You are on page 1of 22

Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2, pp.

220241 (2009)

COMPUTATIONAL FLUID DYNAMICS MODELLING


OF FLOCCULATION IN WATER TREATMENT: A REVIEW
J. Bridgeman*, B. Jefferson** and S. A. Parsons**

* School of Engineering (Civil Engineering), University of Birmingham, Edgbaston,


Birmingham, B15 2TT, UK
E-Mail: j.bridgeman@bham.ac.uk (Corresponding Author)
** Centre for Water Science, School of Applied Sciences, Cranfield University, Cranfield,
Bedfordshire, MK43 0AL, UK

ABSTRACT: The principal focus of this paper is to present a critical review of current approaches to modelling
the inter-related hydrodynamic, physical and chemical processes involved in the flocculation of water using
Computational Fluid Dynamics (CFD). The flows inside both laboratory and full scale mechanically-mixed
flocculators are complex and pose significant challenges to modellers. There exists a body of published work which
considers the bulk flow patterns, primarily at laboratory scale. However, there is little reported multiphase modelling
at either scale. Two-equation turbulence modelling has been found to produce variable results in comparison with
experimental data, due to the anisotropic nature of the swirling flow. However, the computational expense of
combining the sliding mesh treatment for a rotating mesh with the Reynolds Stress Model (RSM) in a full scale unit
is great, even when using a high performance computing facility. Future work should focus more on the multiphase
modelling aspects. Whilst opportunities exist for particle tracking using a Lagrangian model, few workers have
attempted this. The fractal nature of flocs poses limitations on the accuracy of the results generated and, in particular,
the impacts of density and porosity on drag force and settlement characteristics require additional work. There is
significant scope for the use of coupled population balance models and CFD to develop water treatment flocculation
models. Results from related work in the wastewater flocculation field are encouraging.
Keywords: computational fluid dynamics, turbulence, flocculation, mixing, multiphase modelling, fractal
dimension

1. INTRODUCTION between chemical and biological reactions and the


hydraulic conditions within which they occur.
The scope for applying CFD to flow problems in An insight into the relationships between
the water industry is large, with raw water hydrodynamics and water chemistry in particular
reservoirs, water treatment works, distribution can facilitate the optimisation of existing plant
systems (including treated water storage), and machinery and the development of novel
collection systems, and sewage treatment works design criteria for new unit processes. Such new,
all involving large scale fluid flow. Using CFD it optimised design criteria enable water utilities to
is possible to derive the information necessary to derive greater efficiencies by reducing capital
design, optimise or retrofit various treatment expenditure on new assets, reducing operational
processes. Further advantages include reduced expenditure on existing assets, and also reducing
lead-in times and costs for new designs, the chemical and energy input.
ability to examine the behaviour of systems at the The use of CFD, whilst traditionally strong and
limit or beyond design capacity, and the ability to widespread in the chemical, mechanical and
study large systems where controlled experiments manufacturing industries, has only begun to be
at full-scale would be difficult, if not impossible, exploited within the water industry relatively
to perform. recently. However, interest and experience in this
Effective water treatment to potable standards field are both growing apace, and CFD has now
involves a complex system of physical, chemical been successfully used in water and sewage
and biological processes, all occurring in a wide treatment and sewerage applications, both in the
range of hydrodynamic environments. Whilst UK and abroad. Examples of water treatment
much research effort has been devoted to applications are now many and varied. However,
chemical process optimisation, there remains the whilst CFD has proven to be of use in
need to understand fully the inter-relationships understanding the standard hydrodynamics of

Received: 8 May 2008; Revised: 11 Nov. 2008; Accepted: 12 Dec. 2008

220
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

flow processes in structures, there remain certain or inactivation of impurities, ranging in size from
areas which involve greater degrees of millimetres (grit, leaves) to microns (colloids,
computational complexity and where further viruses, protozoa). The principal processes
improvements are still required. One such involved in water treatment are identified in
example is flocculation. Table 1. Successful removal of these impurities
requires a range of different flow regimes
2. FLOW REGIMES AND SCENARIOS throughout a water treatment works (WTW). The
FOUND IN WATER TREATMENT accurate modelling of each process and flow
regime requires careful consideration, as each
2.1 Background presents its own subtleties and issues. Examples
Water treatment is based on a series of unit of the range of flow regimes and possible CFD
processes, each effecting some degree of removal modelling approaches are given in Table 2.

Table 1 Overview of principal water treatment processes.

Treatment Description Purpose


Process

Raw Water Bulk storage of water > 1 day. Backup supply to WTW in event of source pollution.
Storage Some solids removal via sedimentation.

Coagulation Chemical (trivalent inorganic coagulant) Destabilisation of water via neutralization of colloidal
dose and short ( <30 s) rapid mix. material charge and precipitation of soluble compounds.

Flocculation Slow, extended (1545 mins) mix. Encourage agglomeration of particles to form mass
fractal aggregates (flocs) up to 1000 m.

Clarification Sedimentation or flotation (via dissolved Solids removal.


air injection) of larger flocs.

Filtration Flow through porous granular media. Removal of smaller flocs and particles ( <100 m).

Disinfection Chemical (chlorine, UV) dose and Killing or inactivation of potentially harmfully micro-
storage (chlorine only). organisms.

Table 2 Flow regimes and CFD modelling approaches adopted in water treatment.

Flow Characteristics Treatment Process CFD Modelling Approach

Turbulent flow Open channel flow 2-equation turbulence models


Pipe flow Reynolds Stress Model
Mixing chambers Large Eddy Simulation
Direct Numerical Simulation
Laminar flow Settlement tanks Laminar flow model
Multiphase flow Coagulation Eulerian multiphase model
Flocculation Lagrangian particle model
Settlement
Flotation
Filtration
Disinfection
Rotating flow Mixing chambers Sliding Mesh
Flocculation Multiple Reference Frames

221
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

2.2 Turbulence modelling the Reynolds stress model (RSM). In addition,


most codes provide the Spalart-Allmaras model,
Most water treatment processes take place in
and the Large Eddy Simulation (LES) model. A
turbulent flow regimes. The most straightforward
descriptive comparison of the most commonly
complete turbulence models are the two-equation
used turbulence models is provided in Table 3 and
models where the turbulent velocity and length
the formulation of each is presented in
scales are determined via solution of two separate
Appendix 1.
transport equations; one for turbulent kinetic
energy, k, and one for the turbulence length scale
3. FLOCCULATION MODELLING
or some equivalent parameter (, the dissipation
of turbulence kinetic energy per unit time, or , 3.1 Background
the rate at which turbulent energy is dissipated).
These equations account for the production, The destabilisation (via coagulation) and
diffusion and destruction of turbulence within the subsequent agglomeration (via flocculation) of
flow field. The principal assumptions for these fine particles and colloids into larger particles is a
models are that the flow does not depart far from proven means of removing impurities (e.g.
local equilibrium, and that the Reynolds number turbidity and colour) at WTWs. Chemical
is high enough to ensure that local isotropy is coagulant addition brings about a change in the
satisfied. nature of small particles, reducing their negative
Two-equation turbulence models are based on the surface charge and rendering them unstable,
eddy viscosity concept of Boussinesq. This eddy- whilst flocculation encourages particle
viscosity model states that the Reynolds stresses agglomeration via gentle mixing and the
are related to the local shear via eddy viscosity, t , formation of irregularly-shaped, loosely
in the following manner (Rodi, 1993): connected mass fractal aggregates, known as flocs.
The size and structure of flocs are fundamental to
U U j 2
ui u j = t i + k ij = 2 t s ij 2 k ij the efficient operation of WTWs. Ineffective
x xi 3 3 coagulation and flocculation result in poorer
j
(1) quality feed water to downstream treatment
processes, potentially jeopardising treated water
The turbulent eddy viscosity, t , is defined in quality and increasing operational costs. Several
terms of the turbulent kinetic energy, k, and the interrelated criteria govern the efficiency of the
rate of dissipation of turbulent energy per unit coagulation and flocculation stages; viz.
mass, , according to: coagulant type and dosage, pH and mixing
k2 arrangements.
k- model: t = C (2) Flocculation is the transformation of smaller

