You are on page 1of 11

Advances in Engineering Software 41 (2010) 99–109

Contents lists available at ScienceDirect

Advances in Engineering Software


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

The development and application of CFD models for water treatment flocculators
J. Bridgeman a,*, B. Jefferson b, S.A. Parsons b
a
School of Engineering (Civil Engineering), University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK
b
Centre for Water Science, School of Applied Sciences, Cranfield University, Cranfield, Bedfordshire, MK43 0AL, UK

a r t i c l e i n f o a b s t r a c t

Article history: The use of CFD to simulate turbulent flows in laboratory and full scale flocculation processes commonly
Received 21 December 2007 found at water treatment works is reported. The paper considers a range of modelling strategies and sim-
Accepted 2 December 2008 ulation techniques including, inter alia, steady and unsteady flow, two-equation and Reynolds stress tur-
Available online 21 January 2009
bulence modelling, sliding mesh and multiple reference frames approaches to rotational flow simulation,
and mesh density optimisation. Through analysis of turbulence dissipation rate, this work considers the
Keywords: development of models for environmental engineering problem solving and demonstrates clearly the
CFD
benefits to be gained from the use of CFD in flocculation vessels.
Mixing
Rotating mesh
Ó 2008 Civil-Comp Ltd. and Elsevier Ltd. All rights reserved.
Turbulence
Water treatment

1. Introduction has attempted to model the process using CFD. The research which
has been published [3–6] is limited in scope. These papers consid-
The destabilisation (via coagulation) and subsequent agglomer- ered the flow fields generated by small scale flocculators and
ation (via flocculation) of fine particles and colloids into larger par- recognised the limitations of the existing design parameters. How-
ticles is a proven means of removing impurities (e.g. turbidity and ever, none of the papers considered, other than qualitatively, how
colour) at water treatment works (WTW). Chemical coagulant these results would impinge on flocculation performance.
addition brings about a change in the nature of small particles, ren- In this work, the commercially-available CFD code, Fluent v6
dering them unstable, whilst flocculation encourages particle (Ansys Inc., Sheffield, UK) was used to model the flow field within
agglomeration via gentle mixing and the formation of irregularly- flocculators at both laboratory and full scale. The focus of the work
shaped, loosely connected mass fractal aggregates, known as flocs. concentrated on the development and application of simulation
The size and structure of flocs are considered fundamental to the techniques and the analysis of computational models to increase
efficient operation of WTW. Ineffective coagulation and floccula- understanding of floc formation and breakage mechanisms. Previ-
tion result in poorer quality feed water to downstream treatment ous work [7] provided a quantitative analysis of floc growth and
processes, potentially jeopardising treated water quality and breakage. The work reported here complements that previous
increasing operational costs. Several interrelated criteria govern work, providing a fundamental understanding of the hydrody-
the efficiency of the coagulation and flocculation stages; viz. coag- namic environments to which flocs are exposed at a range of
ulant type and dosage, pH and mixing arrangements. scales.
Particle removal efficiency decreases with decreasing particle
size [1]. Therefore, flocs must be able to withstand shear energy 2. Flocculation
applied to them in various different unit processes. When the de-
gree of shear exceeds a threshold value, floc breakage will occur. 2.1. Conceptual view of flocculation
However, quantification of the energy requirements for floc break-
age is not straightforward, and despite much work in this field, no Flocculation is the transformation of smaller destabilised parti-
standard strength test exists. cles into larger aggregates or flocs, with the rate of growth gov-
The literature contains examples of practical investigations into erned by the rate of inter-particle collisions. Flocs are mass
floc strength at laboratory scale using a standard jar test apparatus fractal objects whose density approximates to that of water [8],
and procedure [2]. However, there is little reported research which but is found to decrease as size increases and so as the flocs grow
in size, induced shear forces may give rise to floc breakage. The rate
of floc growth has been expressed conceptually as [2]:
* Corresponding author. Tel.: +44 121 414 5145; fax: +44 121 414 3675.
E-mail address: j.bridgeman@bham.ac.uk (J. Bridgeman). Rfloc ¼ aij Rcol  Rbr ð1Þ

0965-9978/$ - see front matter Ó 2008 Civil-Comp Ltd. and Elsevier Ltd. All rights reserved.
doi:10.1016/j.advengsoft.2008.12.007
100 J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109

