Lemoullec 2008

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Chemical Engineering Science 63 (2008) 2436 – 2449

www.elsevier.com/locate/ces

Flow field and residence time distribution simulation of a cross-flow


gas–liquid wastewater treatment reactor using CFD
Yann Le Moullec, Olivier Potier, Caroline Gentric ∗ , Jean Pierre Leclerc
Laboratoire des Sciences du Génie Chimique, UPR 6811, CNRS-ENSIC-INPL, 1, rue Grandville BP 20451, 54001 Nancy, France

Received 16 January 2007; received in revised form 22 January 2008; accepted 24 January 2008
Available online 6 February 2008

Abstract
A three-dimensional Eulerian–Eulerian two-phase approach has been used for the simulation of a cross-flow gas–liquid wastewater treatment
reactor. Two different turbulence models have been tested: the k. and Reynolds Stress Model (RSM) models. Bubble induced turbulence
source terms have been added to these models. Numerical results have been validated using Laser Doppler Velocimetry (LDV) measurements.
Simulations with both turbulence models successfully predicted the hydrodynamics of the reactor. Then particle tracking with a stochastic
approach has been used to calculate residence time distributions (RTD) with the flow previously simulated. It has been shown that dispersion
in the reactor is primarily due to turbulence. Results have been compared with experimental RTD for various liquid and gas flowrates both on
a bench scale and full scale plant. The RSM model accurately predicted the dispersion whereas the standard k. model slightly underestimated
the dispersion.
䉷 2008 Elsevier Ltd. All rights reserved.

Keywords: Dispersion; Multiphase flow; RTD; Simulation; Turbulence; Wastewater treatment

1. Introduction reactors have been studied more extensively. In particular


bubble column reactors have been extensively investigated
The pollution removal by microorganisms is an essential either with the Euler–Euler approach (Jakobsen et al., 2005;
step in the biodegradable wastewater treatment plants. These Pfleger and Becker, 2001; Sokolichin et al., 2004) or with
biological reactions often take place in gas/liquid reactors the Euler–Lagrange approach (Lain et al., 2002; Lapin and
with a very long length compared to their height and width Lübbert, 1994). Two- or three-dimensional (2D and 3D)
(channel reactors). In this type of reactors, the global water flow and stationary or transient simulations have been performed.
is along the length of the reactor and gas is sparged at the bot- Although transient 3D simulations are necessary to represent
tom. In wastewater treatment reactors, apparent reaction orders the complex flow structure of bubble columns, 2D and station-
are often greater than zero; therefore the efficiency of the pol- ary simulations allow correct representations of the average
lution removal reaction depends on the hydrodynamics (Levin flow field and can be sufficient in certain cases (Sokolichin
and Gealt, 1993). Therefore, in order to accurately predict et al., 2004). But simulation of bubbly flows using CFD is not
conversions, the kinetics modelling must be coupled with the straightforward. Closure terms, in particular interaction forces
hydrodynamics description. But the prediction of bubbly flow between phases and their impact on the simulation results
hydrodynamics is quite difficult due to the complex phenomena have been the subject of numerous studies (Jakobsen et al.,
involved. 1997). Another critical point widely discussed, is the turbu-
Nevertheless, with the improvement of computational lence closure terms. Mainly three modifications of single phase
power and the availability of computational fluid dynamics turbulent viscosity models have been proposed and compared
(CFD) codes adapted to such two-phase flows, gas–liquid in the literature: the addition of a bubble induced turbulence
viscosity term (Sato and Sekoguchi, 1975), a pseudo-bubble
∗ Corresponding author. Tel.: +33 3 83 17 53 38; fax: +33 3 83 32 29 75. induced turbulence tensor (Arnold, 1988) and the addition of
E-mail address: Caroline.Gentric@ensic.inpl-nancy.fr (C. Gentric). bubble induced turbulence source terms in the k. transport
0009-2509/$ - see front matter 䉷 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2008.01.029
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2437

