistry, Vol. 36, No. 9, pp. 840-847. Steinour, H.H. 1944c, ‘Rate of sedimentaion, concentrated flocculated suspensions of powders’, Industrial and Engineering Chemistry, Vol. 36, No. 10, pp. 901-907. Sternberg, R.W., Ogston, A. and Johnson, R. 1996, ‘A video system for in situ measurment of size and settling velocity of suspended particles’, Journal of Sea Research, Vol. 36, pp. 127-130. Stokes, G.G. 1851, ‘On the effect of the intertial friction of fluids on the motion of pendulums’, Trans. Cambridge Philos. Soc., Vol. 9, pp. 8-106. Stoll, S. and Buffle, J. 1998, ‘Computer simulation of flocculation processes: The roles of chain conformation and chain/colloid concentration ratio in the aggregate structures’, Journal of Colloid and Interface Science, Vol. 205, pp. 290-304. Stratton, R.A. 1983, ‘Effect of agitation on polymer additives’, Tappi Journal, Vol. 66, No.3, pp. 141-144. Strenge, K. 1993, ‘Structure formation in disperse systems’, in: Dobias, B. (Ed.), Coagulation and Flocculation: Theory and Applications, Marcel Dekker, USA, pp. 265-320. Stumm, W. 1992, Chemistry of the Solid-Water Iinterface, John Wiley and Sons. Stumm, W. and Morgan, J.J. 1996, Aquatic Chemistry, (third ed.) Wiley-Interscience, New York, Suharyono, H. and Hogg, R. 1996, ‘Flocculation in flow through pipes and in-line mixers’, Minerals and Metallurgical Processing. Vol. 13, pp. 93-97. Svarovsky, L. (Ed.), 2000, Solid-Liquid Separation, 4th edition, Butterworth-Heinmann, UK. Swift, D.L. and Friedlander, S.K. 1964, ‘The coagulation of hydrosols by Brownian motion and laminar shear flow’, Journal of Colloid Interface Science, Vol. 19, pp. 621-647. Tadayyon, A. and Rohani, S. 1998, ‘Determination of particle size distribution by Par-Tec 100: modeling and experimental results’, Part. Part. Syst. Charact. Vol. 15, pp. 127-135. Talmage, W.P. and Fitch, E.B. 1955, ‘Determining thickener unit areas’, Ind. Engng. Chem., Vol. 57, pp. 38-41. Tambo, N. and François, R.J. 1991, ‘Mixing, breakup, and floc characteristics’, in: Amirtharajah, A. et al. (Ed.) Mixing in Coagulation and Flocculation, American Water Works Association Research Foundation, USA. pp. 256-281 Tambo, N. and Watanabe, Y. 1979, ‘Physical aspects of flocculation process: I. Fundamental treatise’, Water Res., Vol. 13, pp. 429-439. Tambo, N. and Watnabe, Y. 1979, ‘Physical characteristics of flocs – 1: The floc density function and aluminium floc’, Water Research, Vol. 13, pp. 409-419. Tence, M., Chevalier, J.P. and Jullien, R. 1986, ‘On the measurement of the fractal dimension of aggregated particles by electron microscopy: experimental method, corrections and comparison with numerical models, J. Phys. - Paris, Vol. 47, No. 11, 1989-1998. Tennekes, H. and Lumley, J.L. 1972, A First Course in Turbulence, MIT Press, London. Thomas, D. G. 1964, ‘Turbulent disruption of flocs in small particle size suspensions’, AIChE Journal, Vol. 10, No. 4, pp. 517-523.