destabilised particles into larger aggregates or
k- model: k flocs which are subsequently removed via
t = C (3)
sedimentation or flotation processes. For all flocs
in their initial growth phase, the floc formation
ui u j U i U i U j process is understood to be a balance between the
k= and = + (4)
2 x j x xi rate of collision-induced aggregation and the rate
j
of breakage for given shear conditions (Bouyer
ij represents the Kronecker delta and is defined et al., 2005; Mikkelsen, 2001; Biggs and Lant,
as: 2000).
1 i= j R floc = ij Rcol Rbr (6)
ij = (5) ij
0 i j
where Rfloc is the overall rate of floc growth, is
A detailed discussion of the terms in the k- the collision efficiency factor (0 < < 1), R col is
models can be found in Launder and Spalding the rate of particle collision and Rbr represents the
(1974). The principal turbulence models which rate of floc breakage.
are generally packaged with commercially
available CFD packages are the standard k-
model, the low Reynolds number k- model (LRN
k-), the renormalized group k- model (RNG k-),
the realizable k- model, the standard k- model,
the shear-stress transport (SST) k- model, and

222
Table 3 Comparison of available turbulence models (from Versteeg and Malalasekera, 1995, Marshall and Bakker, 2004, and Menter, 2003).

Model Comments Advantages Disadvantages

Standard k- Semi-empirical modelling of k and . Simplest model. Poor performance in some scenarios (i.e.
Valid for full turbulence only (molecular Excellent performance for many flows. rotating flows, flow separation, adverse
diffusion ignored). Well established. pressure gradients).
Assumes isotropy in turbulence.
Poor prediction of lateral expansion in 3-d
wall jets.

Renormalized Based on statistical methods, not observed Improved performance for swirling flow Less stable than the standard k- model.
Group (RNG) k- fluid behaviour. compared to the standard k- model.
Mathematics is highly abstruse. Texts only
quote model equations which result from it.

223
Effects of small-scale turbulence
represented by means of a random forcing
function in N-S equations.
Procedure systematically removes small
scales of motion by expressing their effects
in terms of larger scale motions and a
modified viscosity.
Similar in form to sk-, but modified
dissipation equation to describe high-strain
flows better.
Differential equation solved for t (changes
C from 0.09 to 0.0845 at high Re), so good
for transitional flows.
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)
Table 3 (Continued)

Model Comments Advantages Disadvantages

Realizable k- Recent development. In highly strained Suited to round jets, as well as swirling Not recommended for use with multiple
flows, the normal Re stresses, u i2 , become flows and flows where separation occurs. reference frames.
negative (unrealizable condition), so t
uses variable C .
C is function of local strain rate and fluid
rotation.
Different source and sink terms in transport
equations for eddy dissipation.
Good for spreading rate of round jets.

Standard k- = /k = specific dissipation rate. Valid throughout the boundary layer, Separation is typically predicted to be

224
sk- solves for dissipation of turbulent subject to sufficiently fine grid resolution. excessive and early.
kinetic energy, k- solves for rate at which
dissipation occurs.
Resolves near wall without wall functions,
so can be applied through boundary layer.

Shear Stress As k- except for gradual change from k- Suitable for adverse pressure gradients and Less suitable for free shear flows.
Transport (SST) k- in inner region of boundary layer to high pressure-induced separation.
Re version of sk- in outer part. Accounts for the transport of the principal
Modified t formulation to account for turbulent shear stress.
transport effects of principal turbulent shear
stresses.
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

Reynolds Stress The most general of all models. Accurate calculation of mean flow Computationally expensive.
Model properties and all Reynolds stresses. Not always more accurate than two-equation
Yields superior results to k- models for models.
flows with stagnation points. Harder to obtain converged result.
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

More accurately, the ij R col ij term should be specific and interesting challenges for the
modeller.
broken down further into its constituent parts
Flocculation at WTWs is effected either via
where ij R col ij is a function of
mechanical or hydraulic means. For mechanical
flocculation, the energy input is via an agitator
BM
ij BM Rcolij , Sh ij Sh Rcolij , and DS ij DS Rcolij (7)
which generates the necessary shear stress,
whereas in hydraulic flocculation, energy is
where i and j refer to discrete particles, and BM, imparted via the headloss across a baffled,
Sh and DS refer to the collision mechanisms of serpentine channel. The two processes are shown
Brownian Motion, shear and differential schematically at Fig. 1.
settlement respectively. (Derivations of these Coagulation and flocculation optimisation are
mechanisms are to be found in the literature, e.g. generally considered at the laboratory scale, using
Kusters, Wijers and Thoenes, 1997). As a result, a jar test apparatus and procedure. This well-
flocs do not continue to grow throughout the established process optimisation technique allows
flocculation stage, but rather, they attain a a rapid assessment of key variables (e.g.
limiting size beyond which breakage prevents coagulant dose, pH, mixing speed and
further overall growth. The flocculation process, flocculation time). The test apparatus typically
and hence this limiting size, is governed by comprises four or six glass vessels, each with a
physico-chemical conditions and by the shear powered paddle to stir the contents of the vessel
conditions within the containing vessel. Bouyer and each containing the same volume of raw
et al. (2005) investigated the link between water. The paddles are set to rotate at high speed
hydrodynamics, physico-chemical conditions and for a short period, during which time different
floc size, and found a dependency of floc size on quantities of coagulant (and acid or alkali for pH
hydrodynamic history. Consequently, vessel adjustment as necessary) are added to each vessel
hydrodynamics exhibit significant control over and mixed. After a short period of intense mixing
the effectiveness of flocculation. (3060 seconds) to simulate the coagulation
Particle removal efficiency decreases with process, the paddle speed is reduced to produce a
decreasing particle size (Boller and Blaser, 1998). more gentle mixing to simulate the flocculation
Therefore, flocs must be able to withstand shear process. After 20 minutes of flocculation, mixing
energy applied to them in various different unit is terminated, the paddles removed and the
processes; otherwise, when the degree of shear suspensions allowed to settle. The flocs and
exceeds a threshold value, floc breakage will treated water in each vessel are then analysed for
occur. However, quantification of the energy floc size, turbidity removal and organics removal,
requirements for floc breakage is not allowing conclusions to be drawn regarding
straightforward, and despite much work in this optimum coagulant dose and pH. These
field, no standard strength test exists. Thus, conclusions are often then applied to the
effective and efficient modelling of the various operation of the main treatment plant. Two typical
types and configurations of flocculators presents jar test configurations are shown at Fig. 2.
some excellent opportunities, but also some

Fig. 1a Schematic diagram of sectional view of Fig. 1b Schematic diagram of plan view of an hydraulic
a mechanical flocculator. flocculator, from Haarhoff and Van der Walt (2001).