equations (Gu and Guo, 2005; Mudde and Simonin, 1999; (Plexiglas) to allow LDV measurements. Its total length is 3.6 m
Olmos et al., 2003; Pfleger and Becker, 2001). It seems that the with a rectangular section of width and height, respectively,
addition of bubble induced turbulence source terms is the best equal to 0.18 and 0.2 m (Fig. 1). One side of the walls of the
compromise between precision and complexity (Sokolichin reactor is fitted with stainless-steel tubes where 1 mm holes
et al., 2004). Recently, efforts have been made to simulate local have been drilled every centimetre for air sparging. Gas and
and global mass transfer in bubble columns, air-lift and trickle liquid flowrates are the variable parameters.
bed reactors (Ekambara et al., 2005; Gunjal et al., 2003; Vial
et al., 2005). CFD simulations of liquid phase residence time 2.2. LDV measurements
distribution (RTD) in these gas–liquid reactors have also been 2D LDV measurements have been performed to obtain ax-
performed with good results (Ekambara et al., 2005; Gunjal ial (Ux ), lateral (Uy ) and vertical (Uz ) time-averaged velocity
et al., 2003). With these improvements in gas–liquid simula- fields (see Fig. 1 for directions) as well as an estimated value
tions, some other industrial reactors can be investigated and of the turbulent kinetic energy (k). An acquisition time of 180 s
specifically wastewater treatment reactors. The previously pre- is required in order to obtain statistically meaningful results for
sented studies are limited to vertical liquid flow with counter or all measurement positions.
co-current gas flow. Wastewater treatment reactors by activated Photographic technique has shown that the gas hold-up along
sludge are mostly cross-flow reactors in which liquid flow is the reactor length is inhomogeneous though the variations are
horizontal and gas flow is vertical. Two types of reactors are small. In order to verify that these small variations of gas hold-
used: closed-loop oxidation ditch reactors, also called carrousel up are of no importance for the LDV measurements, these
reactors, and channel aerators. In carrousel reactors, the liquid measurements have been carried out at two different planes
horizontal velocity is such that gas has no real impact on the with different apparent gas hold-ups (planes 1 and 2: one is
hydrodynamics and mixing is sometimes performed with im- located between two sparger tubes, the other is located at the
pellers (Roustan and Line, 1996). CFD modelling of the veloc- middle of a sparging device, see Fig. 1).
ity field and mass transfer has been carried out for these reactors
(Cockx et al., 2001). In channel aerators, the liquid horizontal 2.3. RTD experiments
velocity is very low compared to carousel reactors and mixing
is performed by gas injection in the reactor. Therefore an accu- Tracer experiments used for simulation validation have been
rate simulation of the interaction between bubbles and liquid is performed in a previous study in the same bench scale reactor
of primary importance. The purpose of this work is to obtain an (Potier et al., 2005). Approximately 80 RTD experiments have
accurate and predictive hydrodynamics model of a channel aer- been carried out by injecting pulses of a solution of sodium
ator with respect to velocity, turbulence and RTD which is, as chloride (NaCl) at the inlet and monitoring its concentration
in most gas–liquid flows, governed by local scale phenomena with a conductimetric probe at the outlet. A plug flow with
(Vial et al., 2005). Validation is performed with Laser Doppler axial dispersion model (Villermaux, 1993) allows a very pre-
Velocimetry (LDV) measurements and RTD experiments. cise representation of the data. Mean residence times and axial
dispersion coefficients have been determined by model fitting
with the software DTSPRO 4.2 (PROGEPI, France). This pre-
2. Experimental setup and preliminary results
vious study has shown the existence of a correlation between
2.1. Bench scale reactor the axial dispersion coefficient on the one hand and the gas
flowrate and the geometry of the reactor on the second hand:
Experiments have been carried out in a bench scale channel 
Qg −1.99
reactor filled with tap water. It is built in transparent material D = (0.2032.w − 0.008569) (100h)0.00476.w (1)
L

Fig. 1. Schematic representation of the bench scale reactor.


2438 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

with h, w and L, respectively the height, width and length Table 1


of the reactor and Qg the gas flowrate. This correlation has Constants used in the two-phase k . model

been established for Qg (15.65 L min−1 ), Ql (1.5.5 L min−1 ), C C1 C2 Ck C k 


h (0.08.0.2 m) and w (0.05.0.2 m). Besides, this correlation
0.09 1.44 1.92 1.44 1.0 1.0 1.3
has been validated on an industrial reactor successfully. In the
present work, results from simulation and experiments have
been compared for different gas and liquid flowrates for a given
geometry: length of 3.6 m, height of 0.2 m and width of 0.18 m. The drag coefficient is calculated from the following equation
(Jamialahmadi et al., 1994):
3. Hydrodynamics modelling
4 (l − g ) gd g
Cd = with
Because of the high number of bubbles (approximately 7000 3 l u2∞
and more in an industrial size reactor) in the reactor and despite 
2l gd g
the low gas fraction (< 10%), an Euler–Euler approach has u∞ = + for dg > 3 mm. (9)
been used. CFD simulations have been carried out with the dg (l + g ) 2
CFD software FLUENT.
3.2. Two-phase k. turbulence model
3.1. Euler–Euler equations
In order to model turbulence, a two-phase k. turbulence
The mass transfer between phases has been neglected there- model can be used (Elghobashi and Abou-Arab, 1982; Mudde
fore the continuity equation is expressed by and Simonin, 1999). Only the continuous phase turbulence is
taken into account and is considered isotropic. The dispersed
j
(q q ) + ∇ · (q q Uq ) = 0 with q = g or l (2) phase influences the continuous phase through bubble induced
jt turbulence terms. For the liquid phase, the transport equations
besides: for the turbulent kinetic energy k and the turbulent dissipation
rate  are
g + l = 1. (3)    
j t,l
The momentum conservation equation for multiphase flows is (l l kl ) + ∇ · (l l Ul kl ) − ∇ · l l + ∇kl
jt k
given by
j = l (t,l ∇Ul [∇Ul + (∇Ul )T ] − l l ) + k,l , (10)
(q q Uq ) + ∇ · (q q Uq Uq )    
jt j t,l
(l l l ) + ∇ · (l l Ul l ) − ∇ · l l + ∇l
= −q ∇p + q q g + ∇ · q (q + t,q ) + Iq jt 
l
with q = g or l, (4) = l (C1 t,l ∇Ul [∇Ul + (∇Ul )T ] − C2 l l ) + ,l .
kl
where Iq is the interphase momentum exchange term and q and (11)
t,q are, respectively, the viscous stress tensor and the turbulent The turbulent viscosity of the continuous phase is calculated
stress tensor, defined as follows: by Eq. (7).
The bubble induced turbulence source terms are calculated
q = q (∇Uq + ∇Uqt ) + (q − 23 q )∇.Uq I (5)
from (Pfleger and Becker, 2001)
and
k,l = l Ck |Il | · |Ug − Ul |, (12)
t,q = t,q (∇Uq + ∇Uqt ) − 23 (kq + t,q ∇.Uq )I (6)
l
,l = l C1 C |Il | · |Ug − Ul |. (13)
with kl
kq2 All the constants of the two-phase k. turbulence model are
t,q = q C . (7)
q summarized in Table 1.
In most gas–liquid flows virtual mass force is necessary to
3.3. Two-phase RSM turbulence model
reproduce transient flow characteristics (Leon-Becerril et al.,
2002), nevertheless this force has no influence on stationary
The Reynolds Stress Model (RSM) (Daly and Harlow, 1970)
calculation results (Sokolichin et al., 2004; Zhao et al., 2006)
does not assume the isotropy of turbulence and is more adapted
therefore it has been neglected and only the drag force has been
to swirling flows than the k. turbulence model and its variants
considered significant. Iq is thus defined as follows:
(Ranade, 2002). In this model, the required computational time
3 g l is only 60% higher than with a k. model. As for the k.
Il = Cd |Ug − Ul |(Ug − Ul ) and Ig = −Il . (8)
4 dg model, turbulence is only considered for the continuous phase.
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2439