Page 207: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

206

Thomas, D.G. 1963, ‘Transport characteristics of suspensions VII. Relation of hindered-settling floc characteristics to rheological parameters’, AIChE Journal, Vol. 9, No. 3, pp. 310-316. Thomas, D.N., Judd, S.J. and Fawcett, N. 1998, ‘Flocculation modelling of primary sewage effluent’, 8th International Gothenburg Symposium in Prague, 7-9th September. Thomas, D.N., Judd, S.J. and Fawcett, N. 1999, ‘Flocculation modelling: a review’, Wat. Res., Vol. 33, No. 7, pp. 1579-1592. Tomi, D. and Bagster, D.F. 1975, ‘A model of floc strength under hydrodynamic forces’, Chemical Engineering Science, Vol. 30, pp. 269-278. Tomi, D. and Bagster, D.F. 1978, ‘The behaviour of aggregates in stirred vessels’, Trans IChemE, Vol. 56, pp. 1-8. Tong, P., Basu, A., Alexander, K.S. and Dollimore, D. 1998, “Development of a new thoery for hindered settling suspensions’, STP Pharma Sciences, Vol. 8, No. 4, pp. 241-247. Torres, F.E., Russel, W.B. and Schowalter, W.R. 1991a, ‘Simulations of coagulation in viscous flows’, Journal of Colloid Interface Science, Vol. 142, pp. 554-574. Torres, F.E., Russel, W.B. and Schowalter, W.R. 1991b, ‘Floc structure and growth kinetics for rapid shear coagulation of polystyrene colloids’, Journal of Colloid Interface Science, Vol. 145, pp. 51-73. Tory, E.M. (Ed.) 1996, Sedimentation of Small Particles in a Viscous Fluid, Bell & Bain Publishers, Glasgow. Tsao, H.K. and Hsu, J.P. 1989, ‘Simulation of floc breakage in a batch-stirred tank’, Journal of Colloid and Interface Science, Vol. 132, No. 3, pp. 313-318. Turian, R.M. and Yuan, T-F. 1977, ‘Flow of slurries in pipelines’, AIChE Journal, Vol. 23, No. 3, pp. 232-243. Turian, R.M., Ma, T.W., Hsu, F.L.G. and Sung, D.J. 1997, ‘Characterization, settling, and rheology of concentrated fine particulate mineral slurries’, Powder Technology, Vol. 93, pp. 219-233. Uchida, S. 1952, ‘Slow viscous flow through a mass of particles’, Industrial and Engineering Chemistry, Vol. 46, No. 6, pp. 1194-1195. Uhl, V.W. and Gray, J.B. (Ed’s), 1967, Mixing: Theory and Practice, Academic Press, Vol. 1, New York. Valentas, K.J. and Amundson, N.R. 1966, ‘Breakage and coalescence in dispersed phase systems’, Ind. Eng. Chem. Fundam., Vol. 5, pp. 533-542. Valentas, K.J., Gilous, O. and Amundson, N.R. 1966, ‘Analysis of breakage in dispersed phase systems’, Ind. Eng. Chem. Fundam., Vol. 5, pp. 271-279. van de Ven, T.G.M. and Mason, S.G. 1977, ‘The microrheology of colloidal dispersions VII. Orthokinetic doublet formation of spheres’, Colloid Polymer Sci., Vol. 255, pp. 468-479. van Diemen, A.J.G. and Stein H.N. 1984, ‘Energy dissipation during flow of coagulated concentrated suspensions’, Powder Technology, Vol. 37, pp. 275-287. van Diemen, A.J.G., Schreuder, F.W.A.M. and Stein H.N., 1985, ‘Rheology of suspensions of hydrophilic and hydrophobic solid particles in nonaqueous media’, Journal of Colloid and Interface Science, Vol. 104, No. 1, pp. 87-94.