where Rfloc represents the overall rate of floc growth, a is the colli- 2.2. Velocity gradient
sion efficiency factor (0 < a < 1), Rcol is the rate of particle collision
and Rbr is the rate of floc breakage. Previous workers [2] have shown that floc size is dependent on
More accurately, the aijRcol term should be broken down further the turbulence energy dissipation rate. The concept of the absolute
into its constituent parts where aijRcol is a function of velocity gradient, G (s1), which encapsulates the turbulence en-
BM
aij BM Rcolij ; Sh aij Sh Rcolij ; and DS aij DS Rcolij and i and j refer to discrete ergy dissipation rate, was developed by Camp and Stein [13], and
particles, and BM, Sh and DS refer to the collision mechanisms of is defined as:
Brownian Motion, shear and differential settlement, respectively. sffiffiffiffiffiffiffiffiffi rffiffiffi
Floc size may be considered to be a balance between the hydro- P=V e
dynamic forces exerted on a floc and the strength of the floc [9,10]. G¼ ¼ ð8Þ
l m
As a result, flocs do not continue to grow throughout the floccula-
tion stage, but rather, attain a limiting size beyond which breakage where P is the power dissipated, V is tank volume, l is dynamic vis-
prevents further overall growth. Where the floc strength is resis- cosity of the water, e is energy dissipation rate per unit mass, and m
tant to the hydrodynamic forces, one would expect floc size either is the kinematic viscosity of the water.
to remain constant or for growth to occur. Where the hydrody- In theory, the absolute velocity gradient can be calculated at
namic forces exceed floc strength, floc breakage will occur. Conse- any point within a mixing vessel, provided that the power dissi-
quently, the conceptual growth breakage mechanism may be pated is known at that point. In practice, however, the flow charac-
expressed as follows [10]: teristics vary within the mixing vessel from point to point, and so
too does the energy dissipation. Consequently the velocity gradient
Hydrodynamic forces F
B¼ ¼ ð2Þ is a function of both time and position. Given the difficulties asso-
Floc strength J ciated with calculation of G, workers have traditionally replaced
the absolute velocity gradient with an approximation of the exact
where F represents the hydrodynamic forces exerted by the flow,
value; that is its average value throughout the vessel, Gave:
and J represents the strength of the floc. It is clear from Eq. (2) that
breakage will occur when B > 1, and floc size will be maintained or sffiffiffiffiffiffiffiffiffi
increased when B < 1. Floc strength, J, is a function of the physico- Pav e
Gav e ¼ ð9Þ
chemical conditions (raw water type, coagulant type and dose) Vl
and the floc structure [2].
Previous workers [11] suggested that the hydrodynamic force where the average power consumption, Pave, is readily obtained
required to pull apart a floc in tensile mode may be expressed via:
as:
Pav e ¼ P0 qN3 D5 ð10Þ
p 2
F rd ð3Þ and P0 is the impeller power number, q is the fluid density, N is the
4
rotational speed of the impeller, and D is the impeller diameter.
where r represents the hydrodynamic stress exerted on thepfloc, ffiffiffiffiffiffiffiffi The average velocity gradient value, Gave, is used worldwide to
and d is the area of the floc. In the viscous subrange, r ¼ l  e=m. characterise mixing in a wide range of environmental engineering
Substituting into Eq. (2) shows that the breakage mechanism in applications, including flocculators. However, since its introduc-
the viscous sub-range, BVSR, may be expressed as: tion, authors have argued that the concept of the Gave value is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi flawed, as it attempts to represent a complex flow field within a
C1l e=m  d2 single number [14]. The distribution of velocity gradients within
BVSR ¼ ð4Þ a stirred tank is clearly not uniform, and local power consumption
J
at a point of high turbulence within a vessel (e.g. adjacent to the
where C1 is a constant. impeller) can be several orders of magnitude in excess of the rest
In the inertial sub-range of the vessel [14,15]. Thus, the Gave value does not describe ade-
quately the fluctuations in local velocity gradient experienced at
r ¼ q  u2 ð5Þ a specific point within the flow field. This is unfortunate as it is pre-
cisely the magnitude and fluctuations in local shear to which a floc
and is subjected, and not the average value, which determine the like-
 2 ¼ C 2 ðe  dÞ2=3 ½12
u ð6Þ lihood of floc agglomeration, breakage and re-growth. CFD offers
the opportunity to quantify and understand the local impact of
Substituting into Eq. (2) shows that the breakage mechanism in mean flow and turbulence on floc formation and break-up via
the inertial sub-range, BISR, may be expressed as: the concept of the local velocity gradient, GL, where:
C 2 qe2=3  d
8=3 rffiffiffi
BISR ¼ ð7Þ e
J GL ¼ ð11Þ
m
where C2 is a constant.
Consequently, it is deduced from Eqs. (4) and (7) that floc size is Using CFD, for the first time GL can be calculated for any and all
dependent on the turbulence energy dissipation rate and floc points in the mixing vessel, to give a complete distribution of val-
strength, irrespective of subrange. This is clearly of great signifi- ues, rather than a simple average value.
cance when one attempts to gain an understanding of floc break-
age mechanisms and limiting floc size as it directs workers to 2.3. Laboratory scale flocculation
focus their efforts towards e and J. Thus, it is both physico-chemical
and turbulent conditions within the containing vessel which con- Coagulation and flocculation optimisation are generally consid-
trol the flocculation process. In water treatment, the velocity gradi- ered at the laboratory scale, using a jar test apparatus and proce-
ent term is used to characterise mixing and thereby predict dure. The test apparatus typically comprises a glass vessel with a
aggregation kinetics and break-up phenomena. powered paddle to stir the contents. No absolute standard apparatus
J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109 101

Fig. 1. Mechanical and hydraulic flocculation.

or procedure exists for the jar test. However, typically the paddle is ples of CFD analyses of full scale water treatment unit processes,
set to rotate at high speed for a short period, during which time there are very few examples of numerical studies of either
coagulant (and acid or alkali for pH adjustment as necessary) is mechanical or hydraulic flocculators. A thorough search of the lit-
added to each vessel and mixed. After a short period of intense erature found just two published, peer-reviewed articles relating
mixing (30–60 s) to simulate the coagulation process, the paddle to the performance of a full scale hydraulic flocculator [17,18]. In
speed is reduced to a produce more gentle mixing to simulate the view of the apparent lack of attention paid to CFD studies of floc-
flocculation process. After 20 min of flocculation, mixing is termi- culators, it was considered appropriate for this work to consider
nated, the paddle removed and the suspension allowed to settle. one example of each installation in order that their design and
The flocs and treated water are then analysed for floc size and operation might be considered in relation to each other. In com-
strength, allowing conclusions to be drawn regarding optimum mon with some previous research, the work reported in this paper
coagulant dose and pH. These conclusions are often then applied considers flow patterns but, unlike previous work, here we con-
to the operation of the main treatment plant with no account taken sider the impact on floc fate. However, it is not just the hydrody-
of scaling characteristics. namic behaviour of flocculators which is of interest, but the
Various different jar test configurations are in use today and the agglomeration, growth and breakage processes are significant also.
literature contains many examples of practical investigations into Multiphase models which incorporate a particle size distribution
flocculation and floc strength using the jar test apparatus and pro- (PSD), such as those found in a flocculator, require a population
cedure. A review of 34 different experimental approaches used by balance model (PBM) to describe particle population changes. Sev-
several workers in recent years can be found at [16]. Thus, it is eral solution methods exist, however a full description of each is
clear from the literature that many jar test experiments have been outside the scope of this paper. Recent work [19] provided proof
conducted in which the shape of the containing vessel, the shape of concept for the decoupling of fluid flow and flocculation dynam-
and location of the impeller, the volume of water, and the mixing ics but did not provide information regarding the growth and
speed have all differed. Within the literature, data from previous strength of flocs in specific water treatment applications. Other re-
work are often compared with new data, and workers have com- cent work includes that of Nopens [20] (CFD and PBM in wastewa-
pared and contrasted the chemical characteristics of flocs and ter clarifier) and Feng and Li [21] (theoretical PBM approach,
coagulants in detail. However, it is known that physical parame- encapsulating internal body forces and fluid shear stress). Whilst
ters, such as mixing intensity and retention time, bring a signifi- a detailed review of all recent PBM work is not appropriate here,
cant bearing on coagulation and flocculation efficiency [16]. It is it is the work of Nopens which has advanced the application of
interesting to note that there is usually no allowance made for CFD and PBM modelling to the greatest extent thus far. However,
the different conditions under which the experiments have invari- it is clear from the literature that there has been no work under-
ably taken place and one of the key elements of this work was to taken on raw waters abstracted from WTW at either lab or full
examine how, if at all, changes to the configuration of the jar test scale. Consequently, there remains the need to simulate accurately
experiments might affect results. realistic water treatment flocculation processes at both laboratory
scale using raw water samples taken from WTW, and also at full
2.4. Full scale flocculation scale.