225
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

6
6

25

150
127
25

76

53
30

76

100 115

Section A-A Section B-B

6 6
A A B B

115
Plan Plan

Fig. 2 Typical jar test vessel configurations (not to scale, all dimensions in millimetres).

by explicit calculation of the flow pattern in the


vicinity of the blades without recourse to
experimental data. This is to be favoured above
the application of experimental data, since it
avoids the need to extrapolate any such
experimental data in order to apply it to
situations for which no experiments have been,
or can be, performed.
One method of explicitly calculating the flow
Fig. 3 Sliding mesh in two orientations, shown in field in a rotating flow scenario is the sliding
2D, from Marshall and Bakker (2004). mesh method. With the sliding mesh method, the
tank is divided into two regions that are treated
3.2 Rotating flow separately: 1) the impeller region and 2) the tank
region which includes the bulk of the liquid, the
The modelling of rotating flows (such as those tank wall, the tank bottom and the baffles (Fig. 3).
found in mechanical flocculation) is of great The grid in the impeller region rotates with the
importance in mechanical flocculation and a impeller, whilst the grid in the tank remains
number of approaches are available to the stationary. The two grids slide past each other at
modeller to represent this scenario. The a cylindrical interface. The sliding grid model
traditional approach to modelling paddle mixers explicitly calculates the mixer region, and then
was to apply experimentally-obtained velocity rotates this section of the grid relative to the rest
data in the outflow of the impeller. However, of the domain. It is assumed that the flow field is
such methods have now largely been superseded unsteady, and the interactions are modelled as

226
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

they occur. Consequently, the sliding mesh 3.3 Swirling flow


model is the preferred option in instances where
The mechanical mixing of flocculators can often
interaction between rotor and baffle is strong and
result in swirling flow conditions, thus
the most accurate simulation of the system is
introducing anisotropic turbulence conditions,
desired. Whilst the sliding mesh model is the
and so presenting some specific and interesting
most accurate method for simulating flows in
challenges. Although two-equation models are
multiple moving reference frames, it is also the
commonly used by CFD practitioners, these
most computationally demanding. Using the
models exhibit an inherent difficulty when
sliding mesh model to analyse laminar flows in a
modelling swirling flow as they are predicated on
stirred reactor, Bakker et al. (2000) used a time
the assumption that turbulence is isotropic,
step of 0.01 s and 1000 time steps to study the
meaning that the turbulent stress tensor is
flow created by a pitched blade turbine in a tank.
independent of direction (or, more precisely,
With an impeller rotational speed of 3.75 s-1,
invariant with respect to rotation and reflection
37.5 revolutions were simulated. Bakker et al.
of the coordinate axes of the coordinate system
reported that the calculation time was
moving with the mean motion of the fluid). This
approximately 15 minutes per impeller
assumption is not true for swirling flow which is
revolution on a Cray C-90 computer, giving a
anisotropic in nature and where turbulence is
total CPU usage time of 9.75 hours.
generally more in the tangential and axial
An alternative means of modelling the mixing
directions compared to the radial direction.
process is to use the Multiple Reference Frames
Therefore, in order to capture this anisotropic
(MRF) approach (Luo, Issa and Gosman, 1994).
nature of turbulence, other models (e.g. RSM)
This approach adopts two reference frames; one
should be employed as the effects of strong
stationary frame related to the vessel walls; the
turbulence anisotropy can be modelled rigorously
other related to the rotating shaft and impeller.
only by the second-moment closure adopted
The fluid zone is divided into two separate
therein. It is widely acknowledged that the RSM
regions, one of which is related to the stationary
is the most rigorous of the available models
zone, whilst the other, close to the rotating
(Versteeg and Malalasekera, 1995); however,
impeller, is related to the moving zone (Fig. 4).
this additional rigour comes with concomitant
The momentum equations and closure models
additional computational power and time.
are resolved in the separate zones and a steady-
Irrespective of the above comments, it is
state approximation is made at the zone interface.
interesting to note that the literature contains
The mesh does not move in this technique. The
several examples of the use of two-equation
MRF is computationally far more expedient than
models to solve highly swirling flow conditions
the sliding mesh method and for steady state
(e.g. Ducoste and Clark, 1999; Korpijarvi et al.,
applications has been used successfully to model
1999; Korpijarvi, Laine and Ahlstedt, 2000;
mixing systems of various types. However, if an
Essemiani and de Traversay, 2002a & b; Ng,
unsteady solution is required, the sliding mesh
Borrett and Yianneskis, 1999). Korpijarvi et al.
approach is the method of choice.
(1999) modelled the flow patterns found within a
cylindrical jar test device. Grids of 30,000 and
67,000 cells were developed to represent the
vessel and the authors tested the sensitivity of the
flow patterns within the two grids to four
different turbulence models; viz. standard k-,
RNG k-, LRN k- and the RSM, used in
conjunction with the Sliding Mesh approach.
Velocities within the mini-flocculator were found
to be insensitive to changes in turbulence model
or grid density. Korpijarvi et al. demonstrated
that the total dissipated power in the fine grid
mesh was more than twice that found in the
coarser mesh. The difference observed by the
authors when comparing total dissipated power
Fig. 4 Cylindrical mixing tank with an MRF associated with LRN k- and RSM models using
boundary surrounding the impeller, from the coarse grid rose to in excess of ten-fold,
Fluent (2005).

227
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

demonstrating the importance of turbulence if the volume fraction of water in any given cell
model selection. is represented by w , then,
All turbulence models offer advantages and
w = 0 implies that the cell is empty of water,
disadvantages over others in certain applications,
and no single turbulence model can be judged to
w = 1 implies that the cell is full of water, and
be the optimum model in all circumstances. 0 < w < 1 implies that the cell contains the
Consequently, it is necessary to evaluate each air/water interface.
circumstance individually and form a view as to Haque, Mahmud and Roberts (2006) simulated
which model would provide the best fit for the the flow field in unbaffled vessels mixed with a
particular flow under examination. Factors to paddle impeller and a Rushton turbine and
consider include, inter alia, the physics adopted the VOF approach to represent the
encompassed in the flow, the established practice vortex formed at the surface. Using the MRF
for a specific class of problem, the level of approach to rotating flow and the RSM to model
accuracy required, the available computational turbulence, Haque, Mahmud and Roberts (2006)
resources, and the amount of time available for reported a typical run time of 70 hours to achieve
the simulation. It is clear from the above that the convergence (Sun UltraSPARC III processor,
selection of turbulence model is not a 4Gb RAM, 900 MHz clock speed). Thus, it can
straightforward one. Menter (2003) correctly be seen that the computational expense of
suggested that one should resist the temptation to considering the liquid surface shape is significant.
claim that one model is superior to another, as Whilst the authors reported reasonable
the robustness of a turbulence model is correlation between observed and numerically-
undoubtedly a function of the code being used obtained vortex shapes, they found discrepancies
and the scenario being modelled. This has been between the predicted and measured values in the
reaffirmed by several authors (e.g. Freitas, 1995; vicinity of the shaft. Haque, Mahmud and
Iccarino, 2000) who urged caution in the Roberts (2006) also compared axial velocities at
unilateral promotion of one code, turbulence various locations, but did not consider turbulence
model, convergence criteria or discretisation quantities. Despite the reasonable results
technique at the expense of all others. Specific obtained by Haque, Mahmud and Roberts, it is
factors relating to the scenario being analysed noteworthy that Marshall and Bakker (2004)
must always be accounted for. recommended that this approach should not be
The situation is further complicated by the fact adopted for the prediction of vortex shape.
that in a jar tester, the highly swirling, Marshall and Bakker recognized that the VOF
anisotropic nature of the flow often leads to the model has application in the monitoring of liquid
formation of a surface vortex. Using CFD it is surface shape, but suggested that the model is ill-
possible to track the shape of the liquid surface equipped to deal with the bubble formation
during mixing using the Volume of Fluid (VOF) arising from the breaking of the air/water
Model. In this approach, the computational interface when air passes through grid cells
domain is extended beyond the surface to span where large momentum sources exist. As a result,
the air/water interface, and the volume fraction Marshall and Bakker (2004) concluded that it is
of each phase is determined throughout the flow appropriate to use the VOF model to indicate
field by solving the continuity equation for one whether or not vortexing will occur, but that it
of the phases. Although most often applied to an should not be used to predict the flow condition
air/water interface in water treatment afterwards.
applications, the VOF model can be applied to
any number of immiscible fluids, subject to 3.4 Velocity gradient
computational power. For an homogeneous Traditionally, flocculators have been
multiphase system consisting of air and water, characterised on the basis of the velocity gradient
the equation is solved for the volume fraction of (Camp and Stein, 1943). This parameter is used
the liquid phase. In a given computational cell, worldwide to characterize mixing in a wide range
the volume fraction of unity represents pure of environmental engineering applications, and
water and zero represents pure air. Thus, the principally mixing in flocculation basins (e.g.
volume fraction of the air is obtained as the Crittenden et al., 2005). From the consideration
difference between the liquid phase volume of the angular distortion of an elemental volume
fraction and unity. The air/water interface is of water arising from the application of
determined by identifying the cells where the tangential surface forces, G is defined as the root
volume fraction is between zero and unity. Thus, mean square velocity gradient in a mixing vessel:

228
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

u v 2 u w 2 v w 2
0.5
a single number (Cleasby, 1984; Clark, 1985;
G = + + + + + (8) Graber, 1994; Luo, 1997; Jones, 1999).
y x z x z y Furthermore, the distribution of velocity
gradients within a stirred tank is clearly not
where u, v and w are the velocity components in
uniform, and local power consumption at a point
the x, y and z directions of a Cartesian coordinate
of high turbulence within a vessel (e.g. adjacent
system.
to the impeller) can be several orders of
Camp and Stein termed this the absolute velocity
magnitude in excess of the rest of the vessel (Luo,
gradient and related this to work done per unit
1997). When considering flocculation, this is
volume per unit time via:
unfortunate as it is precisely the magnitude and
=G
2
(9) fluctuations in local shear to which a floc is
subjected, and not the average value, which
From which: determine the success of flocculation.
However, using CFD it is possible to quantify
P /V (10) and understand the local impact of mean flow
G= =
and turbulence on floc formation and break-up
using the local velocity gradient, GL , where
where work of shear per unit volume per
unit time at a point
GL = (13)
P is power dissipated
V is tank volume
Korpijarvi, Laine and Ahlstedt (2000)
is the dynamic viscosity of the water,
demonstrated the very large variability of the
is the energy dissipation rate per unit
local velocity gradient value (GL ) within a jar
mass
test vessel, thus questioning the validity of the
is the kinematic viscosity of the
water use of the G value to characterise the vessel.
Unsurprisingly, the largest GL values were to be
In theory, the absolute velocity gradient can be found near the paddles and particularly in the
calculated at any point within a mixing vessel, turbulent area behind the panel, stretching in a
provided that the power dissipated is known at radial direction towards the wall.
that point. In practice, however, the flow
characteristics vary within the mixing vessel
from point to point, and so too does the energy
dissipation. Consequently the velocity gradient is
a function of both time and position. Given the
difficulties associated with calculation of G,
workers have traditionally replaced the absolute
velocity gradient with an approximation of the
exact value; that is its average value throughout
the vessel, G :
Pave
G = (11)
V

where the average power consumption, Pave , is


readily obtained via
Pave = Po N 3 D 5 (12)
Fig. 5 Distribution of local velocity gradients in a
where Po is the impeller power number pilot scale mechanical flocculator.
is the fluid density Flow rate = 2 m3.hr -1, mixing speed = 73 rpm,
N is the rotational speed of the impeller from Essemiani and de Traversay (2002a).
D is the impeller diameter
Essemiani and de Traversay (2002a) modelled a
Since its introduction, authors have argued that pilot scale mechanical flocculator to determine
the concept of the G value is flawed, as it areas of highest shear stress (where floc break-up
attempts to represent a complex flow field within is to be anticipated) and also areas of the lowest

229
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

values where, subject to a de minimus, one absolute terms and normalized by dividing by the
would expect coalescence (i.e. floc growth) to G value. They then introduced the concept of a
occur. Similar to Korpijarvi et al. (1999) and 95th percentile of the normalized G value
Korpijarvi, Laine and Ahlstedt (2000), the distribution in the flocculator as an arbitrary (but,
authors found that the distribution of local in their opinion, realistic) performance indicator
velocity gradients indicated high GL intensity for flocculator optimisation (i.e. a G value which
adjacent to the impeller, decreasing with distance is exceeded only in 5% of the total flocculator
from it (Fig. 5). volume). The authors argued that use of this G95
Essemiani and de Traversay (2002a) found that figure enables the designer to compare absolute
for the system considered, G and the average GL and normalized G values on a more
values exhibited reasonably close correlation, representative basis than by simply using the
although there was a great range of GL values; a average. Using this approach, Haarhoff and van
conclusion corroborated by Bridgeman, Jefferson der Walt were able to conclude that the slot
and Parsons (2008). Essemiani and de Traversay width ratio is the single most important
concluded that G does not account for variables geometric ratio for design purposes. The effect of
such as impeller position, geometry, operating the overlap ratio was found to have a less
mode and residence time. However, these pronounced impact on the G values, whilst the
parameters do affect the velocity gradient effect of the depth ratio was inconsequential if
distribution, and therefore, spatial floc size the channel velocity was maintained.
distribution. As a result, Essemiani and de
Traversay suggested that for the same G value, 3.5 Floc breakage
changing mixing rates and impeller type will Having expended time and energy in developing
affect performance. flocs, it is important that operating conditions do
Craig et al. (2002) developed a model using not subsequently cause their break-up. As a
Fluent to simulate the operation and performance result, floc strength, growth and breakage have
of a simple flocculation tank equipped with an been the subject of detailed research in recent
axial impeller generating turbulent flow years.
conditions. No details of modelling technique or Floc size may be considered to be a balance
turbulence model used were provided. Craig et al. between the hydrodynamic forces exerted on a
used the model to demonstrate variations in the floc and the strength of the floc. Where the floc
GL value, and to identify maximum and strength is resistant to the hydrodynamic forces,
minimum GL values associated with different one would expect floc size either to remain
operational scenarios. The authors identified that constant or for growth to occur. Where the
use of GL values in this manner enabled the hydrodynamic forces exceed floc strength,
identification within the flocculation tank of breakage will occur. Consequently, the
those areas where both coalescence and floc conceptual growth/breakage mechanism may be
break-up would occur. Similar to Essemiani and expressed as follows:
de Traversay (2002a), this led the authors to
Hydrodynamic Forces F
question the validity of the use of the G value as B= = (14)
Floc Strength J
a design parameter.
Haarhoff and van der Walt (2001) used the where F represents the hydrodynamic forces
lesser-used Flo++ code to study flow in an exerted by the flow, and J represents the strength
around-the-end hydraulic flocculator. Building of the floc (Coufort, Bouyer and Lin, 2005). It
on previous work (van der Walt, 1998), Haarhoff is clear from Eq. (14) that breakage will occur
and van der Walt used CFD to develop when B > 1, and floc size will be maintained or
qualitative guidelines for three fundamental increased when B < 1. Floc strength, J, is a
geometric ratios (slot width ratio, overlap ratio function of the physico-chemical conditions (raw
and depth ratio) (see Fig. 1b). No details of water type, coagulant type and dose) and the floc
modelling strategy were provided in the paper. structure.
Haarhoff and van der Walt also studied the Yeung, Gibbs and Pelton (1997) suggested that
variation of the G value in hydraulic flocculators. the hydrodynamic force required to pull apart a
It is known that hydraulic flocculators incur floc in tensile mode may be expressed as:
zones of high turbulence at the baffle edges
which can cause floc break-up. The authors F . .d 2 (15)
4
showed variations in GL in the flocculator in both