Page 208: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

207

Vand, V. 1948, ‘Viscosity of solutions and suspensions. I Theory’, Journal of Physical Chemistry, Vol. 52, pp. 277-299. Vand, V. 1949, ‘Viscosity of solutions and suspensions’, Journal of Physical Chemistry, Vol. 52, pp. 277-299. Vanni, M. 2000, ‘Approximate population balance equations for aggregation-breakage processes’, Journal of Colloid and Interface Science, Vol. 221, pp. 143-160. Veerapaneni, S. and Wiesner, M.R. 1996, ‘Hydrodynamics of fractal aggregates with radially varying permiability’, Journal of Colliod and Interface Science, Vol. 177, pp. 45-57. Verwey, E.J.W and Overbeek, J.T.G. 1948, ‘Theory of the stability of lyophobic colloids’, Elsevier, Amsterdam. Vo, H.X. 1994, ‘Use of an on-line droplet size analyzer to design and scale-up a process for mixing silicone oil and water’, Society of Cosmetic Chemists, September 13, 1994 - Chicago. Vold. M.J. 1963, ‘Computer simulation of floc formation in a colloidal suspension’, J. Colloid Sci., Vol. 18, pp. 684-695. Wagberg, L. and Lindstroem, T. 1987, ‘Kinetics of polymer induced flocculation of cellulosic fibres in turbulent flow’, Colloids Surfaces, Vol. 27, pp. 29-42. Wakeman, R.J. and Holdich, R.G. 1984, ‘Theoretical and experimental modelling of solids and liquid pressures in batch sedimentation’, Filtration and Separation, Downington USA, pp. 420-422. Ward, A.S. 1996, ‘Sedimentation and thickening’, Chapter 7 in: Solid-Liquid Filtration and Separation Technology, VCH, Germany. Wasp, E.J., Kenny, J.P. and Gandhi, R.L. 1977, Solid-Liquid Flow Slurry Pipeline Transportation, Trans Tech Publications, Germany. Webster, J.G. 1990, Electrical Impedance Tomography, Adam Hilger, Bristol. Weitz, D.A., Huang, J.S., Lin, M.Y. and Sung, J. 1985, ‘Limits of the fractal dimension for irreversible kinetic aggregation of gold colloids’, Physical Review Letters, Vol. 54, No. 12, pp. 1416-1419. Wesely, R.J. 1985, ‘Laboratory methods in solid-liquid separation’, Chapter 30 in: SME Mineral Processing Handbook, Ed: Weiss, N.L. American Insitute of Mining, Metallurgical, and Petroleum Engineers, Inc., Kingsport Press, USA. Whittington, P.N. and George, N. 1992, ‘The use of laminar tube flow in the study of hydrodynamic and chemical influences on polymer flocculation of escherichia coli’, Biotech. Bioeng., Vol. 40, pp. 451-458. Wiesner, M.R. 1992, ‘Kinetics of aggregate formation in rapid mix’, Water Res., Vol. 26, pp. 379-387. Wigsten, A.L. and Stratton, R.A. 1984, ‘Polymer adsorption and particle flocculation in turbulent flow’, Polymer Adsorption and Dispersion Stability, ACS Symp. Ser. 240, E.D. Goddard and B. Vincent (Eds.), Am. Chem. Soc., pp. 429-444. Williams, R.A. (Ed.) 1992, Colloid and Surface Engineering, Applications in the Process Industries, Butterworth-Heinmann, UK. Williams, R.A. and Amarasinghe, W.P.K. 1989, ‘Measurement and simulation of sedimentation behaviour of concentrated polydisperse suspensions’, Trans. Instn Min. Metall. Jan.-Apr. No. 98, pp. C68-C82.

Page 209: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

208

Williams, R.A. and Simons, S.J.R. 1992, ‘Handling colloidal materials’, Chapter 2 in: Williams, R.A. (Ed.), Colloid and Surface Engineering, Applications in the Process Industries, Butterworth-Heinmann, UK. Williams, R.A., Peng, S.J. and Naylor, A. 1992, “In situ measurement of particle aggregation and breakage kinetics in a concentrated suspension’, Powder Technology, Vol. 73, pp. 75-83. Wistrom, A. and Farrell, J. 1998, ‘Simulation and system identification of dynamic models for flocculation control’, Wat. Sci. Tech., Vol. 37, No. 12, pp. 181-192. Wolynes, P.G. and McCammon, J.A. 1977, ‘Hydrodynamic effect on the coagulation of porous bipolymers’, Macromolecules, Vol. 10, pp. 86-87. Wynn, J.W. 1996, ‘Improved accuracy and convergence of discretized population balance of Lister et al.’, AIChE Journal, Vol. 42, No. 7, pp. 2084-2086. Yeung, A., Gibbs, A. and Pelton, R. 1997, ‘Effect of shear on the strength of polymer-induced flocs’, Journal of Colloid and Interface Science, Vol. 196, No. 1, pp 113-115. Yoshioka, N., Hotta, Y., Tanaka, S., Naito, S. and Tsugami, S. 1957, ‘Continuous thickening of homogenious flocculated slurries’, Soc. Chem. Engng., Vol. 2, pp. 47-54. Young, S., Stanley, S.J. and Smith, D.W. 2000, ‘Effect of mixing on the kinetics of polymer aided flocculation’, Journal of Water Supply: Research and Technology, Vol. 49, No. 1, pp. 1-8. Zandi, I. and Govatos, G. 1967, ‘Heterogeneous flow of solids in pipelines’, Journal of the Hydraulics Division, Proceedings of the American Society of Civil Engineers, May, pp. 145-159. Zeichner, G.R. and Schowalter, W.R. 1979, ‘Effects of hydrodynamic and colloidal forces on the coagulation of dispersions’, Journal of Colloid Interface Science, Vol. 71, pp. 237-253.