In practice, full scale flocculation at WTW is effected via either


3. CFD modelling of jar tester
mechanical or hydraulic means. For mechanical flocculation, the
energy input is via an agitator which generates the necessary shear
3.1. Jar test configurations
stress (and, hence, velocity gradients), and power dissipation may
be obtained via Eq. (10). In hydraulic flocculation, energy is im-
The Fluent software (v6, Ansys Inc., Sheffield, UK) was used to
parted via the headloss across a baffled, serpentine channel and
model the flow field within two sets of jar test apparatus and to ex-
power is obtained via consideration of that headloss. The velocity
plore possible relationships between the turbulence dissipation
gradient is then obtained via
sffiffiffiffiffiffiffiffiffi rate and mixing speed in the two vessels. The first configuration
qgh comprised an unbaffled 100 mm diameter Perspex stirred vessel,
G¼ ð12Þ liquid depth 127 mm, containing one litre of water (the ‘‘cylindri-
lt
cal vessel”). This replicated the geometry adopted by previous
where h is the headloss across the channel, t is the retention time workers [7]. The second configuration comprised a 115 mm square
within the channel. The two processes are shown schematically in section Perspex vessel, liquid depth 150 mm, containing two litres
Fig. 1. of water (the ‘‘square vessel”), replicating another commonly used
Having expended time and energy in developing flocs, it is geometry (e.g. [15]). The contents of both vessels were mixed with
important that operating conditions do not subsequently cause a 76 mm diameter, 25 mm deep paddle impeller, mounted on a
their break up. As a result, floc strength, growth and breakage have 6 mm diameter shaft. For the cylindrical vessel, the paddle offset
been the subject of detailed research in recent years [1,2,7]. How- from the base of the vessel was 35 mm, whereas the corresponding
ever, much of this work has been undertaken experimentally at distance for the square vessel was 53 mm. The two configurations
laboratory scale. Whilst the literature contains a number of exam- are shown in Fig. 2.
102 J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109

where, um, is velocity magnitude at point m, h is the number of cells


in the mesh and subscripts 1 and 2 refer to coarse and fine mesh,
6
6 respectively.
A factor of safety of 3 (i.e. Fs = 3.0) was applied in accordance
with published recommendations [23]. Taking the 87,777 mesh
as the baseline, CGIs were calculated for the finer grids as 3.75
25

150
127

(98,517), 2.89% (283,470) and 2.31% (337,944). The increase in grid


25
density between the 98,517 and 283,470 cell grids led to an in-
crease in computing time from 16 h to 24 h to achieve a converged
76
solution. It was concluded that the increased computational effort

53
30

76 was not reflected in a corresponding increase in accuracy of results


and so the 98,517 cell grid was selected as the model for validation
100 115 in this study.
Section A-A Section B-B
For each geometry, converged solutions were produced for mix-
ing speeds of 30, 40, 50, 75, 100, 150 and 200 rpm, matching those
used previously [7].

6
3.3. Model validation
A 6 A B B
115

A two phase approach was adopted to model validation; firstly


using laser Doppler anemometry (LDA) measurements to validate
velocity components, and secondly using power input measure-
ments to validate power dissipation.
Plan Plan
Two-channel LDA measurements were undertaken on the jar
Fig. 2. Modelled cylindrical and square vessel configurations (not to scale, all
test apparatus using a laser Doppler anemometer operating in dual
dimensions in millimetres). beam mode. The system included a 3 W argon-ion laser, beamsplit-
ter and Bragg cell for frequency shifting. The signal was fed to a
correlation-type signal processor and the amplified Doppler signal
3.2. Model development was monitored on a digital oscilloscope. Signal processing was per-
formed on a personal computer. The axial and tangential compo-
During the development and use of any CFD model, several nents were measured simultaneously along an axial profile at
project-specific criteria must be taken into account, including x = y = 0.025 m (i.e. midway between shaft and wall). Twenty-three
the choice of turbulence model, grid density, and the selection measurements were taken along the profile at 5 mm intervals and
of appropriate boundary conditions. The case of the jar test, at seven different mixing speeds; viz. 30, 40, 50, 75, 100, 150 and
where there is highly swirling flow within a constricted volume, 200 rpm. (The percentage error in the calibration factor was
presents some specific and interesting challenges for the model- approximately 1.5%).
ler. For the cylindrical vessel, sensitivity analyses were under- CFD model results and LDA axial velocity data for a mixing
taken on four grid densities (87,777, 98,517, 283,470, and speed of 100 rpm in the cylindrical vessel are compared in
337,944 cells), six turbulence models (Standard k–e (Sk–e), Real- Fig. 3. Numerical data are provided for the RSM and Rk–e turbu-
izable k–e (Rk–e), Renormalised Group k–e (RNG k–e), Standard lence models, used in conjunction with both multiple reference
k–x (Sk–x), Shear Stress Turbulence k–x (SST k–x) and Rey- frames (MRF) and sliding mesh (SM) techniques, and also for the
nolds Stress Model (RSM)) and two treatments of the rotating RNG turbulence model used with the sliding mesh technique, all
mesh (sliding mesh and multiple reference frames). Full descrip- applied to the 98,517 cell mesh. Good correlations between the
tions are outside the aegis of this paper, but may be found at observed and MRF RSM data are observed, particularly in the pad-
[22]. dle region and below. There would appear to be some loss of accu-
Free slip conditions were specified at the water surface and no racy around tank mid-depth and upwards for all models except
slip conditions were specified at the walls. The near-wall region the SM RSM simulation. This is likely to be as a result of vortexing
was modelled via standard semi-empirical logarithmic wall func- within the vessel. The correlation between observed data and the
tions. To initialise the model, all velocities were set to zero, whilst Rk– data (both MRF and SM) is rather more inferior, particularly
k and e were set to 0.001 m2 s2 and 0.01 m2 s3, respectively. For for the MRF Rk– simulation, with large errors found at mid-depth
all simulations, the residuals to assess convergence (continuity, and also in the region below the paddle. The SM RNG models
velocity components, turbulent kinetic energy, turbulence dissipa- shows reasonable accuracy in the lower tank region, but this is
tion rate, specific dissipation rate, and Reynolds stresses, where lost in the paddle region and above.
applicable) were set to 1  105. Using the energy dissipation rate calculated from each turbu-
Grid density analysis was undertaken using the grid conver- lence closure model, the overall power consumption was calcu-
gence index (CGI) method to provide an indication of error bands lated for each mixing speed by numerically integrating the local
[23]. Velocity magnitude values were extracted for 1000 individual power consumption over the entire vessel contents, expressed as:
points in the flow field, and the CGI calculated as: Z
P¼q edV ð14Þ
erms
GCI ¼ F s 2
r 1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Actual power input was calculated from applied torque measure-
P1000 2 ments via:
m¼1 jðu m;1  u m;2 Þ=u m;2 j
erms ¼ ð13Þ
1000 P ¼ 2  p  N  Tq ð15Þ
 1=3
h2 where Tq is the applied torque, measured on a torque transducer

