You are on page 1of 10

ARTICLE IN PRESS

Chemical Engineering Science 65 (2010) 3937–3946

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Experimental validation of a 1-D continuous thickening model


using a pilot column
Brendan R. Gladman a, Murray Rudman b, Peter J. Scales a,
a
Particulate Fluids Processing Centre, Department of Chemical and Biomolecular Engineering, The University of Melbourne, 3010 Victoria, Australia
b
CSIRO Mathematical and Information Sciences, Private Bag 33, Clayton South, Victoria 3168, Australia

a r t i c l e in fo abstract

Article history: Predictive modelling of solid–liquid separation can greatly assist the design and operation of thickening
Received 20 April 2009 and filtration equipment, improving water recovery and reducing costs. A phenomenological model
Received in revised form describing continuous thickening has been previously developed with primary inputs being the
28 January 2010
material properties, (compressive yield stress and hindered settling function) derived from routine
Accepted 19 March 2010
laboratory batch settling and filtration tests. This work aimed to validate the model by operating a pilot
Available online 25 March 2010
column continuously and measuring the underflow solids. The column was operated at two different
Keywords: solid fluxes and several bed heights. Additionally, the influence of flocculation conditions (polymer
Thickening dosage and residence time) on thickening performance were studied. The model predicted the
Dewatering
experimental underflow solids concentration at a given flux. For the observed underflow solids
Flocculation
concentration, the ratio of the actual to predicted flux was observed to be between a factor of 1
Model validation
Pilot thickening (accurate) and 10. The model was most accurate for the lowest bed heights. This work confirmed the
model was able to correctly predict trends for the case where minimal bed height and shear forces are
present. Deviation from the model is postulated to be due to changes in the dewatering properties of
flocculated aggregates over time that are not adequately captured using conventional batch
sedimentation tests. The data from these tests are traditionally used as a key input to thickening
models.
& 2010 Elsevier Ltd. All rights reserved.

1. Introduction feed-well and usually mixed with a polymer flocculant to increase


the mean particle size. These aggregated solids settle due to a
In many parts of the world, potable or high quality water is a density difference and are removed as a concentrated underflow.
limited resource and the growing demand by industry imposes It is usual that the aggregated particles form a bed in the
an increasing strain on existing sources. Consequently, it is thickener. Depending on operational conditions, this bed may
imperative that water usage is properly managed, particularly in be a percolating (inter-connected) network of aggregates, a
industry, which often represents a significant proportion non-percolating bed or a combination of both. The particle
of the overall water demand for a community. Dewatering or concentration at the point of percolation is known as the gel
solid–liquid separation is important in this respect and has wide point or fg. Clarified liquor overflows the vessel. A rake or scraper
ranging applications in mineral processing, water and wastewater is used to aid the mechanical transport of solids to the underflow
treatment and food manufacturing. Effective dewatering of the vessel.
minimises the volume of water ultimately consumed and at the Despite its apparent simplicity, there are a number of aspects
same time, reduces the volume of solid waste requiring disposal. of gravity thickening that are poorly understood. For example, it
In some arid parts of the world, improved water recovery can has been recognized for a long time that the solids in an
mean the difference between continued operation and enforced operational thickener may exist in a series of zones from top to
plant shutdown. bottom, including a clarification, hindered settling and bed zone
Gravity thickening is a conceptually simple and effective but as inferred above, this bed may exist in a percolating or
means of recovering water for recycle in industry. In this process, non-percolating form. The transition from a non-percolating to a
a particulate slurry is fed to a gravity thickener vessel through a percolating bed depends on the gel point but the relationship of
this parameter to the morphology of the aggregate is still not
clear. In addition, in predicting thickener performance (usually
 Corresponding author. indicated as the solids flux through the thickener to produce a
E-mail address: peterjs@unimelb.edu.au (P.J. Scales). given underflow solids condition), the role of bed height, rake

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.03.029
ARTICLE IN PRESS
3938 B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946