Page 210: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

209

11. Appendix

11.1 Hounslow et al.’s (1988) population balance Hounslow et al. discretised the size interval according as per Batterham et al. (1981):

vv

i

i

+

=1

2 11-1

or

LL

i

i

+

=1

3 2 11-2

where: v = Particle volume (m3) L = Particle length (m) I.e. 2i ≤ vi ≤ 2i+1 The number density within the channel is taken to be constant across the channel, and is given by: n’ = Ni/(2i+1-2i) = Ni/2i The average particle volume within the channel is then given by:

v diii

i

i

i

i

i

i i

i i

=

=

= −

=−

+

+

+

+

∫ 2

22

22

22

2 22

1

1

1

1

ln

ln ln

ln

=

22

i

ln 11-4

Which is slightly less than the geometric mean (2i/0.6666). *N.B. note from text, if the average channel volume was assumed to be the geometric mean, then a normal distribution of three daughter fragments would have a mean at the boundary of the i-2 and i-1 channels, giving a 2:1 breakage distribution. But since 11-4 < geometric mean the distribution becomes greater than 2:1, with the final distribution being a function of the standard deviation of the distribution.

Page 211: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

210

Mechanism 1 Collision involving a small particle with a large particle to give an aggregate in the ith channel. The larger particle has to be in the i-1th channel (2*2i-k < 2i, when k >1), and probably at the upper end of that channel (see Figure 11-1):

Figure 11-1: Mechanism 1

The number of available particles in the i-1th channel is:

aNii−

−1

12 11-5

i.e. the shaded area in Figure 11-1. The rate of aggregation is then given by:

R aN Ni i ji

i a( )

,1

11

12= −

−−β 11-6

I.e., consider the interval a to da, the population density will be:

Naa2

11-7

I.e. the total number of particles will be:

N daaa2

11-8

I.e. the total number of interactions that will cause a new ith sized particle is:

2a

2i 2j+1 2j 2i-1 2i+1

2i-2a

+

Page 212: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

211

dR aN N dai a i j

ii

aa,

( ),

11

112 2

= −−

−β

Hounslow substitutes Njj2for Na

a2 since j ≤ a ≤ j+1, and 11-7:

dR aN N da

R N Nada

R N N a

R N N

R N N

i j i ji

ij

j

i j i jii

jj

i j i jii

jj

i j i jii

jj

j j

i j i jii

j

j

j

j

j

i

i

j

j

,( )

,

,( )

,

,( )

,

,( )

,

,( )

,

( ) ( )

1

2

2

12

21

1

11

11

2

2

11

11

2

2

2

11

11

1 2 2

11

11

1 1

1

1

2 2

2 2

2 2 2

2 22

222

2

+ +

+

+

∫ ∫

=

=

=

= −⎛⎝⎜

⎞⎠⎟

=

−−

−−−

−−−

−−−

+

−−−

β

β

β

β

β ( )

( )

( )