h1 unit.
J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109 103

140

120

100
Depth, mm

80

60

40

20

0
-0.05 -0.03 -0.01 0.01 0.03 0.05
-1
Axial Velocity, ms

LDA MRF RSM MRF rk- ε SM rk- ε SM RSM SM RNG k- ε

Fig. 3. Simulated axial velocity profiles calculated for various modelling approaches, over tank depth compared to experimentally-derived value (N = 100 rpm).

Values for the vessel average velocity gradient were then calcu- tions of two-equation models in predicting highly swirling flow
lated for both the experimental and CFD results using Eq. (9) and patterns, and the expense of the SM approach, the RSM MRF com-
the results for a mixing speed of 100 rpm are shown at Table 1. bination was selected for all jar test CFD models.
The MRF with Sk–e solution is clearly wholly unrepresentative of Similar exercises were undertaken for the square vessel. Two
the actual flow field, as this modelling combination yields a Gave separate grids were developed, one with 144,718 cells and the
of 989 s1, compared to an experimentally-derived value of other with 281,785 cells. Excellent agreement was observed be-
104 s1. The Rk–e and RS models, used in combination with the tween two sets of axial velocity data. However, the increase in
MRF treatment for rotating meshes, produced reasonable results computing time when undertaken on a modest laptop was signif-
when compared to the work of previous authors [5,6,24]. SM and icant. Convergence for the coarser of the two grids was achieved in
Rk–e also produced acceptable results. 17 h, whereas the finer grid took 30 h to converge. In view of the
It is widely acknowledged that the Reynolds Stress Model is the good agreement between the two datasets, and taking account of
most rigorous of the available models [22], and this was indicated the increased computing time associated with the finer grid, it
by the increased simulation time required when using this model was decided to use the model grid density of 144,718 cells for fu-
(an additional 40–50% when compared to two-equation models ture work.
and using the MRF approach). Furthermore, there is an inherent
difficulty with all two-equation models when modelling this type 3.4. Jar test simulation results
of swirling flow as they are all predicated on the assumption that
turbulence is isotropic, meaning that the turbulent stress tensor Contours of velocity magnitude for the two vessels taken
is independent of direction (or, more precisely, invariant with re- through the horizontal plane corresponding to the centreline of
spect to rotation and reflection of the coordinate axes of the coor- the paddle are shown in Fig. 4. For both vessels it is apparent that
dinate system moving with the mean motion of the fluid). This the highest velocities are limited to the paddle tip zone. There is
assumption is not true for swirling flow which may be anisotropic good distribution of velocities in the cylindrical vessel whereas
in nature. For swirling flow, turbulence is generally more in the the same is not true of the square vessel. High values are apparent
tangential and axial directions as compared to the radial direction.
The effects of strong turbulence anisotropy can be modelled rigor-
ously only by the second-moment closure adopted in the RSM.
The modelling strategy exercise confirmed that the sliding
mesh approach, coupled with the RSM, provided the greatest de-
gree of accuracy for the determination of velocity data but under-
estimated power dissipation. The MRF RSM simulation was found
to offer reasonable correlation with the LDA data and the experi-
mentally-derived power dissipation, at a substantially reduced
computational cost. Consequently, in light of the good correlation
found between the RSM results and the observed data, the limita-

Table 1
Vessel average velocity gradient values, Gave, s1 for cylindrical vessel, N = 100 rpm.

Scenario Expt MRF and MRF and MRF and SM and SM and SM and
Sk– Rk– RSM RNG Rk– RSM
Gave, s1 104 989 77 71 62 72 33
Fig. 4. Contours of velocity magnitude taken at paddle mid-depth N = 50 rpm.
104 J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109

Using the cylindrical vessel model, turbulence dissipation rate


values were extracted from a plane in the wake of the paddle when
mixing at 200 rpm (Fig. 7). It is apparent that much of the turbu-
lence dissipation takes place within, and immediately adjacent
to, the impeller zone. A qualitatively similar result was obtained
previously when considering mixing in the square vessel at
60 rpm [15], but this was not related in any quantitative fashion
to floc breakage and size.
In order to develop the understanding of the magnitude and
distribution of the turbulence dissipation rate in each vessel fur-
ther and to link this to impacts on floc breakage, four volumes of
interest were identified and studied for each of the two vessel
models; specifically:

(i) vessel contents – i.e. all liquid in vessel,


(ii) paddle swept volume (PSV) – i.e. volume of vessel traced by
Fig. 5. Contours of velocity magnitude taken z = 2 mm, N = 50 rpm. paddle,
(iii) continuum – i.e. that part of the vessel not included within
the paddle swept volume,
in the paddle wake for both scenarios, but they dissipate quickly in (iv) paddle tip zone (PTZ) – i.e. an annulus extending 1.5 mm
the square vessel, with evidence of dead spots in the vessel corners. beyond the paddle tip in both the radial and vertical
Similar observations were found in the vertical plane (not shown) directions.
with the cylindrical vessel demonstrating good coverage from the
surface to just above the base. Velocities calculated 2 mm above The volume-specific average turbulence dissipation rate val-
the vessel bases at a mixing speed of 50 rpm are shown in Fig. 5. ues were found to exhibit the hierarchy paddle tip zone > paddle
Velocities in the square vessel are approximately 50% of those in swept volume > vessel > continuum for both the cylindrical and
the cylindrical vessel and there are significant dead spots in the square vessels. Data for the two vessels are compared directly
square vessel. Overall, an increased distribution of the higher in Fig. 8, where the results for the paddle tip zone and vessel
velocities in the cylindrical vessel compared to the square vessel averages are plotted together. It is apparent from Fig. 8 that
was observed. The implication of these observations is that the the square vessel exhibits higher average turbulence dissipation
supplied mechanical energy is dissipated throughout the majority rate values than the cylindrical vessel. It is also clear that the
of the cylindrical vessel, generating a greater degree of mass flow rate of increase in e with increasing mixing speed is greater in
within it. The same is not true of the square vessel indicating im- the square vessel than in the cylindrical vessel; this observation
proved mixing in the cylindrical vessel than the square. being particularly true in the paddle tip zone. Similar shaped
Considering the link between turbulence dissipation rate, e, GL curves for the PSV and vessel average for the square vessel using
and floc strength [2,16], an informed understanding of e in various LDA-derived data have been published and these new CFD data
different scenarios (i.e. test configurations and mixing speeds) can show good agreement [15].
be used to develop important information with regard to floc The differences between the e values found in the different
behaviour. Plots of the variation of e over the mid-planes of the zones were investigated further. For the cylindrical vessel it was
two vessels are shown in Fig. 6 for the 200 rpm scenarios. For both found that, considering the full range of mixing speeds, the average
scenarios, there is a strong non-uniformity of turbulence dissipa- e value in the paddle tip zone was 3.5 times greater than that found
tion. In the impeller zone, the turbulent kinetic energy is highly in the vessel overall (i.e. eptz = 3.5  evessel). However, the corre-
dissipated, whereas outside of the impeller zone, the dissipation sponding relationship for the square vessel shows that
rate is reduced. The higher proportion of larger e values in the eptz = 5.0  evessel. This result agrees well with previous work in
square vessel is evident. which LDA-derived data showed that for the square vessel at mix-
ing speeds of 100 rpm and below 4.0 < eptz: evessel < 6.0 [15]. It is
apparent from the results that the degree of turbulence generated
in the paddle tip zone in the square vessel is significantly greater
than the cylindrical vessel. However, it is interesting to note that
for both vessels econtinuum/evessel = 0.5 (i.e. the average e value in
the continuum is half of that found in the vessel overall). This re-
sult further confirms the conclusion drawn above that although
higher e values are generated in the square vessel, they are less
well dissipated than in the cylindrical vessel.
There is a significant range of velocities and turbulence dissipa-
tion rates, and consequently velocity gradients, found in the two
vessels, with the highest values found in the paddle swept volume
and the paddle tip zone. The above results demonstrate that the
hydrodynamic environments generated by the same paddle at
the same mixing speeds but in different vessels can be significantly
different. The logical extension of this conclusion is that it is nei-
ther appropriate nor valid to compare results from one jar test sce-
nario to another when the configurations are different. It is clear
from the above that comparisons of floc behaviour made between
Fig. 6. Contours of turbulence dissipation rate for cylindrical and square vessels, different jar test configurations (of which there are several in the
N = 200 rpm. literature) must be treated with caution.
J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109 105

0.4000

0.3500
Dissipation Rate, m 2/s3
0.3000

0.2500

0.2000

0.1500

0.1000

0.0500

0.0000
0.0409
0.000

0.013

0.027

0.0234
0.040

0.053

0.067 Radial Distance, m

0.080
0.0058

0.094

0.107

0.120
Depth, m

0-0.05 0.05-0.1 0.1-0.15 0.15-0.2 0.2-0.25 0.25-0.3 0.3-0.35 0.35-0.4

Fig. 7. Variation in local turbulence dissipation rate in cylindrical vessel at 200 rpm.

4. CFD modelling of full scale flocculators the production of high density flocs quickly in the first cell, encour-
ages growth rather than breakage in the second cell, with a mini-
4.1. Description mum amount of turbulence generated in the final cell to avoid
floc breakage but sufficient to maintain suspension in the floccula-
Two full scale flocculators at operational WTW were selected tor. Each cell comprises a rectangular section vessel, 6.4 m 
for modelling and analysis; one a mechanical flocculator and one 7.0 m  4.75 m deep, within which a four blade, 6.0 m diameter
an hydraulic flocculator. The mechanical system considered in this rake paddle rotates slowly.
work comprises three identical vessels and paddles rotating at dif- The hydraulic system considered here is a rudimentary process
ferent, reducing speeds. This way of tapering flocculation enhances of coagulation and hydraulic flocculation through a mixing

1.00
Average Turbulence Dissipation Rate, ε, in Zone, m2/s3

0.10

0.01

0.00

0.00
25 50 75 100 125 150 175 200
Mixing Speed, N, rpm

Square Vessel Square Paddle Tip Zone Cyl Vessel Cyl Paddle Tip Zone

Fig. 8. Comparison of average turbulence dissipation rate in paddle tip zone and vessel for the cylindrical and square vessels.
106 J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109

Table 2 To develop a converged solution for one flocculation cell over 15


Model and convergence time summary. revolutions using either the Sk–e or Rk–e model took approxi-
Vessel Mesh Rotation Turbulence model Machine CPU time mately five days’ CPU time using a high performance computing
size (Hours) (HPC) facility. The work undertaken here made use of a Sun cluster
Cylindrical 98,517 MRF RSM Laptop 16 grid, comprising 19 2  3.06 dual processor computers, each with
Cylindrical 283,470 MRF RSM Laptop 24 6GB RAM available, all managed by the Sun grid engine from which
Square 144,718 MRF RSM Laptop 17 100% of a processor could be made available for up to five days at a
Square 281,785 MRF RSM Laptop 30
Cylindrical 98,517 MRF Sk–e, Rk–e, k–x, Laptop 12
time. Modelling in excess of one vessel at any one time was not
SST k–x achievable with the available resources. For each mixing speed,
Cylindrical 98,517 SM RNG HPC 17 three different flow rates were simulated (21.1, 24.0 and 16.0 Ml/
Mechanical 338,213 MRF Sk–e, HPC 24 d) representing the average, maximum and minimum flows
Mechanical 338,213 SM Sk–e, Rk–e, HPC 120
through the vessels. Thus, nine separate simulations were
Hydraulic 395,328 – Sk–e HPC 6
Hydraulic 776,187 – Sk–e HPC 12.25 undertaken.
It was considered that the flow generated in the hydraulic floc-
culator would not follow a swirling path which would have neces-
chamber and single baffled channel. The channel is 2.5 m wide, sitated the use of a rigorous turbulence modelling approach such
4.165 m deep and 50 m long. Vertical baffles are positioned at reg- as the Reynolds Stress Model. The Sk–e model has been used suc-
ular intervals along the length of the channel, causing the flow to cessfully in previous work undertaken on similar structures,
move in a serpentine manner. [17,18,25], and was chosen in this instance also. As the hydraulic
flocculator runs at average output for the majority of the time of
4.2. Model development the time, just one flow rate was modelled; that being 160 Ml/d.