Table 2 Table 3
Constants used in the two-phase RSM model Constants used in the stochastic model

C C 1 C2 Ck C  CL k . model CL RSM model

0.09 1.44 1.92 1.44 1.92 1.0 0.15 0.3

Bubble induced turbulence source terms are added to the trans- turbulent dispersion of particles is also taken into account: par-
port equations. The transport equation for the Reynolds stress ticles in a turbulent flow are submitted to a randomly varying
tensor component Rij is flow field. This phenomenon is modelled by a stochastic ap-
proach, assuming that the particles interact with a sequence of
j j
(l l Rij ) + (l l Uk Rij ) turbulent eddies, the fluctuating velocity within each eddy be-
jt jxk ing isotropic and obeying a Gaussian probability density func-
    tion. The interaction time I is assumed to be sufficiently short
jUj jUi j jRij
= −l l Rik + Rj k + l l so that the fluid velocity in a given eddy is constant during
jxk jxk jxk jxk
this process. The particle trajectory is therefore calculated as
 
j jui juj follows:
−   
(l l ui uj uk ) + l p +
jxk jxj jxi dup 3 
= Cd l |up −ul |(up −ul ) (19)
dt 4 dp p
− l l ij + R,ij (14)
with
with 
2
R,ij = 3
ij k,l
2
with k,l from Eq. (12) (15) ul =Ul +ul and ul composed of ux,l =uy,l =uz,l = kl .
3
and (20)

ij = 23
ij l with l from Eq. (17). (16) is a normally distributed random number and Cd is calculated
by the Morsi and Alexander correlation (1972).
The third and fourth term on the right-hand side of Eq. (14) are A new value of is applied each time the interaction time is
modeled (see Fluent Inc., 2005). reached. This interaction time is taken as the minimum of the
The turbulence dissipation rate l is calculated using the eddy lifetime e and the eddy crossing time ct .
following transport equation:
    kl
j j j t,l j e = −CL log( ), (21)
(l l l ) + (l l Ui l ) − l  l + l l
jt jxi jxj  jxj
    where is a uniform random number between 0 and 1.
l jui jui   
=−l C1 l Rik +Rik +C2 l l + ,l (17) Le
kl jxk jxk ct = − ln 1 − (22)
|up − ul |
with
with the particle relaxation time:
kl = 21 (Rii + Rjj + Rkk ) (18)
4 2p dg
and ,l is given by Eq. (13) and t,l by Eq. (7). =
3 2l Cd |up − ul |
All the constants of the two-phase RSM turbulence model
are summarized in Table 2. and the eddy lifetime:
 3/2
3.4. Lagrangian particle motion model 3 kl
Le = C .
2 l
In order to perform the simulation of RTD, particle track-
ing has been used (Stropky et al., 2007; Thyn et al., 1998). Different values for CL are recommended depending on the
RTD is deduced from the simulation of a pulse injection of a turbulence model (Fluent Inc., 2005) (Table 3).
sufficient number of particles. The trajectory of a particle is
predicted by integrating the force balance on the particle in a 3.5. Transport equation of a tracer
Lagrangian reference frame (Fluent Inc., 2005). The character-
istics of these particles have been chosen such as they can be Another approach used to simulate RTD is the solution of a
considered as perfect tracers of the fluid flow. Their diameter is transport equation for a passive tracer with the same physical
10−6 m and they have the same density as water. This ensures properties as the continuous phase (Talvy et al., 2007; Zhang
that they can perfectly follow all the simulated timescales. The et al., 2007). The transport equation for the concentration of a
2440 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