230
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

where F represents the floc rupture force, performance of existing flocculators or to design
represents the floc strength, and d is the floc new installations, thus permitting optimisation on
diameter. In the viscous sub-range, = . . the basis of implied floc strength. Indeed,
Substituting into Eq. (14) shows that the Bridgeman, Jefferson and Parsons (2007)
breakage mechanism in the viscous sub-range, modelled two full scale flocculators (one
BVSR , may be expressed as: mechanically mixed, the other an hydraulic
installation) and assessed their performance in
C1 .d 2 terms of the breakage threshold concept.
BVSR = (16)
J However, the computational expense of using
sliding mesh to model a three dimensional full
where C1 is a constant. scale flocculator using the realizable k- model
Thomas (1964) suggested that in the inertial sub- was significant. (Using a high performance
range computing facility, the full scale mechanical
2 2 2 flocculator sliding mesh simulations required up
= .u and u = C2 ( .d ) 3
(17) & (18) to 12 days CPU usage).
Substituting into Eq. (14) shows that the
breakage mechanism in the inertial sub-range, 3.6 Residence time distribution
BISR , may be expressed as: One means by which flocculator performance
2 8 can be assessed is by considering the residence
C2 3 .d 3
(19) time of flocs. Empirical guidelines exist
BISR =
J regarding optimum residence times for flocs, and
where C2 is a constant. CFD can be used to gain an improved
Consequently, it is apparent from Eqs. (16) and understanding of retention times in vessels.
(19) that floc size is dependent on the turbulence Although it is possible to employ the discrete
energy dissipation rate and floc strength, phase model to undertake a Lagrangian particle
irrespective of sub-range. This is clearly of great tracking analysis on a number of particles to
significance when one attempts to gain an determine overall residence times, this method
understanding of floc breakage mechanisms and requires a large number of particles to be
limiting floc size as it directs workers to focus analysed in order for the results to be statistically
their towards and J. valid and so introduces significant additional
Essemiani and de Traversay (2002a & b) argued computational expense and time. An alternative
that the GL distribution (and, hence, distribution) approach is to treat a tracer fluid as a continuum
in a vessel is significant as it controls particle by solving a transport equation for the tracer
suspension, distribution, coalescence and break- species. The transport and decay of a chemical
up efficiency. The authors did not, however, species in a flow can be described by a general
attempt to quantify the effects of local shear on convection/diffusion equation:
floc characteristics; a subject addressed by 2
Bridgeman, Jefferson and Parsons (2008) who + u i D + t 2 = S (20)
t xi xi
used CFD coupled with a Lagrangian particle
trajectory model to model the flow field within a where is any scalar variable, D is the
standard jar test apparatus to study the effects of
molecular diffusivity of the scalar, is the
turbulence on individual flocs. Combining
turbulent Schmidt number of the scalar, and S is
numerical and experimental data, the authors
the source/sink term for the scalar variable
were able to postulate velocity gradient values at
(= 0 for tracer transport).
which floc breakage occurs for three different
The residence time distribution (RTD), E(t) is a
floc suspensions. Although the threshold values
measure of the bulk flow patterns in a vessel and
were determined using jar test and CFD data in
is defined such that the RTD function represents
combination, they were based on the flocs
the fraction of fluid at the outlet that has a
resistance to induced velocity gradients. This is a
residence time in the vessel of between t and
significant result, as previous breakage
t + t. Consequently,
thresholds had always been expressed in terms of
t1
mixing speed and so could not be applied at full
scale. The results of Bridgeman, Jefferson and E (t )dt = 1 and E (t )dt = 1 E (t )dt
0 t1 0
(21)
Parsons (2008) can be adopted for use in other
situations and can be used to assess the

231
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

To assess RTD numerically, a small amount of 3.7 Multiphase modelling


non-reactive tracer (approximately 0.5% of
vessel volume), with fluid properties set to the 3.7.1 Background
same as water, is injected into the flow via the Much of the CFD modelling of flocculators has
inlet. All reactions between the species and bulk considered the bulk flow characteristics of
flow are turned off, and the species concentration specific vessels. Examples of CFD flocculation
monitored at the outlet. It is normal for the tracer modelling are to be found in Table 5. It is
study data to be normalised on the basis of t and possible, however, to use CFD to consider floc
CN where transport analysis in flocculators. Using a
Lagrangian approach, the predicted pathlines of
t=
0
Ct dt

Ct t (22) particles, varying in both size and density, can be

C t studied.
0
C dt

3.7.2 Lagrangian particle model


CN =
t
Cd =
0
Cdt

Ct
(23) Treating each particle individually, the
0
t t t
Lagrangian particle model solves the equations
C of motion for each particle to obtain its trajectory
E ( ) = (24) in space and time. Equations for temperature and
CN
species concentration can be added to model heat
where t = mean residence time in vessel, and mass transfers between the particles and
C = concentration exiting vessel at time the surrounding fluid, thus yielding detailed
t, information regarding the particles, their
t = time since addition of tracer pulse positions (in both space and time), trajectories,
to vessel. temperature and species concentration.

Crittenden et al. (2005) presented the key terms


used to characterise RTD curves and these are
presented in Table 4.

Table 4 RTD curve characterisation terms, from Crittenden et al. (2005).

Term Definition

t Theoretical hydraulic retention time, = V/Q

ti Time at which tracer first appears

Normalised retention time = t


t

tp Time at which peak concentration of tracer is observed

t Mean residence time = centroid of E() curve

t10 , t50 , t90 Time at which 10, 50 and 90% of tracer has passed through reactor

t50 / t90 Morrill dispersion index

ti / Index of short-circuiting. Ideal plug flow reactor = 1. Tends to 0 with increased short-
circuiting

t50 / Index of mean retention time. Measure of skew of E() curve

232
Table 5 CFD modelling of flocculation processes.

Author Code Aim Scale Dim Turb. Extra Validation


Model Model method

Ducoste & Clark, 1999 FIDAP Code evaluation for Lab 3D sk- None LDV
modelling flocculator fluid
mechanics.
Bridgeman, Jefferson & Parsons, 2008 Fluent Evaluation of floc strength Lab 3D sk-, Rz k-, RNG DPM LDA
and trajectory in jar tester. k-, k-, SST k-,
RSM
Bridgeman, Jefferson & Parsons, 2007 Fluent Flow characteristics in two Lab 3D sk-, Rz k-, RNG DPM LDA, PIV
jar test configurations. k-, k-, SST k-,
RSM
Flow characteristics in

233
mechanical flocculator. Full sk-, Rz k-, RSM DPM, RTD, Power
RTD measurements