The software packages, Gambit and Fluent (Ansys Inc., Sheffield, 4.3. Full scale flocculator results
UK) were used to generate the mesh and solve the equations of
flow in both cases. The mesh, turbulence modelling and rotating 4.3.1. Velocity
mesh details (where appropriate) are shown at Table 2. Plan and Simulated velocity contours for one quadrant of the mechanical
sectional views of the two flocculators are shown in Figs. 9 and 10. flocculator are shown in Fig. 11 (mixing speed = 2.4 rpm, flow
Appropriate boundary conditions were specified; in particular, rate = 21.1 Ml/d). Some small dead spots at the vessel corner and
free slip conditions at the top water level, no slip conditions at around the shaft are apparent; otherwise there is a reasonable de-
the walls, standard logarithm wall boundary layer profile, normal gree of movement within the vessel. Fig. 11 indicates that the high-
velocity inlet and pressure outlet. est velocities are located towards the tips of the paddles. The
The mechanical flocculator offered the most significant chal- velocity varies from 0 to 0.89 ms1 within the vessel; the upper
lenges in deriving a converged solution. In order to derive an initial limit being some 1.83 times greater than the inlet velocity of
solution, the rotating grid was first modelled using the Sk–e turbu- 0.486 ms1. The corresponding values for the 1.6 and 3.2 rpm sim-
lence model and the MRF method. (Convergence criteria were set ulations were 1.23 and 2.42, respectively.
at 1  105). These data were then interpolated into a sliding mesh For the hydraulic flocculator, a plot of velocity vectors calcu-
version of the grid and the analysis refined using that method of lated on a vertical plane along a section of the centreline of the
solution. 100 time steps per revolution were specified, and, in channel is shown in Fig. 12. Working from the left hand side of
the first instance, 15 revolutions were specified in order to ensure the plot, it is apparent that the flow at the surface strikes the baffle
a steady solution had been reached. Point velocities and pressures wall and is deflected vertically downwards. A significant propor-
were monitored to ensure that a time-independent solution had tion of the flow continues under the baffle wall and into the next
been reached. The Sk–e turbulence model was then replaced by cell. However, it is also clear that in each cell a recirculation loop
the Rk–e turbulence model and a converged solution obtained once is set up, with a proportion of the flow returning, at low speed
more. The selection of the Rk– model for the mechanical floccula- along the top or bottom of the vessel. This recirculation flow pat-
tor was made on the basis of its suitability to round jets, such as tern is observed in all flocculation cells of the channel. Velocities
may be found at the flocculator inlet. To reduce computational ex- range from 1.29 ms1 along the channel floor to 0 ms1 in the
pense, just one vessel was modelled at any one time, and each of centre of cells. Dead areas are also found at the upper and lower
the three mixing speeds (3.2, 2.4 and 1.6 rpm) were simulated. corners of each cell.

Fig. 9. Plan and sectional views of meshed mechanical flocculation cell.


J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109 107

Table 3
Turbulence dissipation rate summary.

eave, 103, m2 s3 GL ave s1 GL max s1 Gave s1


Hydraulic, 160 Ml/d 4.11 45.7 734.9 46.6
1.6 rpm, 16.0 Ml/d 0.42 16.9 268.7 19.4
1.6 rpm, 21.1 Ml/d 0.48 18.2 286.5
1.6 rpm, 24.0 Ml/d 0.53 19.2 289.6
2.4 rpm, 16.0 Ml/d 1.31 29.6 544.3 31.8
2.4 rpm, 21.1 Ml/d 1.36 30.4 538.6
2.4 rpm, 24.0 Ml/d 1.45 31.6 526.0
3.2 rpm, 16.0 Ml/d 3.07 44.9 839.9 45.4
Fig. 10. Plan and elevation views of hydraulic flocculator channel. 3.2 rpm, 21.1 Ml/d 3.11 45.5 837.9
3.2 rpm, 24.0 Ml/d 3.19 46.5 832.8