tracer in a turbulent flow is Therefore a refined mesh is necessary near that wall. Differ-
   ent grid sizes have been tested: cubic cells of 1 cm3 (130 000
j 
(l Ctr ) + ∇.(l Ul Ctr ) = ∇. l Dm + t ∇Ctr (23) cells) or 0.125 cm3 (1 000 000 cells) and cells of 1 cm3 with a
jt Sct refined mesh near the walls (350 000 cells) (Fig. 2). Grid in-
with Sct , the turbulent Schmidt number considered constant and dependency has been verified (Fig. 3) and finally this last grid
equal to 0.7. The simulation of the RTD is obtained following offers the best compromise between precision and computa-
the outlet concentration of the tracer injected as a pulse at the tional effort.
inlet. Inlets of gas and liquid have been modelled with a velocity
inlet boundary condition. The turbulence boundary conditions
at the inlets are given through the turbulence intensity (10%)
3.6. Grid, boundary conditions and discretization scheme
and inlet hydraulic diameter. The outlet of liquid has also been
modelled with a velocity inlet boundary condition; it allowed
It is commonly admitted in bubble induced flows that there
us to impose the outflow velocity in order to avoid backflow
is no need for boundary layer or advanced wall treatment
which causes mass balance problem when RTD simulations are
(Troshko, 2006). In our case, bubbles are only located near one
performed. In order to simulate the degassing boundary condi-
wall which induces a high velocity gradient at this location.
tion at the top of the reactor, an outlet boundary condition has
been used for the gas phase and a symmetry boundary condi-
tion has been used for the liquid phase, this degassing condition
has been implemented through a user defined function.
A second order discretization scheme (QUICK: Quadrative
Upwind Interpolation for Convective Kinematics) has been cho-
sen for the turbulent dissipation rate and void fraction equations
in order to limit their numerical diffusion (calculation errors
induced by grids and discretization schemes). A first order
discretization scheme has been tested for the solution of the mo-
mentum equations but it leads to unstable results; consequently
a second order discretization scheme (QUICK) has been used
to solve all equations. The SIMPLE (Semi-Implicit Method for
Pressure-Linked Equations) pressure–velocity coupling scheme
has been used. The under relaxation parameters used are given
in Table 4.
Table 4
Under relaxation parameter used for each variable

Pressure Momentum Volume fraction k  Rij

0.3 0.7 0.2 0.8 0.8 0.5


Fig. 2. Projection of the grid on a vertical plane (y, z).

0.4
Time-averaged liquid vertical velocity Uz (m/s)

0.3

0.2

0.1

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
-0.1

-0.2 129600 cells grid


345600 cells grid
-0.3 1036800 cells grid

-0.4
y (m)

Fig. 3. Simulated average liquid vertical velocity on a transversal line (x = 1.8 m, z = 0.1 m) for three different grid sizes.
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2441

4. Results

This CFD study will be divided into three main parts. First,
a simulation of the complete turbulent hydrodynamics for dif-
ferent liquid and gas flowrates is carried out. The second step
is the simulation of the RTD. And finally, the numerical param-
eters used for the bench scale reactor are validated on a plant
scale reactor.
A test case experiment is chosen as a reference for the vali-
dation of the simulations. It is carried out for a liquid flowrate
QL of 3.6 L min−1 and a gas flowrate QG of 15 L min−1 . The
bubble diameter for this air–water system has been estimated
using a photographic technique: an average value of 4 mm has
been used and since the gas hold-up is very low (< 10%),
bubble break-up and coalescence are neglected.

4.1. Velocity profiles and turbulence characteristics

Simulations have been performed with the k. and RSM tur-
bulence models. Fig. 4 presents the projection of the liquid ve-
locity field on a vertical plane for the experimental data and Fig. 4. Overall representation of the experimental and simulated average
velocity fields on a vertical plane (y, z) for both turbulence models.
the results of simulations with the k. and RSM turbulence
models. Simulated velocity fields with the k. and RSM mod-
els are superimposed, both models giving similar results. The
may have an impact on dispersion; this will be investigated in
overall agreement is good, however, some discrepancies can be
Section 4.2.
observed near the bubble injection position and near the free
Table 5 presents results relative to turbulence. The experi-
surface. Discrepancies observed near the sparger are probably
mental results for turbulent kinetic energy k are surface aver-
due to the simplification made for the gas inlet boundary con-
ages on the measurement plane (plane 1). It can be highlighted
ditions (velocity magnitude of 0.162 m s−1 and void fraction
that experiments have shown an isotropic turbulence. The tur-
of 1) whereas those observed near the free surface are probably
bulent dissipation rate  of the liquid phase can be estimated
due to the simplified representation of the surface which is not
with an energy balance (Eq. (24)), assuming that all the energy
planar in the experimental setup.
brought by the isothermal expansion of the gas is dissipated by
Fig. 5 compares the measured and simulated profiles of
turbulence:
the vertical liquid velocity at three different heights. These  
quantitative comparisons show some reasonable agreement be- Patm QG  gh
l theoretical = ln 1 + l . (24)
tween experiments and simulations (an average difference of l V Patm
0.035 m s−1 is observed which represent 18% of the maximum
liquid velocity) whereas the two simulations results are very Simulation results for the turbulent kinetic energy k are also
close. The LDV measurements on the two measurement planes surface averages on plane 1 and turbulent dissipation rate  is
(defined in Section 2.1) show that there is no significant differ- a volume average over the whole reactor. The turbulent kinetic
ence, with respect to the vertical liquid velocity, between these energy k is estimated by simulation with an underestimation
two positions despite their apparent different gas hold-ups (ex- of around 26% by the k. model and a 33% overestimation
perimental data planes 1 and 2). As it will be discussed later, by the RSM model. The turbulent dissipation rate  obtained
despite the discrepancies between experimental and simulated with the k. model is also underestimated by 20% whereas the
results, the liquid velocity fields calculated are satisfying for turbulent dissipation rate  obtained with the RSM model is
RTD calculations. more importantly underestimated by 25%. The underestimation
LDV measurements revealed a non-uniform time-averaged of  could be explained by the energy dissipation at the free
axial velocity profile along the length of the reactor (Fig. 6). surface due to bubble break-up. This dissipation can reach 30%
Simulations predict these spatial oscillations (Fig. 6) but fail of the total energy for some configurations (Desai et al., 1995).
to predict the period and amplitude of the phenomenon. The
maximum experimental value of these oscillations (0.02 m s−1 ) 4.2. Residence time distribution
is ten times the mean axial velocity (0.0017 m s−1 ). Because
of these very small velocity values, measurement uncertainty All our experimental RTD can be described by the plug
is high. But, at the same time, the maximum axial velocity flow with axial dispersion model. Therefore the dispersion
is still ten times smaller than the maximum vertical veloc- coefficient is a good estimation of the overall mixing within
ity (0.2 m s−1 ). The amplitude of these fluctuations is thus the reactor. The simulated dispersion coefficient can be
negligible compared to the highest velocities encountered but described as the sum of four contributions, three of them are
2442 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