Hydraulic flocculator. Full sk- RTD Turbine meter


velocity
measurements
Craig et al., 2002 Fluent Flow characteristics. Pilot ND ND ND ND
Essemiani & de Traversay, 2002a Fluent Mixing efficiency. Pilot 3D sk- None RTD, Power No.,
Pumping No.
Essemiani & de Traversay, 2002b Fluent Comparison of mixer Pilot 3D sk- None RTD, Power No.,
configurations. Pumping No.
Haarhoff & van der Walt, 2001 Flo+ + Hydraulic flocculator design Full 3D sk- RTD Turbine meter
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

optimization. velocity
measurements
Korpijarvi, Laine & Ahlstedt, 2000 CFX GL variations in jar tester. Lab 3D sk-, RNG k-, None LDA, PIV
RSM
Table 5 (Continued)

Author Code Aim Scale Dim Turb. Extra Validation


Model Model method

Korpijarvi et al., 1999 CFX Turbulence modelling and Lab 3D sk-, RNG k-, None LDA, PIV
mesh density analysis jar LRN k-, RSM
tester.
Lain et al., 1999 Fluent Flow characteristics to Full 2D RNG k- DPM None
determine cause of
underperformance.
Nopens, 2007 Fluent Coupling CFD and PBM in Full 2D ske PBM None
wastewater clarifier. (QMOM)
Prat & Ducoste, 2007 Phoenics Evaluation of Lagrangian Lab 3D ske Lagrangian, Experimental data
and Eulerian approaches Eulerian,

234
for simulating flocculation PBM
in stirred vessels.

N/A not applicable, ND not discussed, ADV acoustic Doppler velocimetry, LDV laser Doppler velocimetry, LDA laser Doppler anemometry,
PIV particle image velocimetry, RTD residence time distribution, MFM multi-fluid model.
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

Most commercially-available CFD software has mass fractal aggregatesi.e. they demonstrate
an in-built discrete phase model which can be self-similarity irrespective of scale and
used to model the trajectories of individual demonstrate a power law relationship between
particles within a flow via integration of the force mass (or volume) and length, such that:
balance on a particle in a Lagrangian reference df
frame. Equating particle inertia with the forces ML (31)
acting on it, and applying a Cartesian co-ordinate where df is a non-integer mass fractal dimension.
system, yields: A low fractal dimension (< 2.0) indicates an open
du p g x ( p ) structure, whereas a higher fractal dimension
= FD (u u p ) + + Fx (25) indicates a more compact structure. (For regular,
dt
three-dimensional objects, df = 3.0).
where, u is the fluid velocity, up is the particle This fractal nature has some very important
velocity, is the dynamic viscosity of the fluid, consequences on floc settlement and transport and
and and p are the fluid and particle densities, the modelling thereof. For example, as floc size
respectively, FD (u u p ) is the drag force per unit increases, the density decreases (Tambo and
Watanabe, 1979), with the floc effective density
particle mass, and
represented by
18 CD Re
FD = . (26) E = B . a y (32)
p d p 2 24
where a is floc size and B and y are constants.
where dp is the particle diameter, and Re is the The effective density is proportional to the
relative Reynolds number, defined as: volume fraction of solid in the floc, such that
d p u p u E = s ( p ) (33)
Re = (27)

and hence y is related to df via
The drag coefficient, CD , is defined as:
df = 3 y (34)
2 3
C D = 1 + + (28)
Re Re 2 Furthermore, because flocs are not spherical, the
drag coefficient expression must be amended.
where n are constants that apply to smooth Assuming the sphericity of flocs to be
spherical particles over several ranges of Re given approximately 0.8, at low Reynolds numbers,
by Morsi and Alexander (1972) ( n values are Tambo and Watanabe (1979) suggested that
shown in Appendix 2). CD 45 / Re (35)
The term Fx in Eq. (25) relates to additional forces
in the particle force balance that are relevant only i.e. almost twice the value predicted via Stokes
under special circumstances; for example, forces law (= 24/Re).
on particles that arise due to the rotation of the Whilst this deviation from sphericity may be
reference frame. Considering rotation about the z easily addressed via a simple adjustment to
axis, the forces on the particles in the x and y Eq. (28), the porosity of flocs (and hence the
directions may be expressed as: possibility of flow through the floc itself) poses
additional modelling challenges. The net effect of

1 2 x + 2 u y , p u y (29) porosity is that a floc will experience reduced
p drag compared to an impermeable sphere of the
p
same size and density. This effect becomes
where uy , p and uy are the particle and fluid significant for more open flocs with low fractal
velocities in the y direction, and dimensions (df < 2.0), settling faster than solids of
the same size and density.

1 2 y + 2 u x , p u x (30) However, the Lagrangian particle models treat all
p
p particles as point masses, and therefore cannot
consider the effect of porosity directly. One
where ux , p and ux are the particle and fluid
means of addressing this is to assume that the
velocities in the x direction.
porosity has an effect on the drag forces and
However, the work of Morsi and Alexander (1972)
consequently amend the drag coefficient
was predicated on the assumption that particles
accordingly. However, the means by which the
are spherical, whereas flocs are irregularly-shaped,
information required to do this with any degree of

235
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

accuracy is unclear. This also involves further use of the QMOM. The QMOM offers reduced
amendment of a coefficient to which adjustment computational expense compared to the discrete
has already been made to accommodate deviation model, but it differs from the SMM by replacing
from sphericity. This clearly adds an additional the exact closure with an approximate closure,
degree of uncertainty into the modelling process. requiring only a small number of scalar equations
to track population moments (McGraw, 1997).
3.8 Population balance modelling (PBM) The reduced computational cost of the moment
Referring back to the modelling work undertaken methods means that they can be combined with
on water treatment flocculation, it can be seen that CFD calculations such that the effects of local
much work in the water treatment field has (rather than global average) flow field
addressed flow patterns. However, it is not just characteristics can be considered.
the hydrodynamic behaviour of flocculators Whilst the development of PBM techniques has
which is of interest, but the agglomeration, advanced, the coupling of PBMs with CFD
growth and breakage processes are of great remains in its infancy, particularly in water
significance also. Multiphase models which treatment applications with most (but not all)
incorporate a particle size distribution (PSD), work being undertaken using wastewater. Prat
such as those found in a flocculator, require a and Ducoste (2007) considered the evolution of
population balance model (PBM) to describe flocs using the QMOM and used CFD to solve the
particle population changes. Several solution turbulent flow field within a 28 litre stirred vessel
methods exist, including, inter alia, the discrete, with clay particles. Whilst the results provide
class size method (Hounslow, Ryall and Marshall, proof of concept for the decoupling of fluid
1998), the standard method of moments (SMM) flow and flocculation dynamics in the
(Randolph and Larson, 1971), and the quadrature Lagrangian/QMOM approach, they do not
method of moments (QMOM) (Marchisio, Vigil provide information regarding the growth and
and Fox, 2003). A full description of each is strength of flocs in specific water treatment
outside the scope of this paper and only brief applications. Other recent work includes that of
comments on each are made. Nopens (2007) (CFD and QMOM-based PBM in
The discrete method requires the particle wastewater clarifier) and Feng and Li (2008)
population to be discretised into a finite number (theoretical PBM approach, encapsulating internal
of size bins and is based on the representation of body forces and fluid shear stress). A detailed
the PSD in terms of those bins. This method has review of all recent PBM work is outside the
the advantage that the PSD is calculated directly; scope of this paper; the interested reader is
however, the bins must be defined from the outset referred to Hounslow, Ryall and Marshall (1988),
and a large number may be required. Further, the Nopens and Vanrolleghem (2006) and Coufort
coupling of CFD and PSD models requires the et al. (2007). It is the work of Nopens which
incorporation of transport equations in each bin, would appear to have advanced the application of
making the process computationally expensive. CFD and PBM modelling to the greatest extent
Unlike the discrete method, moment methods thus far. However, it is clear from the literature
simulate statistical information about the PSD, that there has been no work undertaken on raw
rather than deriving an accurate description of the waters abstracted from WTW at either lab or full
PSD. With the SMM, the population balance scale. Consequently, there remains the need to
equation is transformed into a set of transport simulate accurately realistic water treatment
equations for moments of the distribution which, flocculation processes at both laboratory scale
for lower-order moments, are relatively easy to using raw water samples taken from WTW, and
solve. No prior assumptions are made with regard also at full scale.
to the size distribution and the moment equations
4. CONCLUSIONS
are closed (involving only functions of the
moments themselves). However, this latter point
The use and application of CFD within water
prevents particle aggregation and breakage being
treatment has expanded significantly over the
written as functions of moments, thus meaning
period 19952008. Examples of the techniques
that the method is only applicable in situations of
application can be found for most unit processes,
constant aggregation, size independent, and
from the storage of raw water, through detailed
growth. For the simulation of flocculation in
and technical water treatment processes, to
water treatment, these are significant limitations.
chemical disinfection. Model complexity varies
However, these limitations can be overcome via
from application of two-equation turbulence