geometry, rake speed, the slope of the walls and the way that the in the material gel point, fg (Gladman et al., 2005). This is
solids are flocculated are usually not considered in an explicit postulated to be due to intra-aggregate densification such that
manner. Whilst any one of these additional parameters may not R(f) is lowered and network failure occurs at a lower applied
be dominant, it is clear that many of the original concepts of stress for a given volume fraction. This postulate has operational
thickening, developed in an era when the use of high molecular support, with the prediction of thickener output observed to
weight polymers to promote fast settling aggregates and high rate depend on the rotational speed of the thickener rake, type of
or paste thickener geometries was not the norm, may need a thickener and cone angle (Shannon et al., 1964; Tory and
re-think. Indeed, in the absence of flocculation and/or raking, the Shannon, 1965; Gladman et al., 2005; Shannon and Tory, 1966;
prediction of thickening based on laboratory sedimentation data Tiller and Chen, 1988; Usher et al., 2005). There is also growing
following the approaches of Coe and Clevenger and thereafter, evidence of improvements in R(f) for systems with an average
Kynch, has proved to be adequate and a good basis for design solids volume fraction less than the gel point. A phenomenolo-
(Shannon et al., 1964; Tory and Shannon, 1965). gical description of dewatering that includes aggregate densifica-
Over the past two decades, several models have been tion at low shear stresses and concentrations below fg has been
developed to simulate dewatering in gravitational thickening presented (Usher et al., 2009).
(Landman et al., 1994; Garrido et al., 2003; Tiller et al., 1987; The aim of this paper is to establish the range of validity of a
Landman et al., 1988). The aim here was to replace the graphical 1-D steady state thickening model by demonstrating that under
methodologies utilised in the original thickening descriptions controlled conditions with low shear stresses the 1-D model
with a numerical approach. All models share a common correctly predicts dewatering. By implication, the work has the
phenomenological basis in that dewatering is described by two aim of providing an operational range of validity for all literature
key compressional rheology functions; a concentration depen- models of thickening where the feed to the thickener is
dent, compressive yield stress, Py(f) and a hindered settling flocculated.
function, R(f). That most existing literature uses this methodol- To approximate one dimensional, steady state conditions with
ogy is not immediately apparent since the terminology utilised low shear stresses, a tall narrow column with no moving internal
here (that of Landman et al., 1994), is not common to all authors. parts was used. Pilot scale experiments provide greater control
A recent paper (de Kretser and Scales, 2007) demonstrates, none over experimental conditions such that feed material and
the less, that they share a common phenomenological basis. flocculation state are both constant and the assumption of steady
The compressive yield stress Py(f) governs the strength of a state is satisfied. Although laboratory scale measurements were
particulate network under unidirectional compression while R(f) considered, they were dismissed as not useful since the
quantifies the hydrodynamic drag experienced by particles compressional and residence time effects typical of full scale
taking into account interaction between other particles. thickening could not be mimicked. Laboratory scale batch settling
The hindered settling function R(f) is inversely related to the experiments were used to establish inputs to the model to give
permeability. Collectively, these material functions govern the the expected thickener performance.
rate and extent to which a suspension will dewater.
The continuous gravitational thickening model, developed by
Landman et al. (1988) and demonstrated for a range of materials
by Usher and Scales (2005), has been applied to several operating 2. Steady state model development
thickeners. Unfortunately, the predictions do not usually match
reality. Comparison of results with operational data showed that Fig. 1 shows schematically, a one-dimensional continuous
for a given underflow solids concentration, the solids flux gravity thickener. Feed is introduced at a height¼z ¼H, with a
predicted by the model was usually under-predicted, in some volumetric flux Qf and solid volume fraction ff. At some height,
instances, by up to two orders of magnitude (Usher, 2002; z¼hb, the solids can reach the gel point, i.e. a solids concentration
Gladman et al., 2006; Usher and Scales, 2009). Data comparing at which the particulate suspension is percolating and can
laboratory measurement and model prediction to full scale output
are surprisingly scarce, suggesting that this observed shortcoming Qo
is widespread. A more in-depth analysis suggests that the
discrepancy predominately occurs when the feed to the system
is flocculated. Given that the majority of slurries fed to thickeners
are now flocculated and further, given that all thickener models
use a similar fundamental basis, the implication is that none are
useful to current operational practice. Qf φf z=H
There is a growing body of evidence that shows that shearing a
consolidating bed, as is the case in a raked thickener, vastly φφ00
improves dewaterability. The effect appears to be minor for un- φ
0
flocculated slurries (Tory and Shannon, 1965) but is more
significant for the case where the thickener feed is flocculated.
φ 0φ<0 <φφ << φφg γ
Although an effect might be expected for un-flocculated slurries z
based on the work of Channell and Zukoski (1997) using ceramic z = hb
particulates in filtration, the fractal nature of flocculated aggre- φφ >> φgφ γ
gates from typical thickener feeds might be expected to yield
greater changes in aggregate morphology upon the introduction
of shear (Heath et al., 2006a). Although flocculated aggregates
experience shear in the thickener feed-well, it is appropriate to
consider shear effects post-formation, where there is no longer
un-adsorbed flocculant and reformation of aggregates after Qu φu
breakage is unlikely. Shearing of flocculated suspensions has
been associated with a measured decrease in R(f) and an increase Fig. 1. Ideal 1-D continuous flow thickener.
ARTICLE IN PRESS
B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946 3939

support a normal stress. At heights greater than z¼ hb, solids fall a function of bed height. A limiting (maximum) solids flux is
in a hindered settling zone, i.e. the non-zero solids concentration obtained by combining free settling and compression calculations
affects the permeability and drag but the solids concentration is and choosing the minimum value (Usher and Scales, 2005).
below the gel point and the compressive yield stress is zero. At In general, it is possible to operate a thickener fed at a constant
heights below z ¼hb, the solids concentration is above the gel solids flux in two steady state modes:
point and the compressive yield stress is non-zero. This zone is
termed the consolidation zone. Fluid, displaced by the settling  Constant bed height (the underflow flow rate is adjusted to
solids, exits the thickener via an overflow stream, at a flux Qo. achieve a constant bed height).
Under steady-state conditions, the solid and fluid fluxes are by  Constant underflow solids (the bed height is adjusted to
definition constant at all heights in the thickener. achieve steady state operation).
If the feed flux (Qf) is fixed, the only relevant operational
variable is the underflow flux, Qu. The magnitude of Qu is
important as it dictates the height of bed, the solids residence Fig. 2 shows the model output using, as an example, a flocculated
time and the solid volume fraction in the underflow, fu. The solid calcite suspension for bed heights ranging between 0.1 and 10 m.
volume fraction in the consolidation zone (0 ozohb) increases The gel point concentration in this instance is at a solids volume
non-linearly with increasing compression and thus bed depth. fraction of approximately 0.18. In this particular example, at
The inputs for the steady state model described by Usher and solids fluxes above  0.3 t h  1 m, 2 the model predicts that bed
Scales (2005) include the following: height will not influence the solids concentration in the
underflow. The solids flux is sufficiently high that solids do not
 constitutive relationships for the compressibility and hindered spend enough time in the thickener to achieve a solids
settling function as functions of the solids volume fraction; concentration at which compressive forces act (i.e. the solids
 the thickener diameter (if the area is not constant for all concentration is below the gel point everywhere). In this case, the
heights, the diameter as a function of height is required); underflow concentration depends solely on the suspension
 the solids volume fraction of the feed (ff); permeability. Thickeners operated in this solids flux regime are
 the density of the solid and fluid phases; and said to be permeability limited whereby the predicted flux at a
 the bed height (hb). given underflow solids depends only on the solids settling
velocity and underflow pumping rate. Operating the thickener
at this limit is not advisable, since a small change in the system,
For these inputs, the model predicts the solids throughput q (i.e.
such as a change in feed composition or flocculant activity, could
flux) as a function of underflow solids concentration in two stages.
result in solids carryover. Moreover, the short residence time
The first stage solves the governing dewatering equation under
limits or excludes, compressive dewatering and consequently,
hindered settling, where Py(f) is universally zero. The second
underflow solids volume fractions may not even exceed the gel
stage considers the consolidation zone, where the solid stress
point. In terms of water recovery for recycle, this is undesirable.
gradient is non-zero.
When the solid flux is less than  0.3 t h  1 m  2, the model
From a material balance, the solids flux is given by
predicts that the underflow concentration will show a depen-
q uðfÞ dence on the bed height. For higher bed heights, the longer solids
¼  ð1Þ
aðhb Þ 1 1 residence times in the device means that compression becomes