to grow flocs [7]. Therefore, it is instructive to consider the range of


GL values found in that vessel with those found in the full scale
flocculators. GL values were extracted for the jar tester, mechanical
and hydraulic flocculator models and their distributions are pre-
sented in Fig. 13. At a mixing speed of 1.6 rpm, the mechanical floc-
culator vessel has a 90% coverage of GL values below 50 s1, and
95% coverage of GL values of 65 s1 or below. The three mechanical
flocculator curves are all of a similar shape, shifting towards higher
GL values as the mixing speed increases. For the 3.2 rpm simula-
tion, 90% vessel coverage is attained at a GL value of 117 s1, and
95% coverage is attained at a GL value of 170 s1. For the hydraulic
flocculator, 90% vessel coverage is attained at GL = 104 s1, and 95%
coverage is attained at GL = 130 s1 (Table 4). There is only a small
percentage of the overall channel volume where the local velocity
Fig. 11. Velocity contours z = 3.0 m, N = 2.4 rpm, Q = 21.1 Ml/d.
gradient exceeds 100 s1, with some 82% of the channel having a GL
value of 100 s1 or less.
Comparison of the hydraulic flocculator curve with the three
mechanical curves is interesting. At low velocity gradient values
(i.e. GL < 15 s1), the hydraulic velocity gradient distribution is
approximately midway between the mechanical 1.6 and 2.4 rpm
scenarios. However, beyond 15 s1, the rate of change of volume
decreases, indicating that a greater volume of higher local velocity
gradient values are found within the channel. The rate of volume
change is greater for the mechanical flocculator than for the
hydraulic one, and thus a greater volumetric change occurs over
a reduced band of GL values. The curve showing the distribution
of velocity gradients in the cylindrical jar tester when mixing at
Fig. 12. Velocity vectors along centreline of channel.
30 rpm is shifted to the left of all mechanical and hydraulic floccu-
lator curves, indicating a less turbulent mixing regime. However,
4.3.2. Turbulence dissipation rate previous work postulated that flocs formed from upland water
For the mechanical flocculator, local turbulence dissipation rate (such as those found at these two WTW) will withstand velocity
values above the vessel average were found to occur around the gradient values of up to 150 s1 and below [26]. It is apparent that
paddle edges and in the wake of the paddle, whereas for the the vast majority of the velocity gradient values within the both
hydraulic flocculator local velocity gradient values greater than the flocculation channel and the mechanical flocculator are at or
the average were generally to be found at the baffle walls tips below the postulated breakage threshold, thus indicating that the
and on the facing walls. The average and maximum local velocity magnitudes of velocity gradients induced are acceptable for ade-
gradients, GLave and GLmax, respectively, vessel average velocity gra- quate flocculation. Therefore, this work demonstrates the develop-
dient, Gave, and average turbulent eddy dissipation values, eave, ment and novel use of CFD modelling strategies and techniques to
were calculated for each mixing speed and flow rate and the results assess the performance of pilot and full scale water treatment
are shown at Table 3. For the mechanical flocculator, it is apparent plant, thus negating the need for detailed on-site assessments.
that both the average and maximum local velocity gradients are af-
fected by the mixing speed, but are effectively independent of flow 5. Use of CFD in the water industry
rate at each mixing speed. For both flocculators, the average local
velocity gradient is similar to the overall average; however, a sig- As part of this work, a number of CFD simulations of laboratory
nificant variation of GL value is noted (similar to the jar test data), and full scale unit processes commonly found at water treatment
once again demonstrating the ineffectiveness of the Gave value in works have been undertaken. The early work (MRF, two-equation
describing accurately the flow field characteristics. modelling of jar testers) was undertaken using a 1.4 GHz, 250 MB
RAM laptop computer. The most straightforward of these runs took
4.3.3. Velocity gradient at least 12 h of CPU time to obtain a converged solution and it be-
It is known that the hydrodynamic environment induced by came apparent that modelling anything more complex would re-
mixing the contents of the cylindrical vessel at 30 rpm is sufficient quire significantly enhanced computing facilities. As a result, a
108 J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109

100

Cummulative % vessel volume occupied


90

80

70

60

50

40

30

20

10

0
1 10 100 1.000
Local Velocity Gradient, s -1

1.6 rpm 2.4 rpm 3.2 rpm Hydraulic Cyl Jar Tester - 30rpm

Fig. 13. Cumulative distribution of local velocity gradient for full and lab scale flocculators.

tional commercial IT management in the water industry perceives


Table 4
Local velocity gradient values at which 90% and 95% volume coverage is attained. HPC systems to be a niche environment, and so will not be dis-
posed to significant capital investment, or the training of staff, in
Volumetric GL value at which stated coverage attained
that area. This is borne out to some extent by the fact that in the
coverage (%)
Mechanical: Mechanical: Mechanical: Hydraulic UK, it is only Anglian Water and Thames Water which are known
3.2 (rpm) 2.4 (rpm) 1.6 (rpm)
to have specialist CFD capabilities; this, from a cohort of ten major
95 170 111 65 130 water and sewerage companies. It is even less surprising to note
90 117 80 50 104
that the water-only companies have no specialist in-house CFD
expertise. This issue was also raised by Morvan [27] who stated
high performance computing (HPC) facility was used for the more that ‘‘CFD requires a background knowledge whose cost is still high
complex jar tests scenarios (SM and RSM) and all full scale simula- to industrialists for whom CFD only provides either an occasional
tions. Using the HPC, the full scale mechanical flocculator SM sim- tool or a tool that is too sophisticated to be used in its most general
ulations still required up to 12 days’ CPU usage. As a test, an form in their daily business”. Gujer [28] also questioned the likeli-
attempt was made to run the same model on a laptop. The test hood of widespread CFD modelling in the water industry and
was abandoned after two weeks of continuous running, with the wrote ‘‘. . . to what extent this ever more detailed research will lead
solution still distant from convergence. Mixing scenarios and the towards models that will be applicable in daily consulting busi-
associated turbulence model and rotating mesh treatment are tab- ness. Who can afford to identify accurately parameters that are
ulated with time to convergence at Table 3. subject to high sensitivity but are only amenable to expensive
Several articles have been published in trade magazines espous- and time-consuming analytical procedures?”
ing the benefits of CFD to the water industry and it is accepted that Although CFD offers undoubted benefits for those seeking to
the cost of undertaking simple CFD modelling has fallen in real understand better the flow in and around complex geometries,
terms as the cost of basic hardware and memory has fallen and, the level of complexity demands extensive computing power and
although annual licence fees for commercially-available codes re- trained staff to such an extent that this author believes its wide-
main relatively high, short-term licences have made the use of spread use for water industry applications outside academic and
CFD more cost-effective. However, this work has shown how mod- specialised consultancies is unlikely. Nevertheless, this work has
est computing hardware is capable of running only the most demonstrated clearly the benefits to be gained from this powerful
straightforward models. Challenges facing the modeller are, there- tool.
fore, a function of system complexity. For example, the modelling
of species concentration in a chlorine contact tank or ozone reactor 6. Conclusions
is a relatively computationally inexpensive procedure. However, as
can be seen from the work reported here, the development of tur- This paper has demonstrated the range of scales and applica-
bulence dissipation profiles in mechanically mixed vessels is rather tions to which CFD can be applied successfully to flocculation pro-
more challenging. Furthermore, even when runs are possible on cesses used for water treatment.
such machines, the machine is likely to be redundant for all other For straightforward unbaffled paddle mixing in a small vessel,
purposes during that time. This observation leads one to question the computational expense of the sliding mesh technique to model
the appropriateness of CFD in the water industry. Accepting the rotating flows was found not to be necessary where the user’s
fact that complex modelling can only be undertaken by skilled scope of interest is limited to a broad consideration and apprecia-
operatives using HPC facilities, one must consider the costs in- tion of the magnitude of shear zones and turbulent effects within a
volved. Specialist consultants and software developers do offer mixing system. The MRF approach was found to be sufficient to
CFD expertise and access to remote simulation facilities, all of develop an understanding of the magnitude of velocity gradients
which attract commercial rates. This author believes that tradi- to which a floc may be subjected within a vessel.
J. Bridgeman et al. / Advances in Engineering Software 41 (2010) 99–109 109