22 2 2 2

2 22 2

2 22 2

21

2 22 2 2 1

2 22 2

2 2 1 2 1

11

11

2 1 2 1

11

11

2 12 1

2 1

11

11

2 1 2 1 2 1

11

11

2 1 2

jj j

i j i jii

jj

j j

i j i jii

jj

jj

j

i j i jii

jj

j j j

i j i jii

jj

j

R N N

R N N

R N N

R N N

+ − −

−−−

+ −

−−−

−+

−−−

− + − +

−−−

= −

= −⎛⎝⎜

⎞⎠⎟

= −

= −

,( )

,

,( )

,

,( )

,

,( )

,

β

β

β

β ( )1

32 2

2

3 2

11

11

2 1

11 1

R N N

R N N

i j i jii

jj

j

i j i j i jj i

,( )

,

,( )

,

=

=

−−−

− −−

β

β

but, want to sum all particles from j = 1 to i-2, i.e.:

R N Ni i j i jj i

j

i( )

,1

1 11

2

3 2= − −−

=

∑ β 11-9

Page 213: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

212

Mechanism 2 This is similar to mechanism 1, but aggregation is between two particles in i-1th size. All aggregation events between two i-1th particles must give an ith sized particle since 2(2i-1) = 2i (i.e. bottom of the ith channel), and 2(2i) = 2i+1 (i.e. the top of the ith channel) - see Figure 11-2.

Figure 11-2: Mechanism 2

( )

( )

dR N N da

dR N da

R N a

R N

R N

R N

i i i iii

i i iii

i i iii

i i iii

i i

i i iii

i

i i iii

i

i

i

i

( ),

( ),

( ),

( ),

( ),

( ),

2 12 1 1 1

11

2 12 1 1

12

12

2

2 12 1 1

12

1 2

2

2 12 1 1

12

11

2 12 1 1

12

112

2 12 1 1

12

2

2

2

22 2

22 1

2

1

1

=

=

=

= −

= −

=

− − −−−

− −−−

− −−−

− −−−

− −−−

− −−

∫∫

β

β

β

β

β

β −−

112i

R Ni i i i

( ),

2 12 1 1 1

2= − − −β 11-10 Once again the half stops the double counting as per 11-1

2i 2i-1 2i+1

Page 214: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

213

Mechanism 3 Loss of particles in the ith size channel when they collide with particles in i-1th size range or smaller. I.e. forming particles in the i+1th channel.

Figure 11-3: Mechanism 3

dR aN Nda

dRNN

a da

RNN

R NN

i j i ji

ijj

i j i ji ji j

j

j

i j i ji ji j

j

i j i j i jj i

,( )

,

,( )

,

,( )

,

,( )

,

.

3

31

3 2 1

3 1

2 2

2

23 2

3 2

=

=

=

=

∫ ∫+

+

+−

− −

β

β

β

β

as per mechanism 1 11-11

Summing for all j:

∑−

=

−−β=1

1

13 23i

j

ijjij,i

)(j,i NNR 11-12

2a

2i+1 2j+1 2j 2i 2i+2

+

Page 215: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

214

Mechanism 4 Death of an ith sized particle by collision with a particle in ith or larger size range. All successful collisions will cause death of ith sized particle.

Figure 11-4: Mechanism 4

R NNi j i j i j,( )

,4 = β

I.e. summing over all j:

∑∞

=

β=ij

jij,i)(

i NNR 4 11-12

In this case the constant of ½ for j = i isn’t needed because each successful collision removes 2 particles. It would be expected then that the overall rate of change in the ith channel would be the sum of 11-9, 11-10, 11-11 and 11-12:

)(i

)(i

)(i

)(i

i RRRRdt

dN 4321 −−+= 11-13

However as the figures above show there is a tendency for the particles to be taken from the top of each channel, and through aggregation with small particles be moved to the bottom of the next channel in mechanisms 1 and 3 (Kumar and Ramkrishna 1997c). Since the model assumes that the distribution is even across the channel (11-3), some adjustment is required to compensate. To do this, Hounslow et al. multiplied mechanisms 1 and 3 by 2/3 giving:

)(i

)(i

)(i

)(i

i RRRRdt

dN 433221

32 −−+= 11-14

2i 2j+1 2j 2i+1

+

Page 216: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

215

Alternatively (Mechanism 1):