f fu the limiting factor in determining underflow solids concentration.
The thickener shape factor (a), defined as In this operational regime, if the compressive force is increased by
 2 increasing the bed height, the residence time will also increase
dðzÞ and the underflow solids concentration is also predicted to
aðzÞ ¼ ð2Þ
dmax increase. The compressibility of the material is crucial; choosing
accounts for the effect of varying thickener area with height while
the settling rate u(f) is given as 10

Drgð1fÞ
uðfÞ ¼ ð3Þ
RðfÞ Permeability Limited
1 Flux Range
Solids Flux, (tonnes hour-1 m-2)

Once the solids form a network, the change in volume fraction Df,
with height is described by a function which relates R(f) and the
solid stress gradient dPy ðfÞ=df to the volume averaged solid
velocity, q(t): 0.1
 
RðfðzÞÞ q fðzÞ
2 aðzÞ
1 Drg fðzÞ
dfðzÞ ð1fðzÞÞ fu
¼ ð4Þ 0.01
dz dPy ðfðzÞÞ Compressibility
dfðzÞ Limited Flux Range
At steady state, z¼ hb and f ¼ fg (denoted as the gel point) at the
0.001
top of the bed, while z¼0 and f ¼ fu at the base of the thickener.
In general, an analytical solution of Eq. (4) for q does not exist.
Instead, the solids flux is evaluated numerically using an iterative
process. The algorithm takes an initial estimate for q and 0.0001
integrates Eq. (4) from z¼0 to hb subject to the boundary 0 0.1 0.2 0.3 0.4 0.5
condition f(0) ¼ fu. The value of q is adjusted until the condition Underflow Solids Concentration,
(v/v)
at the top of the bed (i.e. f(hb) ¼ fg) is satisfied. This is repeated Fig. 2. 1-D thickener model output for a flocculated calcite material; solids flux
for a range of underflow concentrations and bed heights, (dry tonnes solid) versus underflow solids concentration (v/v). The data are for bed
generating a series of curves in which q is plotted against fu as heights ranging from 0.1 to 10 m.
ARTICLE IN PRESS
3940 B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946

conditions to minimise the compressive yield stress, will was introduced via a centrally located pipe at a height of
maximise the solids volume fraction in the thickener underflow. approximately 1.8 m. Overflow was collected by a peripheral
However, as consolidation causes fluid to be displaced, the launder. To minimise shear stresses, the column was operated
permeability or R(f) must also be considered. To validate the without any moving components (i.e. no rakes, pickets or scrapers).
model, fluxes corresponding to both compressibility and perme- A narrow diameter was chosen so that the assumption of 1-D flow
ability limited operational behaviour must be considered. was a good approximation. A schematic of the pilot column flow
circuit is shown in Fig. 3 and 4 shows a photograph of the column.
The interface between the feed and clarified water can be observed.
3. Experimental methods

The experimental program involved two major steps: Table 1


Tall column dimensions.
1. initial experiments in the laboratory to characterise the
Height (m) 2.2
flocculated calcite suspensions, used in the work to produce Truncated cone height (m) 0.205
model parameter inputs and Column diameter (m) 0.285
2. pilot scale continuous dewatering experiments on these Cone angle (degrees) 60
suspensions to validate the range of applicability of the model. Diameter of truncated cone at base (m) 0.05
Cone volume (L) 4.4

3.1. Laboratory characterisation


0.1 % w/w
Fi Flocculant solution
The solid material used in this work was calcite with a
nominal mean particle size of 10 mm (Omyacarb 10). A high Pipe
reactor
molecular weight anionic polyacrylamide/polyacrylate copolymer
∞ ∞
(AN934SH from SNF Floerger) was used to flocculate the calcite.
Calcite suspensions were prepared at a solids concentration of Stock CaC03 slurry Holding tank
0.05 w/w in 40 L batches by gradually adding 2.06 kg of calcite to 5 % w/w
39 L of local tap water (conductivity 0.177 mS cm  1). Using an
overhead stirrer, suspensions were gently agitated for 30 min to Sample point
allow the mixture to achieve thermal equilibrium.
Fig. 3. Pilot tall column flow circuit used to validate the 1-D thickener model.
A 2 g L  1 stock solution of polymer was prepared following a
standard procedure recommended by the supplier. Working solutions
were prepared just prior to flocculation by adding de-ionised water to
produce a concentration of 0.1 g L  1. Excess working polymer
solution was discarded at the end of each day. Stock solutions were
kept at 4 1C and stored for a maximum period of four weeks.
Flocculation was achieved using a linear pipe reactor. Previous
pipe reactor studies have shown that flocculation is highly repro-
ducible (Heath et al., 2006a, b). In this work, a centrifugal pump
delivered 6.3 L min  1 of calcite slurry. A 12.5 mm ID pipe, with two
different pipe lengths, 1.8 and 9.66 m was chosen, resulting in pipe
residence times of 2 and 10 s, respectively. Flocculant was delivered settling /clarification
into the pipe using a peristaltic pump (Masterflex LS). Four flocculant
interface
dose rates were employed (25, 50, 75 and 100 g t  1) requiring flow
rates between 81 and 323 mL min  1 of flocculant solution.
The compressional rheology (i.e. Py(f), R(f)) and gel point of the
resultant suspensions were measured using a combination of settling
tests and pressure filtration experiments as described in previous
work (Lester et al., 2005; de Kretser et al., 2001). Five 1 L measuring
cylinders were filled directly from the pipe outlet limiting the amount bed/settling
of handling and possible shear degradation of the fragile floc interface
structures. The first cylinder was filled to a height (h0) of 0.4 m while
subsequent cylinders were filled to progressively lower heights (0.8h0,
0.6h0, 0.4h0 and 0.2h0). The sediment–supernatant interface h(t) was
monitored for 2 h, after which settling was almost complete. A final
sediment height was measured after 24 h. The transform of h(t) data
to determine Py(f) and R(f) is described in detail elsewhere (Lester
et al., 2005). The functional form of the Py(f) and R(f) inputs to the
thickener model, developed through this experimental and theoretical
analysis, are also described elsewhere (Gladman et al., 2005).