Fig. 5. Comparison between experimental and simulated average vertical velocities at three different heights: 5, 10 and 15 cm.

0.025

0.02 Experimental data


RSM simulation
0.015
k-eps simulation
0.01

0.005
Ux (m/s)

-0.005

-0.01

-0.015

-0.02

-0.025
0 5 10 15 20 25 30 35 40 45 50
x (cm)

Fig. 6. Comparison between experimental and simulated average axial velocities along a fraction of the length of the reactor (at y = 4 cm, z = 10 cm).
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2443

physical: molecular diffusion, dispersion due to velocity field 4.2.1. RTD simulation with scalar transport
(convection) and turbulent dispersion, and the fourth term is An integration time step of 4 s (0.2% of mean residence time)
the numerical diffusion (Eq. (25)). has been selected. Fig. 7 presents the RTD obtained with the
RSM turbulence model and the experimental one. Simulation
Dtotal = Dmolecular + Dconvection + Dturbulent + Dnumerical . (25) results obtained with the k. turbulence model are not pre-
sented because they are similar to the ones obtained from RSM
There are two main methods to simulate a RTD with CFD:
simulations.
• Particle tracking: the motion equations for numerous parti- The mean residence time is correctly simulated whereas the
cles are solved from the entrance of each particle in the reac- dispersion coefficient is 30% underestimated. The simulated
tor to its escape from the reactor (Section 3.4). A significant dispersion coefficient is 4.26 × 10−4 m2 s−1 whereas the exper-
statistical RTD can be obtained as long as enough particles imental one is 6.01 × 10−4 m2 s−1 .
trajectories are calculated (approximately 4000 in our case).
Using this method avoids solving a transport equation and, 4.2.2. Influence of the different dispersion sources on the
therefore, eliminates numerical diffusion because numerical RTD simulated via tracer transport
diffusion is due to the discretization of a transport equation Talvy et al. (2007) proposed a method to identify the part
(numerical schemes and mesh size). However, it does not al- of the two different aspects of the dispersion: Dconvection and
low to take the molecular diffusion of a tracer into account. Dturbulent (the two other aspects are negligible in our case:
This has no consequence here since the molecular diffusion Dmolecular = 1.48 × 10−10 m2 s−1 and an estimation of the
of our experimental tracer is negligible compared to the dis- numerical dispersion gives Dnumerical = 5.0 × 10−6 m2 s−1 ).
persion coefficient. This will be verified later. Eq. (26) proposes an estimation of the turbulent dispersion
• Tracer transport: the transport equation for a passive tracer is (Talvy et al., 2007):
solved (Section 3.4). A pulse of tracer is injected at the inlet t
and its concentration is followed at the water outlet. This Dturbulent = . (26)
l Sct
method includes theoretically all the dispersion sources. Its
implementation can be performed in two steps in order to In our case this turbulent dispersion is equal to 3.81 ×
reduce the calculation time: first the solution of the stationary 10−4 m2 s−1 , approximately 90% of the total dispersion.
flow field and second the solution of the transient transport The convective dispersion can be estimated from CFD
equation for the tracer. simulation with Eq. (27) (Talvy et al., 2007).

Both methods have been tested in this work. ul,x Ctr y,z − ul,x y,z Ctr y,z
Dconvection = . (27)
jCtr y,z
Table 5
Comparison between experimental and simulated turbulence variables jx z

Experimental/ k . RSM This equation estimates the convective dispersion to 1.14 ×


theoretical simulation simulation 10−5 m2 s−1 , approximately 2% of the total dispersion. These
results indicate that the convective dispersion is negligible
k (10−3 m2 s−2 ) 2.42 (exp) 1.78 3.31
compared to the turbulent dispersion. This conclusion will be
 (10−3 m2 s−3 ) 8.75 (Eq. (25)) 6.84 6.10
discussed later. Thus the main dispersion source is turbulence.