236
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

models to straightforward geometries, to the use http://www.bakker.org/cfm. (Accessed 7th


of the RSM and sliding mesh technique, coupled November 2006).
with Eulerian or Lagrangian discrete phase model, 2. Biggs CA, Lant PA (2000). Activated sludge
in mixing scenarios. flocculation: On-line determination of floc
There are several published examples of where size and the effect of shear. Water Research
CFD has been applied to model bulk flow patterns 34:25422550.
in the complex, swirling flow found in 3. Boller M, Blaser S (1998). Particles under
mechanically mixed flocculators. However, there stress. Water Science and Technology
are few examples which extend the analysis to 37(10):929.
consider floc trajectory or fate. Lagrangian 4. Bouyer D, Coufort C, Lin A, Do-Quang Z
techniques are available, but are somewhat (2005). Experimental analysis of floc size
limited by the fractal nature of flocs and, in distributions in a 1-L jar under different
particular, the impacts of density and porosity on hydrodynamics and physico-chemical
drag force and settlement characteristics. conditions. J. Colloid Interface Sci. 292:413
CFD has been used effectively to demonstrate the 428.
limitations of the average velocity gradient 5. Bridgeman J, Jefferson B, Parsons SA (2007).
approach to classifying flocculators. The development and use of CFD models for
Whilst a Volume of Fluid model can be used to water treatment processes. CC2007: The
indicate the likelihood of vortex formation, it is Eleventh International Conference on Civil,
ill-equipped to deal with the bubble formation Structural and Environmental Engineering
arising from the breaking of the air/water Computing, St. Julians, Malta, 1821
interface when air passes through grid cells where September.
large momentum sources exist and so should not 6. Bridgeman J, Jefferson B, Parsons SA (2008).
be used to predict post vortex formation flow Assessing floc strength using CFD to improve
conditions. organics removal. Chemical Engineering
There remain several areas which require further Research and Design 86(8):941950.
study to facilitate their accurate representation 7. Camp TR, Stein PC (1943). Velocity
using CFD. In particular, the robust coupling of gradients and internal work in fluid motion.
population balance modelling for flocculation Journal of the Boston Society of Civil
processes with CFD models for water treatment Engineers 30(4):219237.
processes remains a key challenge. Progress has 8. Clark MM (1985). Critique of Camp
been made in wastewater flocculation; however, and Steins RMS velocity gradient.
there remains the need to simulate accurately J. Environmental Engineering 111(6):741
realistic water treatment flocculation processes at 754.
both laboratory scale using raw water samples 9. Cleasby JL (1984). Is velocity gradient a
taken from WTW, and also at full scale. valid turbulent flocculation parameter?
J. Environmental Engineering 110(5):875
ACKNOWLEDGEMENTS 897.
10. Coufort C, Bouyer D, Lin A (2005).
The submitted manuscript has been made possible Flocculation related to local hydrodynamics
through the funding from the American Water in a Taylor-Couette reactor and in a jar.
Works Association Research Foundation and Co- Chemical Engineering Science 60:21792192.
funding Utilities. The authors would also like to 11. Coufort C, Bouyer D, Lin A, Haut B (2007).
thank the Engineering and Physical Sciences Modelling of flocculation using a population
Research Council (EPSRC), Yorkshire Water, balance equation. Chemical Engineering and
Thames Water, United Utilities, Severn Trent Processing 46:12641273.
Water and Scottish Water and Fort Collins 12. Craig K, de Traversay C, Bowen B,
Utilities for their financial support. Essemiani K, Levecq C, Naylor R (2002).
Hydraulic study and optimisation of water
treatment processes using numerical
REFERENCES simulation. Water Science and Technology:
Water Supply 2(56):135142.
1. Bakker A, LaRoche RD, Wang MH, 13. Crittenden JC, Trussell RR, Hand DW,
Calabrese RV (2000). Sliding Mesh Howe KJ, Tchobanoglous G (2005). Water
Simulation of Laminar Flow in Stirred Treatment: Principles and Design. 2nd Ed.,
Reactors. In The Online CFM Book, John Wiley and Son.

237
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

14. Ducoste JJ, Clark MM (1999). Turbulence in and Wastewater Treatment VI, Proceedings
flocculation: comparison of measurements of the 9th Gothenburg Symposium. Ed. Hahn
and CFD simulations. AIChE J. 45(2):432 HH, Hoffmann E, Odegaard H, Springer, 89
436. 99.
15. Essemiani K, de Traversay C (2002a). 28. Kusters K, Wijers J, Thoenes D (1997).
Optimisation of the flocculation process using Aggregation kinetics of small particles in
computational fluid dynamics. Chemical agitated vessels. Chem. Eng. Sci. 52(1)107
Water and Wastewater Treatment VII, 121.
Proceedings of the 10th Gothenburg 29. Lain S, Phan L, Pellarin P, Robert P (1999).
Symposium. Ed. Hahn HH, Hoffmann E, Operating diagnostics on a flocculator-
Odegaard H, IWA Publishing. settling tank using FLUENT CFD software.
16. Essemiani K, de Traversay C (2002b). Water Science and Technology 39(4):155
Optimum design of coagulation / flocculation 162.
vessels. WQTC 2002 conference, Seattle, 30. Launder BE, Spalding DB (1974). Lectures in
USA. Mathematical Models of Turbulence.
17. Feng X, Li XY (2008). Modelling the kinetics Academic Press, London, UK.
of aggregate breakage using improved 31. Luo C (1997). Distribution of velocities and
breakage kernel. Wat. Sci. Tech. 57(1):151 velocity gradients in mixing and flocculation
157. vessels: comparison between LDV data and
18. Fluent Inc. (2005). Fluent 6 Users' Guide. CFD predictions. PhD Thesis, New Jersey
19. Freitas CJ (1995). Perspective: Selected Institute of Technology.
benchmarks from commercial CFD codes. J. 32. Luo JY, Issa RI, Gosman AD (1994).
Fluids Engineering 117:210218. Prediction of impeller induced flows in
20. Graber SD (1994). A critical review of the mixing vessels using multiple frames of
use of the G-value (RMS velocity gradient) in reference. IChemE Symposium Series No. 136,
environmental engineering. Dev. Theor. Appl. 549556.
Mech. 17:533556. 33. Marchisio D, Vigil R, Fox R (2003).
21. Haarhoff J, van der Walt JJ (2001). Towards Quadrature method of moments for
optimal design parameters for around-the-end aggregation-breakage processes. J. Col. Int.
hydraulic flocculators. J. Water Supply: Sci. 258(2):322334.
Research and Technology-AQUA 50(3):149 34. Marshall EM, Bakker A (2004).
159. Computational Fluid Mixing. In Handbook
22. Haque JN, Mahmud T, Roberts KJ (2006). of Industrial Mixing: Science and Practice.
Modelling flows with free-surface in Ed. Paul EL, Atiemo-Obeng VA, Kresta SM,
unbaffled agitated channels. Ind. Eng. Chem. 257343.
Res. 45:28812891. 35. McGraw R (1997). Description of aerosol
23. Hounslow M, Ryall R, Marshall V (1988). A dynamics by the quadrature method of
discretized population balance for nucleation, moments. Aero. Sci. and Tech. 27(2):255265.
growth, and aggregation. AIChE. J. 36. Menter FR (2003). Turbulence modelling for
34(11):18211832. turbomachinery. QNET-CFD Network
24. Iaccarino G (2000). Prediction of the Newsletter 2(3):1013.
turbulent flow in a diffuser with commercial 37. Mikkelsen L (2001). The shear sensitivity of
CFD codes. Centre for Turbulence Research activated sludge: relations to filterability,
Annual Research Briefs, 271278. rheology and surface chemistry. Coll. Surf. A
25. Jones SC (1999). Static mixers for water 182:114.
treatment. A computational fluid dynamics 38. Morsi SA, Alexander A (1972). An
model. PhD Thesis, Georgia Institute of investigation of particle trajectories in two-
Technology. phase flow systems. J. Fluid Mechanics
26. Korpijarvi J, Ahlstedt H, Saarenrinne P, 55(2):193208.
Renanen J (1999). Modelling of flow field in 39. Ng K, Borrett NA, Yianneskis M (1999). On
the mini-flocculator. IChemE Symposium the distribution of turbulence energy
Series No. 146, Fluid Mixing 6. Ed. Benkreira dissipation in stirred vessels. IChemE
H, 361372. Symposium Series No. 146, Fluid Mixing 6.
27. Korpijarvi J, Laine E, Ahlstedt H (2000). Ed. Benkreira H, 6980.
Using CFD in the study of mixing in 40. Nopens I (2007). Improved prediction of
coagulation and flocculation. Chemical Water effluent suspended solids in clarifiers through