3.2. Pilot column

The pilot column was 2.2 m in height and 0.285 m in internal


diameter. Additional column information is provided in Table 1. Feed Fig. 4. Photo of pilot column used to validate the thickener model.
ARTICLE IN PRESS
B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946 3941

A second interface between settling and consolidating material, which Solids flux density selection: For flocculated calcite, a solids flux
is defined by hb is also noticeable, (although less apparent) just above density of 0.3 t h  1 m  2 is expected to result in permeability
the second ring in Fig. 4. limited thickening (see Fig. 2). The value of the permeability
Calcite suspensions (0.05 w/w) for the pilot column were limited solids flux density for other materials may be different as
prepared in a 500 L make-up tank using tap water. Suspensions particle size, shape, surface properties and density change. Most
were transferred to a feed tank equipped with an overhead mixer. thickeners in the minerals industry tend to operate in a
The feed tank allowed additional batches to be prepared without permeability limited flux regime or at the transition between
interruption to supply or suspension density. The suspension was permeability and compressibility limited operation. For these
fed to the column using a positive displacement pump. As in the conditions, the rheology of the flocculated underflow solids is
laboratory scale tests, the feed was flocculated in a linear pipe such that the suspension yield stress is typically o10 Pa and rake
reactor. The diameter and length of the pipe was the same as for torque will generally be low (Boger et al., 2002). In the pilot
laboratory experiments in all but the lowest flow rate situations. column, a flux density of 0.3 t h  1 m  2 corresponds to a
Shortly after filling the column, three zones became distin- volumetric flow rate of 6.8 L min  1 for a 0.05 w/w concentration
guishable. The top of the column was characterised by a clarified feed slurry.
zone. In the middle zone individual aggregates could be observed A second solids flux density of 0.0375 t h  1 m  2 was chosen to
settling and the lower zone was an opaque region in which solids represent a compressibility limited operational regime. At this
settled at a slower rate. There was a marked interface between the low flux density, the size of the pipe reactor had to be decreased
middle and lower zones that is termed here the ‘‘bed height’’. from 12.5 mm, as the pipe velocity was insufficient to keep the
Note that it is not necessarily the point at which the solids calcite suspended. In addition, the aggregates that formed were
become networked. An apparent (un-networked) bed will form if significantly larger as a consequence of lower shear stresses in the
a concentration is reached at which the hindered settling rate of pipe. To address this issue, a smaller 6 mm ID pipe was chosen.
the aggregates is equal to the downward velocity of the The volumetric flow rate corresponding to a solids flux density of
aggregates due to the action of the underflow solids pump and 0.037 t h  1 m  2 was 0.7 L min  1. Even in the smaller pipe the
if this concentration is less than the gel point. At lower underflow aggregates forming at the lower flux density were still consider-
pumping rates, the bed is likely to be networked. The bed height ably larger and settled significantly faster. For each solids flux
was allowed to increase to approximately 80% of the target height density, we aimed to produce a mean aggregate size that settled
at which point the underflow pump was turned on. This was done with the same velocity. This required modifying the flocculant
to compensate for the slow dynamical response in the column. dosage from 25 to 5.5 g t  1 of solids.
The underflow rate was calculated via a mass balance: Bed height: The effect of varying bed height was examined at
the two flux density conditions. At the higher flux density
f0 F (0.3 t h  1 m  2), four bed heights were investigated: 0.4, 0.6, 0.8
U¼ ð5Þ
fu and 1.0 m. At the lower flux density, (0.037 t h  1 m  2), three bed
heights were investigated; 0.3, 0.5 and 0.75 m. In the lower flux
In practice, the underflow pump speed needed to be periodically density case, a bed height 40.75 m was difficult to maintain due
adjusted to keep the bed height constant due to variation in fu. to operational instabilities and steady state behaviour could not
Over the course of a run, the bed height was maintained within be attained.
10% of the nominal height. Flocculation state: The effect of varying flocculation conditions
Several column experiments were undertaken. Information on was also examined. Two flocculant dose rates were investigated
the solids flux, bed height and flocculation condition for each run for the higher solids flux density (0.3 t h  1 m  2): 25 and 75 g t  1.
is presented in Table 2. For each run, the column was allowed to For each dose rate, 2 pipe residence times (2 and 10 s) were
operate for sufficient time to reach steady state before sampling considered. The bed was maintained at 1 m. Table 2 summarises
commenced. As a guide to estimate the time scale to achieve the experiments and the parameters considered.
steady state, two times the average residence of solids in the
column was used. On this basis, steady state conditions would
take anywhere between 30 min and 2 h depending on the bed 4. Results and discussion
height. After this time had elapsed, underflow samples were
collected at 30 min intervals. The solids concentration in 4.1. 1-D modelling results
underflow samples was determined by drying at 100 1C
overnight. No solids were recorded in the overflow for any Fig. 5 shows the 1-D model results for solids flux density
conditions used in the experiments and the supernatant was versus underflow concentration (on a solid volume basis) for
always observed to be clear. flocculated calcite using material inputs generated through
laboratory characterisation. The curves shown relate to a calcite
Table 2 feed flocculated inline with a polymer dose of 25 g t  1, 2 s
Summary of pilot column experiments. residence time in the pipe reactor and volumetric flow rate of
6.3 L min  1 (i.e. runs 1–4). The results for six different bed
Run Flux Bed height Flocculant dosage rate Reaction time in
(t m  2 h  1) (m) (g t  1) pipe (s)
heights, (0.1, 0.2, 0.4, 0.6, 0.8 and 1.0 m) are shown. The number of
bed heights covered is in excess of the experimental cases. The
1 0.3 0.4 25 2 model predicts that with the exception of the 0.1 m case, the
2 0.3 0.6 25 2 height of the bed should not affect the final solids concentration
3 0.3 0.8 25 2
for solid fluxes 40.3 t h  1 m  2. The obvious difference between
4 0.3 1.0 25 2
5 0.3 1.0 25 10 the 0.1 m and higher bed height predictions is related to the fact
6 0.3 1.0 75 10 that the bed resides within the cone of the pilot column and there
7 0.037 0.3 5 2 is a reduced bed surface area in this case.
8 0.037 0.5 5 2 Fig. 6 shows a series of curves of predicted solids flux density
9 0.037 0.7 5 2
versus underflow solids concentration for the various pilot
column experiments outlined in Table 2 (Runs 4–7). The model
ARTICLE IN PRESS
3942 B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946