0.0006
Experimental RTD

0.0005 Simulated RTD realised with 100 s


sample count (particles tracking)
Simulated RTD realised with 200 s
0.0004 sample count (particles tracking)
Simulated RTD fitted with DTS PRO
(particles tracking)
E (t)

0.0003 Simulated RTD with RSM model


(transport equation)

0.0002

0.0001

0.0000
0 1000 2000 3000 4000 5000 6000
Time (s)

Fig. 7. Comparison between experimental and simulated RTD obtained with the RSM turbulence model and the transport equation method or the RSM
turbulence model and the particle tracking method for different sample times for a liquid flowrate of 3.6 L min−1 and a gas flowrate of 15 L min−1 .
2444 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

Consequently the turbulent Schmidt number allows to adjust In order to verify the negligibility of the molecular diffusion,
the dispersion coefficient. It has been verified that a turbulent two simulations have been performed: two RTD simulations
Schmidt number of 0.44 would give a perfect match between with tracer transport method, the first one with a molecular
experimental and simulated RTD. diffusion coefficient of 10−20 m2 s−1 (in order to simulate a
non-diffusive tracer) and the second one with a molecular dif-
Table 6
fusion coefficient of 1.48 × 10−10 m2 s−1 (NaCl in pure water
Numerical experiments: Comparison of the influence of the main parameters
on the simulated dispersion coefficient with the k . turbulence model at 20 ◦ C). The results of these two experiments are presented
in Table 6 (Ref 6 and 7) and they show that the molecular dif-
Simulation method Ref Characteristics D (10−4 m2 s−1 ) fusion has no significant impact on the overall dispersion.
Particle tracking 1 Reference 2.51
2 2.k 5.02 4.2.3. RTD simulation with particle tracking
3 2. 3.18 Experimental molecular diffusion is negligible compared to
4 2.uy , 2.uz 3.16 the turbulent and convective dispersions therefore particle track-
5 jux /jxi = 0 2.37
ing can also be used to estimate RTD. Because of the discrete
Transport of 6 Dm = 10−20 m2 s−1 3.34 nature of particle tracking, particles must be sampled in order
passive tracer to exploit data: particles escaping the reactor during the same
7 Dm = 14.8 10−10 m2 s−1 3.35 time period are counted; RTD is obtained from the curve giving

0.0008
Experimental RTD
0.0007
Simulated RTD with the RSM
model fitted with DTS PRO
0.0006
Simulated RTD with the k-eps
model fitted with DTS PRO
0.0005
E (t)

0.0004

0.0003

0.0002

0.0001

0
0 1000 2000 3000 4000 5000 6000
Time (s)

Fig. 8. Comparison between experimental and simulated RTD obtained with the RSM and the k . turbulence models and the particle tracking method for a
liquid flowrate of 3.6 L min−1 and a gas flowrate of 15 L min−1 .

12 45
40
10
35
Residence time (min)

8 30
D (10-4 m2/s)

25
6
20
4 15
Simulated dispersion coefficient
Experimental dispersion coefficient 10
2 Correlated dispersion coefficient (Potier 2005)
Simulated residence time 5
Experimental residence time
0 0
0 10 20 30 40 50 60 70
Gas flowrate (l/min)

Fig. 9. Experimental and simulated dispersion coefficients and mean residence times for different gas flowrates with a constant liquid flowrate of 3.6 L min−1
with the RSM model and the particle tracking method.
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2445

20

Dispersion coeffient (10-4 m2/s2)


18 50

Mean residence time (min)


16
14 40

12
30
10
8
20
6 Simulated dispersion coefficient
4 Experimental dispersion coefficient 10
Simulated residence time
2 Experimental residence time
0 0
2 2.5 3 3.5 4 4.5
Liquid flowrate (l/min)

Fig. 10. Experimental and simulated dispersion coefficients and mean residence times for different liquid flowrates with a constant gas flowrate of 65 L min−1
with the RSM model and the particle tracking method.