CFD analysis has demonstrated that, when the jar test vessel [4] Korpijarvi J, Laine E, Ahlstedt H. Using CFD in the study of mixing in
coagulation and flocculation, in chemical water and wastewater treatment VI.
contents were mixed at the same speed, the turbulence dissipation
In: Hahn HH, Hoffmann E, Odegaard H, editors. Proceedings of the ninth
rates generated in a square section vessel were significantly higher Gothenburg symposium, Springer; 2000. p. 89–99.
than those found in a circular section vessel, albeit they were less [5] Essemiani K, de Traversay C. Optimisation of the flocculation process using
well-distributed. computational fluid dynamics, in chemical water and wastewater treatment
VII. In: Hahn HH, Hoffmann E, Odegaard H, editors. Proceedings of the 10th
For the full scale mechanical flocculator, changes to the flow Gothenburg symposium, IWA Publishing; 2002.
rate were shown to exert only a small influence on the distribution [6] Essemiani K, de Traversay C. Optimum design of coagulation/flocculation
of local velocity gradient values. Changes to the mixing speed did vessels, WQTC 2002 conference, Seattle, USA; 2002.
[7] Jarvis P, Jefferson B, Parsons SA. The duplicity of floc strength. Water Sci
impose a significant impact on the local velocity gradient Technol 2004;50:63–70.
distribution. [8] Gregory J. The density of particle aggregates. Water Sci Technol
For the hydraulic flocculator, large recirculation loops and dead 1997;36(4):1–13.
[9] Biggs CA, Lant PA. Activated sludge flocculation: on-line determination of floc
spots were identified in flocculation cells between baffles. size and the effect of shear. Water Res 2000;34:2542–50.
For both the mechanical and hydraulic flocculators, the distri- [10] Coufort C, Bouyer D, Liné A. Flocculation related to local hydrodynamics in a
bution of turbulence dissipation rate was found to be broadly sim- Taylor–Couette reactor and in a jar. Chem Eng Sci 2005;60:2179–92.
[11] Yeung AKC, Pelton R. Micromechanics: a new approach to studying the
ilar to that found in the cylindrical jar tester at 30 rpm and so strength and breakup of flocs. J Colloid Interface Sci 1996;184:579–85.
conducive to floc growth. [12] Thomas DG. Turbulent disruption of flocs in small particle size suspensions. J
Using modest computing hardware (e.g. desktop personal com- AIChE 1964;19(4):517–23.
[13] Camp TR, Stein PC. Velocity gradients and internal work in fluid motion. J
puters) CFD modelling is restricted to the analysis of relatively sim-
Boston Soc Civ Eng 1943;30:219–37.
ple flows. More complex scenarios, such as rotating meshes, [14] Luo C. Distribution of velocities and velocity gradients in mixing and
require greater computing power which is not generally found out- flocculation vessels: comparison between LDV data and CFD predictions.
side academic environments or specialised consultancies. CFD PhD thesis, New Jersey Institute of Technology; 1997.
[15] Stanley SJ, Smith DW. Measurement of turbulent flow in standard jar test
modelling also requires trained staff to generate robust models apparatus. J Environ Eng 1995(December):902–10.
and solutions. The combination of these two factors is likely to pre- [16] Bridgeman J. Investigating the relationship between shear and floc fate using
clude the widespread use of CFD in most water companies and CFD. PhD thesis, Cranfield University; 2006.
[17] Haarhoff J. Design of around-the-end hydraulic flocculators. J Water Supply:
mainstream consultancies in the foreseeable future. However, this Res Technol AQUA 1998;47(3):142–52.
work has demonstrated clearly the benefits to be gained from this [18] Haarhoff J, van der Walt JJ. Towards optimal design parameters for around-
powerful tool. the-end hydraulic flocculators. J Water Supply: Res Technol AQUA 2001;
50(3):149–59.
[19] Prat OP, Ducoste JJ. Simulation of flocculation in stirred vessels – Lagrangian
Acknowledgements versus Eulerian. Chem Eng Res Des 2007;85(A2):207–19.
[20] Nopens I. Improved prediction of effluent suspended solids in clarifiers
through integration of a population balance model and a CFD model. In:
The submitted manuscript has been made possible through the Proceedings of PS-IWA 2007: particle separation, Toulouse; July 2007.
funding from the American Water Works Association Research [21] Feng X, Li XY. Modelling the kinetics of aggregate breakage using improved
Foundation and Co-funding Utilities. The authors would also like breakage kernel. Water Sci Technol 2008;57(1):151–7.
[22] Versteeg HK, Malalasekera W. An introduction to computational fluid
to thank the Engineering and Physical Sciences Research Council
dynamics. The finite volume method. Prentice Hall; 1995.
(EPSRC), Yorkshire Water, Thames Water, United Utilities, Severn [23] Roache PJ. Verification of codes and calculations. AIAA J 1998;36(5):
Trent Water, Scottish Water and Fort Collins Utilities for their 696–702.
financial support. [24] Armenante PM, Luo C, Chou C-C, Fort I, Medek J. Velocity profiles in a closed,
unbaffled vessel: comparison between experimental LDV data and numerical
CFD predictions. Chem Eng Sci 1997;52(20):3483–92.
References [25] Yeung H. Modelling of service reservoirs. J Hydroinformatics 2001;3(3):
165–72.
[1] Boller S, Blaser S. Particles under stress. Water Sci Technol 1998;37(10):9–29. [26] Bridgeman J, Jefferson B, Parsons SA. Assessing floc strength using CFD to
[2] Jarvis P, Jefferson B, Gregory J, Parsons SA. A review of floc strength and improve organics removal. Chem Eng Res Des 2008;86(8):941–50.
breakage. Water Res 2005;39:3121–37. [27] Morvan HP. Automating CFD for non-experts. J Hydroinformatics 2005;7(1):
[3] Korpijarvi J, Ahlstedt H, Saarenrinne P, Renanen J. Modelling of flow field in the 17–29.
mini-flocculator. In: Benkreira H, editor. IChemE symposium series no. 146, [28] Gujer W. Activated sludge modelling: past, present and future. Water Sci
fluid mixing 6; 1999. p. 361–72. Technol 2006;53(3):111–9.

You might also like