1 The thickening model predicts that, for solid fluxes 40.003 t


h  1 m  2, an increase in polymer dose will result in a higher
underflow solids concentration (Fig. 6). At solid fluxes less than
0.003 t h  1 m  2 the curves appear to asymptote to the same
Solids Flux (tonnes hour-1 m-2)

solids concentration. This is representative of the compressibility


limit of the material. The largest shift in dewatering behaviour
from a change in flocculation conditions occurs in the solid range
0.15–0.2 v/v and is consistent with the postulate that the main
role of flocculation in dewatering is to aid suspension perme-
0.1 ability at low solids, near to or below the gel point.
The predictions shown in Fig. 6 clearly show the importance of
flocculation conditions on dewatering. For example, at a solids
flux density of 0.1 t h  1 m  2, the predicted underflow concentra-
tion for a 10 s residence time and 25 g t  1 polymer dose is 0.1 v/v.
When the polymer dosage is increased to 75 g t  1 the solid
underflow concentration increases to 0.27 v/v. Looking at it
another way, to achieve a desired underflow solids concentration
of say 0.3 v/v, the predicted solids flux density with a 75 g t  1 dose
0.01
is approximately 10 times higher than the equivalent flux density
0 0.1 0.2 0.3 0.4 0.5
if a 25 g t  1 dose was applied. The flocculation residence time is
Underflow Solids Concentration, φ (v/v)
also significant and depends on the polymer dose. For the 25 g t  1
Fig. 5. Flux curve predicted from characterisation of calcite flocculated in a polymer dosage, a longer residence time (10 s) produced a less
12.5 mm linear pipe reactor with a 2 s residence time. Data represent the following optimum result than the shorter (2 s) residence time. For the
bed heights: & 0.1 m; } 0.3 m; D 0.4 m; x 0.5 m; * 0.6 m; 3 0.8 m; and + 1.0 m. 75 g t  1 polymer dosage, a reverse trend was observed although
the effect was not as significant. This suggests that the mechanical
strength of the aggregates is important in producing an optimal
100 dewatering result under different flux conditions.
25 g/t 2 s
4.2. Tall column validation
10 75 g/t, 2 s
Solids Flux, (tonnes hour-1 m-2)

25 g/t, 10 s Fig. 7 plots the underflow solids concentration against time for
pilot column runs 1–4 and Fig. 8 for runs 7–9. Fig. 9 plots the data
1 75 g/t, 10 s
for run 6. The results show that under certain operational
conditions, even after 6 h of operation, the underflow solids
0.1 concentration was still fluctuating. Therefore, it is possible that
steady state conditions were not fully developed in some
instances. The fluctuation was more significant at the higher
0.01 bed heights. Table 3 presents that for a one-metre bed, a 6 h run
time corresponds to more than six times the average particle
residence time in the column, well beyond the criteria for steady
0.001

0.19
0.0001
0 0.1 0.2 0.3 0.4 0.5
Underflow Solids Concentration, φ (v/v)

Underflow Solids Concentration,


(v/v) 0.17

Fig. 6. Predicted solids flux as a function of underflow solids concentration for


four different flocculation conditions (two flocculant dose rates and two residence 0.15
times). All data are for a bed height of 1 m.

0.13

inputs for these predictions come from the laboratory batch 0.11
settling experiments. For clarity, only data for a 1.0 m bed height
are shown. The flocculation conditions were chosen to produce
0.09 0.4 m bed height
different aggregate structures with significantly different
0.6 m bed height
dewatering characteristics. A population balance (PB) model
0.8 m bed height
that predicts the evolution of aggregate size as a function of 0.07 1.0 m bed height
pipe residence time was used to assist the choice of residence
times (Heath et al., 2006a, b). Based on the hydrodynamic
conditions in the pipe reactor, the PB model predicted a 0.05
maximum aggregate size at t ¼1.8 s. Thus, a 2 s residence time 0 1 2 3 4 5 6
was chosen, to ensure aggregate size was no longer increasing. Time (hours)
The 10 s residence time was chosen to produce smaller and Fig. 7. Underflow solids concentration as a function of time and bed height for
potentially denser aggregates that were presumably more robust flocculated calcium carbonate in the pilot scale column at a solids flux density of
to breakage. 0.3 t h  1 m  2 (runs 1–4).
ARTICLE IN PRESS
B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946 3943