the number of particles counted during one period as a func- the dispersion coefficient even if it does not vary much with
tion of time. Fig. 7 shows the curves obtained with the RSM the liquid flowrate in this domain. The mean residence time is
turbulence model for two different sampling times: 100 and also correctly simulated.
200 s and an experimental curve fitted with the plug flow with
closed–closed axial dispersion model. As the sampling time in- 4.2.4. Influence of the different dispersion sources on the
creases, fluctuations decrease because of an integration effect. RTD simulated via particle tracking
The mean residence time and dispersion coefficient D are de- Simulated RTD with transport equation solution shows that a
termined using the raw data: first, the mean residence time is very high proportion of the axial dispersion is due to turbulent
determined and in a second step the dispersion coefficient is op- dispersion (90%). But the solution of the transport equation fails
timized in order to fit a plug flow model with axial dispersion. to accurately simulate experimental RTD. On the other hand,
Comparison between experimental and simulated RTD is particle tracking allows a better RTD estimation, but the deter-
shown in Fig. 8 for the k. and RSM turbulence models. The mination of the different dispersion sources is not as straightfor-
RSM turbulence model simulation leads to very good results ward. With CFD it is possible to investigate the influence of one
and the k. turbulence model leads to an underestimated value of the dispersion source at once, with the other parameters kept
of the dispersion coefficient of around 50%. These results are constant. This is impossible with physical experiments. These
due to the stochastic model and especially the constant CL “numerical experiments” do not have real physical meaning but
(Table 3). In fact, in the simulation of turbulent dispersion, they allow a better understanding of the physical phenomena
there is a direct linear relation between CL and the disper- responsible for the dispersion in the reactor. These numerical
sion coefficient D. Consequently the value of CL can be opti- experiments have been performed based on our reference case
mized to obtain a good dispersion coefficient. Empirical value (Sections 4.1 and 4.2).
of CL = 3C = 0.27, which is close to the one used for the Our main objective, here, is to determine which of turbu-
RSM model (Table 3), can be found in literature for the k. lence, convection or molecular diffusion is responsible for the
turbulence (CD-Adapco-Group, 2003); with this value of CL observed dispersion. All the following “numerical experiments”
both the k. and RSM turbulence models give the same results are performed with the k. turbulence model in order to avoid
(not shown here). too long calculations.
Additional experiments and simulations have been performed To study the impact of the velocity field, two simulations
for different gas and liquid flowrates and similar good agree- have been performed: the first one with an axial velocity set
ment between experimental and simulated RTD with the parti- to a constant value of 1.667 mm s−1 (i.e. the average value
cle tracking method have been obtained. All simulations were of the axial liquid velocity) in all grid cells (jux /jxi = 0)
carried out with the RSM turbulence model. Comparison be- in order to evaluate the impact of axial velocity oscillations,
tween experimental and simulated dispersion coefficients and the second one with the velocity of the gas induced loop
mean residence times for different gas flowrates are presented doubled (see Fig. 3) in order to evaluate the contribution of
in Fig. 9. The dispersion coefficient seems to depend on Qg these loops in the dispersion (i.e. uz = 2uz,ref , uy = 2uy,ref ).
according to a square root law, the same as in correlation Results, presented in Table 6 (Ref 5 and 4), show that the
(Eq. (1)). It can be highlighted that the calculations diverge axial velocity oscillations (see Fig. 6) have negligible influence
when the gas flowrate increases beyond 70 L min−1 . on D and gas induced loops have a quite significant impact
The same comparison, with different liquid flowrates, is pre- on D (increase of approximately 25% by doubling the loop
sented in Fig. 10. A good agreement can also be observed for velocity).
2446 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

Fig. 11. Projection of the simulated liquid velocity field on a vertical plane for the industrial reactor (x = 51 m).

1.2

1
Time-averaged vertical liquid velocity Uz (m/s)

91936 cells grid


0.8
151776 cells grid
0.6 194208 cells grid

0.4

0.2

0
0 1 2 3 4 5 6 7 8 9
-0.2

-0.4

-0.6

-0.8

-1
y (m)

Fig. 12. Simulated average vertical liquid velocity on a transversal line (x = 51 m, z = 1.8 m) for three different grid sizes.

Finally, in order to study the impact of turbulence, a simu- 4.3. Scale up on industrial size reactor
lation with a doubled value of  and another with a doubled
value of k has been performed. Results are also presented in In order to validate the chosen hydrodynamics model, RTD
Table 6 (ref 2 and 3). The modification of  induces a small simulations of a plant scale reactor have been carried out. RTD
difference on D (increase of around 25%) whereas doubled the experiments (Potier et al., 2005) have been conducted on a
value of k doubles the value of D (impact of 100%). D seems 3300 m3 (102 m long, 9 m wide, 3.6 high) channel reactor, at
to be strongly linked to turbulence and especially to the value the Nancy Maxéville plant (France). Lithium chloride has been
of the turbulent kinetic energy k in accordance with Eq. (26) used as an inert tracer and injected as a pulse near the liquid
where t is given by Eq. (7). inlet. Lithium chloride concentration was measured by atomic
With all these “numerical experiments”, we validate the fact absorption. It was verified that the tracer was not absorbed
that turbulence is the main cause of the overall dispersion. Tur- on solid matter. The liquid flowrates were 1280, 1650 and
bulent kinetic energy k is the variable which mainly impacts on 2200 m3 h−1 and the linear gas flowrate was approximately
the overall dispersion. But the particle tracking method shows 0.0175 m3 m−1 s−1 ; aeration takes place on both sides of the
that convective dispersion is not negligible, contrary to the con- reactor which causes a double gas induced loop, as seen in the
clusion of the tracer method. simulated flow field presented in Fig. 11. The vertical plane
We can conclude that the RSM turbulence model coupled (x, z) in the middle of the reactor is a symmetry plane there-
with the particle tracking method gives very good RTD results. fore a symmetry boundary condition is used in order to shorten
As it has been said in the RTD simulation part of Section 4.2, calculation time. The grids for this kind of wastewater plant
CL is the preponderant constant in the modelling of turbulent scale reactor simulations are very coarse (average cell volume
dispersion therefore with an empirically deduced CL such as of 0.016 m3 ) but give satisfactory results (Glover et al., 2006;
CL = 3C , good RTD results can also be achieved with the k. Stropky et al., 2007). Grid independency has been verified
turbulence model. (Fig. 12) and a 151 776 cells grid has been used with
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2447