111

111

1

111

2

22

22

1

1

+−−−

−−−

−−

β=

⎟⎟⎠

⎞⎜⎜⎝

⎛β=

⎟⎟

⎜⎜

⎛β=

ijijj,i

j

iijj,i

j

i

ijj,i)(

ij

NN

lnln

NN

N.VolN.Vol

NNR

but, want to sum all particles from j = 1 to i-2, ie:

1

2

11

11 2 −

=−

+−∑ β= ij

i

jj,i

ij)(ij NNR 11-15

Page 217: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

216

11.2 Fractal geometry, calculation of dagg and ρagg from dm, dp & Df Given (Abel et al. 1994, Cohen & Wiesner 1990, Manning & Dyer 1999, Flesch et al.1999, Gregory 1997): fD

agg5s dkm = -A1 where: ms = Mass of solid in aggregate (kg) dagg = Effective diameter of the aggregate (m) Df = Mass – Length fractal dimension and for solid objects: 3

ms6s dkm ρ= -A2 where: k6 = Shape dependent constant (e.g. spheres = π/6) ρs = Solid density (calcite = 2710 kg m-3) dm = Mass equivalent diameter (m) Taking the common point at dp the primary particle size, combining Equations A1 & A2 gives:

3p

sDp5p d

6dkm f

πρ== -A3

where: mp = Mass of primary particle (kg) i.e. (Jiang & Logan 1991):

fD3p

s5 d

6k −πρ

= -A4

i.e., combining A1 & A4

ff Dagg

D3p

ss dd

6m −πρ

= -A5

and, equating with A3:

3m

sDagg

D3p

s d6

dd6

ffπρ

=πρ − -A7

i.e.:

3

p

mDpD3

p

3mD

agg dd

dd

dd f

f

f

⎟⎟⎠

⎞⎜⎜⎝

⎛==

− -A8

i.e.:

fD/3

p

mpagg d

ddd ⎟

⎟⎠

⎞⎜⎜⎝

⎛= -A9

Now, considering ρagg, where:

Page 218: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

217

agg

sagg V

m=ρ -A10

substituting A5 and with the effective volume simply given by (again, assuming spherical): 3

aggagg d6

V π= -A11

i.e.:

3agg

Dagg

D3p

s

agg

sagg

d6

dd6

Vm

ff

π

πρ

==ρ

-A12

i.e. (Mills et al. 1991, Jiang & Logan 1991, Kusters et al. 1997, Serra & Casamitjana 1998):

3D

p

aggsagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ=ρ -A13

11.3 Fractal geometry, calculation of effective mean aggregate density For the settling rate equation (Equation 6-10) the density of the aggregate relative to the fluid is required. Starting with equation A10:

lagg

lslagg V

mmρ−

+=ρ−ρ -A14

where: ml = Mass of liquid in the pores of aggregate, size i (kg) ρl = Density of the liquid (water = 1000 kg m-3) The mass of water in the pores is:

( ) ⎟⎠⎞

⎜⎝⎛ π

−π

ρ=−ρ= − ff Di,agg

D3p

3i,aggli,mi,aggli,l dd

6d

6VVm -A15

Combining Equations A5, A12, A14 and A15:

3

i,agg

Di,agg

D3p

3i,aggl

D3p

li,agg

d6

dd6

d6

d6

fff

π

⎟⎠⎞

⎜⎝⎛ π

−π

ρ+ρπ

=ρ−ρ

−−

-A16

i.e.:

Page 219: The Simulation of Polymer Aggregation/Breakage at High ...Hindered settling velocity based on Richardson & Zaki’s (1955) extension of Stokes’ (1851) law ... except with liquid

218

( ) ( )3D

p

agglslagg

f

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛ρ−ρ=ρ−ρ -A17

and, substituting volume weighted mean sizes (Jiang & Logan 1991, Manning & Dyer 1999, Ellis & Glasgow 1999, Huang 1994):

( ) ( )3D

p

agglslagg

f

d

d−

⎟⎟

⎜⎜

⎛ρ−ρ=ρ−ρ -A18


Recommended