0.45 two times the average particle residence time in the column
equates to 2000 min or 33.3 h. Fig. 9 shows the underflow solids
concentration was reasonably constant during run 6, varying
0.4
Underflow Solids Concentration, φ (v/v)

between 0.20 and 0.24 v/v with an average of 0.21 v/v. When the
flocculant dosage was reduced to 25 g t  1 the underflow density
0.35 decreased to 0.14 v/v.
Fig. 10 reproduces the solid flux density curves of Fig. 5 with
the measured underflow results from the pilot column
0.3 superimposed. The discrete data points correspond to fu at the
completion of each run at 0.3 t h  1 m  2 (runs 1–4) and
0.25 0.037 t h  1 m  2 (runs 6–9).
For both low and high fluxes, the data point corresponding to
the lowest solids concentration is associated with the lowest bed
0.2 0.3 m bed height height and the data points corresponding to higher solids
concentrations are associated with progressively increasing bed
0.5 m bed height
heights (as shown in Figs. 7 and 8). In terms of the solids flux
0.15 0.75 m bed height density associated with a given underflow solids concentration,
the difference between the experimental data and the model
0.1 prediction was very small for the lowest bed heights (0.35 and
0 1 2 3 4 5 0.4 m). At higher bed heights, this difference increased to a ratio of
Time (hours) 1.5 and 1.25 for the highest and lowest solids flux density cases,
respectively. This ratio can be viewed as an apparent permeability
Fig. 8. Underflow solids concentration as a function of time and bed height for
enhancement (PE). Values of the PE ratio of up to 200 have been
flocculated calcium carbonate in the pilot scale column at a solids flux density of
0.0375 t h  1 m  2 (runs 7–9). measured for full scale operations involving paste thickening
(Stickland et al., 2005).
The data of Fig. 10 are interesting on a number of counts. The
0.4 first point of interest is the bed height dependence of the data at
the higher flux density condition of 0.3 t h  1 m  2 (runs 1–4). This
0.35 occurs in a regime where dewatering is believed to be perme-
Underflow Solids Concentration, φ (v/v)

ability limited and consequently no bed height dependence is


predicted. The underflow solids concentration in this case
0.3
is always less than the gel point for flocculated calcite (fg is
nominally around 0.18 v/v). Given that it is not possible to have a
0.25 networked bed of particles at concentrations less than the gel
point, then the traditional concept of a ‘bed’ of particles as being
0.2 networked must be abandoned. The concentration at which a bed
forms in the column was discussed earlier in terms of the point at
0.15 which the settling rate of the individual aggregates is less than the
downward flux due to underflow pumping. At low solids flux

0.1

1
0.05
Operating solids flux
0 (0.3 t h-1 m-2)
200 250 300 350 400
Time (minutes)
Solids Flux (t h-1 m-2)

Fig. 9. Underflow solids concentration as a function of time for the pilot column
operated with a constant 1 m bed and feed flocculated at 75 g t  1 in a pipe reactor
for a residence time of 10 s (run 6).
0.1

Table 3
Calculated residence time of solids in the pilot column for aggregates formed
through flocculation in a pipe reactor with a residence time of 2 s.

Solids flux Residence time (min) Operating solids flux


0.4 m 0.6 m 0.8 m 1.0 m (0.0375 t h-1 m-2)

2 1
0.3 t m h 17 32 48 57
0.037 t m  2 h  1 397 698 1000 – 0.01
0 0.1 0.2 0.3 0.4 0.5
Underflow Solids Concentration, φ (v/v)

Fig. 10. Solids flux of a flocculated calcium carbonate slurry as a function of


state that was used in setting up the experimentation. The lower underflow solids concentration for: 0.1 m; 0.3 m; 0.4 m; 0.5 m; 0.6 m; 0.8 m; and
flux experiments also showed variation with respect to underflow 1.0 m bed heights (shown as lines in order from left to right—see Fig. 5 for details).
concentration. However, this is not surprising as at the lower flux, The 0.3 and 0.037 t h  1 m  2 flux density column data are given as discrete circles.
ARTICLE IN PRESS
3944 B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946

density, it is possible that a networked bed forms but at high 100


solids flux density, it is likely that the bed will be un-networked
(certainly at the top of the bed and perhaps throughout). In the
operation of the column, an opaque interface is observed, above 10

Solids Flux, (tonnes hour-1 m-2)