238
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

integration of a population balance model and density function and aluminium floc. Water
a CFD model. PS-IWA 2007: Particle Research 13:409419.
Separation, Toulouse. 47. Thomas, DG (1964). Turbulent disruption of
41. Nopens I, Vanrolleghem PA (2006). flocs in small particle size suspensions.
Comparison of discretization methods to J. AIChE 19(4):517523.
solve a population balance model of activated 48. Van der Walt JJ (1998). The application of
sludge flocculation including aggregation and computational fluid dynamics in the
breakage. Mathematical and Computer calculation of local G values in hydraulic
Modelling of Dynamical Systems 12(5):441 flocculators. Proceedings of the Biennial
454. Conference of the Water Institute of Southern
42. Paul EL, Atiemo-Obeng VA, Kresta SM Africa, 1998, Cape Town, South Africa.
(2004). Handbook of Industrial Mixing: (Available at: http://www.ewisa.co.za/
Science and Practice. Wiley-Interscience. literature/files/1998%20-%2079.pdf, accessed
43. Prat OP, Ducoste JJ (2007). Simulation of February 2009.)
flocculation in stirred vesselsLagrangian 49. Versteeg HK, Malalasekera W (1995). An
versus Eulerian. Chem Eng Res. and Des. Introduction to Computational Fluid
85(A2):207219. Dynamics. The Finite Volume Method.
44. Randolf AD, Larson MA (1971). Theory of Prentice Hall.
Particulate Processes. Academic Press, New 50. Yeung AKC, Gibbs A, Pelton RP (1997).
York. Effect of Shear on the Strength of Polymer-
45. Rodi W (1993). Turbulence Models and Their Induced Flocs. Journal of Colloid and
Application in Hydraulics. 3rd Ed., IAHR. Interface Science 196:113115.
46. Tambo N, Wanatabe Y (1979). Physical
aspects of flocculationI The flocculation

APPENDIX 1

Standard k- Model

t k
( k ) + ( kui ) = + + Gk (A1)
t xi x j k x j

t 2
( ) + ( ui ) = + + C1 Gk C2 (A2)
t xi x j x j k k
where the turbulent viscosity, t , is calculated according to
k2
t = C (A3)

and the generation of turbulence kinetic energy due to mean velocity gradients, Gk ,
u u u
Gk = t i + j j (A4)
x
j xi xi
k and are the effective Prandtl numbers for k and , respectively.
The empirical model constant values are generally accepted as C = 0.09 , C1 = 1.44, C2 = 1.92, k = 1.0 and
= 1.3.

Renormalized Group (RNG) k- Model

k
( k ) + ( kui ) = ( k eff ) + Gk (A5)
t xi x j x j

239
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

2
( ) + ( ui ) = ( eff ) + C1 Gk C2 R (A6)
t xi x j x j k k

eff = + t
k and are the inverse effective Prandtl numbers for k and , respectively.

Realizable k- Model

t k
( k ) + ( ku j ) = + + Gk + S k (A7)
t x j x j k x j

t 2
( ) + ( u j ) = + + C1 S C 2 + C1 C3 Gb + S (A8)
t x j x j x j k + k

C2 is a constant. k and are the turbulent Prandtl numbers for k and , respectively. Sk and S are user-
defined source terms. The empirical model constant values are generally accepted as C1 = 1.44,
C1 = max 0.43, , C2 = 1.9, k = 1.0 and = 1.2.
+ 5

k- Model

k
( k ) + ( kui ) = k + Gk Yk + S k (A9)
t xi x j x j


( ) + ( ui ) = + G Y + S (A10)
t xi x j x j

k and represent the effective diffusivity of k and respectively


t
k = + (A11)
k
t
= + (A12)

Gk represents the generation of turbulence kinetic energy due to mean velocity gradients. G represents the
generation of . Yk and Y represent the dissipation of k and due to turbulence. Sk and S are user-defined
source terms.

SST k- Model

k
( k ) + ( kui ) = k + Gk Yk + S k (A13)
t xi x j x j


( ) + ( ui ) = + G Y + D + S (A14)
t xi x j x j

D represents the cross-diffusion term.

240
Engineering Applications of Computational Fluid Mechanics Vol. 3, No. 2 (2009)

Reynolds Stress Model

Considering an incompressible flow without body forces and ignoring transport due to rotation, the exact
equation for the transport of the Reynolds stresses (= u i u j ) is

t
( uiuj ) +

xk
( uk uiuj ) =

xk
[ ]
uiuj uk + p( kj ui + ik uj ) +

xk



(uiuj )
xk
local time convection turbulent diffusion molecular diffusion
derivative
u u u u u uj
uiuk j + uj uk i + p i + j 2 i
xk xk x x x x
j i k k
stress production pressure strain dissipation

APPENDIX 2

Drag coefficient parameters (from Morsi and Alexander, 1972):

2 3
C D = 1 + +
Re Re 2

Re 1 2 3
< 0.1 0 24 0
0.11.0 3.69 22.73 0.0903
1.010 1.222 29.1667 -3.8889
10100 0.6167 46.5 -116.67
1001,000 0.3644 98.33 -2778
1,0005,000 0.357 148.62 -4.75 x 104
5,00010,000 0.46 -490.546 57.87 x 104
10,00050,000 0.5191 -1662.5 5.4167 x 106

241

You might also like