which individual aggregates are observed to settle. In this
instance, the bed is not networked. The only way that a bed
height dependence of the underflow density can be achieved in 1
the pilot column at this flux density condition is if the aggregate
structure is changing as a result of particle interaction and/or
0.1
residence time in the un-networked bed. These differences that
result from bed height implies that this phenomenon is not
present in the simple batch settling tests that are used in the
0.01
model prediction.
For low bed heights, the agreement between the predicted and
Feed
measured underflow solids is very good and the PE ratio is close to 0.001 Concentration
unity. The improved permeability could be explained by osten-
sibly, the aggregates becoming slowly denser. Alternative ex-
planations such as aggregate breakup is less likely as an 0.0001
explanation since smaller particles of the same density would 0 0.1 0.2 0.3 0.4 0.5
likely settle slower. Since the model prediction is based on batch Underflow Solids Concentration,
(v/v)
sedimentation tests, where the time scale of the experiment for
Fig. 11. Predicted underflow solid concentration versus solids feed flux for a pipe
which solids concentrations are below the gel point is short (of reactor with 10 s of residence time;  75 g t  1 flocculant;+ 25 g t  1 flocculant; ’
order seconds), a clear difference/discrepancy emerges between underflow solids from pilot column (75 g t  1 flocculant); m underflow solids from
the characterisation methodology utilised in the laboratory and pilot column (25 g t  1 flocculant).
settling behaviour at larger scale in the tall column where the
residence time of the solids at concentrations below the gel point
is of order minutes to hours.
The above analysis assumes that the physics of the calcite 25 g t  1 dose, the predicted underflow (0.075 v/v) was
sedimentation is well-characterised by R(f) and Py(f). In the context considerably lower than the measured value (0.139 v/v). The
of the agreement of the data for the lowest bed heights, the model difference for the 25 g t  1 case is significant (the PE ratio is  10)
appears valid. Since there is no networked bed in the high flux density and once again, is representative of a case where the underflow
case, this in itself merely represents correlation with the basic physics solids of the column are less than the suspension gel point. The
proposed by Coe and Clevenger (1916). Therefore, for the low bed residence time in the column in this case is calculated to be of
height scenario, the model and characterisation methodology are order 1 h but the majority of this time is spent in an un-
sound. The lack of correlation at higher bed heights (PE ratio of up to networked bed where there is ample opportunity to achieve
1.5 in the pilot column experiments) is reflective of either a new piece aggregate densification from shear stresses associated with
of physics that is not included in the original sedimentation analysis aggregate interaction. Little is known of the mechanics of this
or a change in the material properties with time. Since the extraction process and it is the subject of on-going research. Initial
of the R(f) data from the batch settling data is internally self unreported results suggest that the aggregates become less
consistent (the process re-predicts the original data), the latter fractal during this process.
explanation is the more likely. In the case studied herein and presented in Fig. 11, the
In the lower flux density 0.037 t h  1 m  2 (runs 6–9) case, a aggregates formed at 25 g t  1 polymer dose are weaker and thus
slight dependence on bed height is predicted but the observed are postulated to be more susceptible to densification. The higher
difference between the model output and data from the pilot polymer dose, while offering benefits with respect to preventing
column is once again greatest at the largest bed heights (and shear breakage during aggregate formation, will also inhibit
longest residence times). The agreement at low bed heights is restructuring at the low shear stresses associated with sedimen-
very good (PE ratio close to 1) and in this instance, it is not simply tation. The difference in underflow solids in this instance is thus
a correlation with Coe and Clevenger (1916) since the data for the explained not in terms of compression, but rather in terms of the
underflow solids are above the suspension gel point (i.e. a average time flocculated solids spend in the column and the
networked bed exists) and the column was being operated in propensity of the aggregates to restructure at low shear stress.
the compressibility limited regime. The longer the time, the greater the extent to which the
On the basis of the comparison presented, the model is valid aggregates can densify. The limit of this effect is unknown but
for low bed heights and in the context of the discussion, appears from the data, the time scale is of the order minutes to hours (not
valid across a broad range of operational conditions. The poor seconds). This is very different to the time scale associated with
agreement between model and pilot column data at higher bed aggregate growth and breakage that occurs at high shear stresses
heights is concluded to be a materials characterisation issue that (relatively) in the pipe reactor. At the lower solids flux density
is related to changes in the dewatering parameters as a function condition, a bed height dependence of the underflow solids is also
of time in the un-networked bed in the high flux density case and noted. It is not possible in this instance to decouple effects in the
the networked bed in the low flux density case. These material hindered settling and bed zones (above the gel point in this
changes that are not well considered in the short time scale of instance).
hindered settling in simple batch settling tests. In terms of modelling continuous gravity thickening, the data
Fig. 11 shows the predictions of the experimentally measured presented herein are consistent with the statement that the
solids underflow density at two different polymer dose rates (runs methods used to characterise the dewaterability of suspensions,
6 and 7). The model predicted underflow density for the higher namely batch sedimentation and pressure filtration, do not
polymer dose rate 75 g t  1 was quite close to the measured adequately capture the transient effects of hydrodynamic condi-
underflow density (0.21 (v/v)) (PE ratio was  1.25). For the tions on the permeability and compressibility at longer residence
ARTICLE IN PRESS
B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946 3945

times. This was magnified in this case where flocculant dosages jg value of j at the gel point, m3 m  3
were low or solid residence times were high. j0 initial value of j, m3 m  3

5. Conclusions

Pilot scale continuous gravity settling tests have been conducted Acknowledgements
for a calcite suspension and the data compared to model predictions
based on a dewatering characterisation of flocculated calcium This work was performed with the support of a consortium of
carbonate performed in the laboratory. Two flux conditions were companies coordinated through the Australian Mineral Industries
investigated, these being in the permeability (high flux) and Research Association (AMIRA) as project 266D and 266E. The
compressibility limited (low flux) operational regime. financial and in-kind support of Albian Sands Energy, Alcan
The agreement between model prediction and the column International, Alcoa World Alumina, AngloGold Ashanti, Anglo
underflow solids concentration for a range of operational Platinum, Aughinish Alumina, Bateman Engineering BV, BHP-
conditions was observed to be reasonable with the predicted Billiton, Cable Sands, Ciba Specialty Chemicals, Cytec Australia
solids flux density showing a good agreement at the shortest Holdings, De Beers Consolidated Mines, FL Smidth Minerals,
residence times and lowest bed heights in the column. The Glencore AG, Hatch Engineers, Iluka Resources, Kumba Resources,
agreement became poorer for longer residence time and higher Lightnin, Minara Resources, Metso Minerals, Nalco Chemicals,
beds. The solids underflow concentration from the column was OMG, Outotec, Pechiney Aluminium, Phelps Dodge, Queensland
always the same or higher than that predicted on the basis of Alumina, Queensland Nickel, Rio Tinto, WMC Resources and Zinifex
laboratory dewaterability characterisation. The deviation from the Limited through the AMIRA coordinated project is gratefully
model at longer residence times is postulated to be due to acknowledged. The support of the Particulate Fluids Processing
changes in the calcite aggregates over time, a factor that is not Centre, a Special Research Centre of the Australian Research
incorporated in the current model. It is concluded that the batch Council is also acknowledged. The help and advice of Kosta Simic
sedimentation and filtration methods that are commonly used to in the pilot experiments and Phil Fawell is greatly appreciated.
characterise dewatering for the prediction of gravity thickening
are not adequate for the task in the case of flocculated aggregates.
Additional theory is required to describe the change in aggregate References
structure with time and new experimental techniques are
Boger, D.V., Scales, P.J., Sofra, F., 2002. Rheological concepts. In: Jewell, R., Fourie,
required to characterise this change. Such aggregates are common
A., Lord, E. (Eds.), Paste and Thickened Tails A Guide. UWA Press, Perth,
to a wide range of dewatering operations in a wide range of Australia, pp. 23–34.
industries and suitable models and characterisation techniques Channell, G.M., Zukoski, C.F., 1997. Shear and compressive rheology of aggregated
are both pressing requirements. alumina suspensions. A.I.Ch.E. Journal 43, 1700.
Coe, H.S., Clevenger, G.H., 1916. Methods for determining the capacities of slime
settling tanks. AIME Transactions 55, 356–384.
de Kretser, R.G., Scales, P.J., 2007. Linking local and average filtration and
Notation dewatering parameters from traditional, fluid mechanical and geotechnical
theories. Transactions of the Filtration Society 7, 60–66.
de Kretser, R.G., Usher, S.P., Scales, P.J., Landman, K.A., Boger, D.V., 2001. Rapid
d thickener diameter, m measurement of dewatering design and optimization parameters. A.I.Ch.E.
dmax maximum value of the thickener diameter, m Journal 47, 1758–1769.
F solids throughput in the thickener feed, m3 s  1 Garrido, P., Concha, F., Bürger, R., 2003. Settling velocities of particulate systems:
14. Unified model of sedimentation, centrifugation and filtration of flocculated
g gravitational acceleration, m s  2 suspensions. International Journal of Mineral Processing 72, 57.
hb bed (interface) height, m Gladman, B., de Kretser, R.G., Rudman, M., Scales, P.J., 2005. Effect of shear on
hN equilibrium bed (interface) height, m particulate suspension dewatering. Chemical Engineering Research & Design
83, 1–4.
h0 initial suspension (interface) height, m
Gladman, B., Usher, S.P., Scales, P.J., 2006. Understanding the thickening process.
Dp applied pressure, kPa In: Ninth International Seminar on Paste and Thickened Tailings, Paste 2006,
Py compressive yield stress, kPa Limerick, Ireland.
Heath, A.R., Bahri, P.A., Fawell, P.D., Farrow, J.B., 2006a. Polymer flocculation of
q solids throughput flux density, kg s  1 m  2
calcite: experimental results from turbulent pipe flow. A.I.Ch.E. Journal 52,
Q volumetric throughput, m3 s  1 1284–1293.
Qf volumetric throughput of the feed stream, m3 s  1 Heath, A.R., Bahri, P.A., Fawell, P.D., Farrow, J.B., 2006b. Polymer flocculation of
calcite: population balance model. A.I.Ch.E. Journal 52, 1.
Qo volumetric throughput of the overflow stream, m3 s  1
Landman, K.A., White, L.R., Buscall, R., 1988. The continuous flow gravity
Qu volumetric throughput of the underflow stream, m3 s  1 thickener: steady state behaviour. A.I.Ch.E. Journal 34, 239–252.
R hindered settling function (resistance to dewatering), Landman, K.A., White, L.R., Eberl, M., 1994. Solid/liquid separation of flocculated
Pa s m  2 suspensions. Advances in Colloid and Interface Science 51, 175–246.
Lester, D., Usher, S.P., Scales, P.J., 2005. Estimation of the hindered settling function
t time, s R(f) from batch settling tests. A.I.Ch.E. Journal 51, 1158–1168.
u particle settling rate, m s  1 Shannon, P.T., Dehaas, R.D., Stroupe, E.P., Tory, E.M., 1964. Batch and continuous
U solids throughput in the thickener underflow, m3 s  1 thickening. Prediction of batch settling behavior from initial rate data with
results for rigid spheres. Industrial and Engineering Chemistry Fundamentals
z spatial ordinate, m 3, 250–260.
Shannon, P.T., Tory, E.M., 1966. Analysis of continuous thickening. Transactions of
the Society of Mining Engineers of the AIME 235, 375–382.
Greek letters Stickland, A.D., Usher, S.P., de Kretser, R.G., Scales, P.J., Hillis, P., Tillotson, M.R.,
2005. Validation of numerical filtration modelling for industrially relevant
a thickener shape factor, m2 m  2 materials. In: Proceedings of the Annual Exposition of the American Filtration
Society, Atlanta, USA.
Dr density difference between solid and liquid phases,
Tiller, C.L., Chen, W., 1988. Limiting operating conditions for continuous
kg m  3 thickeners. Chemical Engineering Science 43, 1695.
ff value of j in the thickener feed, m3 m  3 Tiller, F., Yeh, C.S., Tsai, C.D., Chen, W., 1987. Generalised approach to thickening,
fu value of j in the thickener underflow, m3 m  3 filtration and centrifugation. Filtration and Separation 24, 121–126.
Tory, E.M., Shannon, P.T., 1965. Reappraisal of the concept of settling in
j volume fraction of solids, m3 m  3 compression. Settling behavior and concentration profiles for initially
ARTICLE IN PRESS
3946 B.R. Gladman et al. / Chemical Engineering Science 65 (2010) 3937–3946

concentrated calcium carbonate slurries. Industrial and Engineering Chemistry Usher, S.P., Scales, P.J., 2005. Steady state thickener modelling from the
Fundamentals 4, 194–204. compressive yield stress and hindered settling function. Chemical Engineering
Usher, S.P., 2002. Suspension dewatering characterisation and optimisation. Ph.D. Journal 111, 253–261.
Thesis, Department of Chemical and Biomolecular Engineering, The University Usher, S.P., Scales, P.J., 2009. Predicting settler/clarifier behaviour: the role of shear
of Melbourne, Australia. effects. Transactions of the Filtration Society 9, 308–314.
Usher, S.P., Gladman, B., Scales, P.J., 2005. New techniques for understanding Usher, S.P., Spehar, R., Scales, P.J., 2009. Theoretical analysis of aggregate
settler/clarifier behaviour. In: American Filtration and Separations Society, densification; impact on thickener performance. Chemical Engineering Journal
18th Annual Conference, Atlanta, USA, April 10–13. 151, 202–208.

You might also like