0.00030
experimental low flowrate
simulated low flowrate
experimental medium flowrate
0.00025
simulated medium flowrate
experimental high flowrate
simulated high flowrate
0.00020

E (t) 0.00015

0.00010

0.00005

0.00000
0 4000 8000 12000 16000 20000
Time (s)

Fig. 13. Comparison between experimental and simulated RTD obtained with the RSM model and the particle tracking method on a full scale reactor for
three different liquid flowrates.

boundary conditions similar to those used for the bench scale estimated axial velocity does not have significant effects and
reactor. The physical properties of the mixture of wastewater the molecular diffusion is negligible.
and sludge are approximated to those of water, which is ac- The hydrodynamics model obtained can predict accurately
ceptable because of the low concentration of sludge (less than the dispersion for different gas and liquid flowrates in the bench
8 g L−1 ) in the reactor and the bubble diameter is approximated scale reactor and can predict the dispersion in an industrial
to 1 cm as observed in experiments. sized wastewater reactor with the same precision than the pre-
Experimental results have been treated by curve fitting with viously established empirical correlation (Potier et al., 2005).
a plug flow with axial dispersion model. Simulated results have This correct estimation of the global hydrodynamics allows
been obtained as in Section 4.2 with a particle tracking method. the future implementation of a kinetic model coupled with the
This method avoids possible numerical diffusion due to the hydrodynamics model.
coarse grid. Fig. 13 shows very good agreement between exper-
imental and simulated data. There is less than 5% of error on Notation
the mean residence time and the dispersion coefficient. It can
be highlighted that the greater discrepancy is observed with the C concentration, kg m−3
lowest flowrate, which is due to the difficulty to keep constant Cd drag coefficient, dimensionless
the inlet flowrate of a real wastewater treatment plant during CL constant in the Lagrangian particle motion
such a long period. In fact, some small flowrate variations have model, dimensionless
been noticed. C , C1 , constants in the k. model or RSM model,
C2 , Ck , C dimensionless
d diameter of the dispersed phase (bubble or
5. Conclusion
particle), m
3D Euler–Euler numerical simulations of a cross-flow D dispersion coefficient, m2 s−1
gas–liquid reactor have been presented. These simulations have Dm molecular diffusivity, m2 s−1
been carried out using the CFD code FLUENT. Grid depen- g gravity acceleration, m s−2
dence, discretization schemes and different turbulence models h reactor height, m
have been tested. It has been found that the use of the RSM I momentum transfer term between phases,
turbulence model coupled with a second order discretization kg m−2 s−2
scheme allows to successfully predict the dispersion coefficient I unit tensor, dimensionless
of the reactor. RTD simulations with particle tracking method k turbulent kinetic energy, m2 s−2
with standard parameters give more accurate results than those L reactor length, m
with transport equation solution. But it is also possible to ad- Le characteristic length of turbulent eddies, m
just the simulation to the experiment with the modification of p pressure, Pa
key model parameters (Sct or CL ). CFD allowed to investi- Patm atmospheric pressure, 101 300 Pa
gate the different effects responsible for the dispersion. It has Qg gas flowrate, m3 s−1
been shown that the dispersion is mainly due to turbulence but Ql liquid flowrate, m3 s−1
the dispersion due to convection is not negligible. The poorly Rij Reynolds tensor component, m2 s−2
2448 Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449

Sct turbulent Schmidt number, dimensionless computational fluid dynamics. Chemical Engineering and Processing 40,
t time, s 187–194.
Daly, B.J., Harlow, F.H., 1970. Transport equations in turbulence. The Physics
u velocity, m s−1 of Fluids 15 (11), 2634–2649.
U statistical average velocity, m s−1 Desai, R.B., Kolhatkar, R.V., Joshi, J.B., Ranade, V.V., Malshelkar, R.A.,
u∞ terminal vertical velocity, m s−1 1995. Turbulence structure in bubble disengagement zone: role of polymer
V reactor volume, m3 addition. A.I.Ch.E. Journal 41 (5), 1329–1332.
Ekambara, K., Dhotre, M.T., Joshi, J.B., 2005. CFD simulations of bubble
w reactor width, m column reactors: 1D, 2D and 3D approach. Chemical Engineering Science
x, y, z spatial coordinates 60, 6733–6746.
Elghobashi, S.E., Abou-Arab, T.W., 1982. A two-equation turbulence model
Greek letters for two-phase flows. The Physics of Fluids 26 (4), 931–938.
Fluent Inc., 2005. Fluent user’s guide 6.2. Discrete Phase Models 23, 1–14.
 volume fraction, dimensionless
Glover, G.C., Printemps, C., Essemiani, K., Meinhold, J., 2006. Modeling

ij Kronecker factor, dimensionless of wastewater treatment plants—how far shall we go with sophisticated
 turbulent dissipation rate, m2 s−3 modeling tools? Water Science and Technology 53, 79–89.
l  volume average of turbulent dissipation rate, Gu, H.-Y., Guo, L.-J., 2005. Modelling and simulation of the dynamic
m2 s−3 flow behaviour in a rectangular bubble column. Journal of Engineering
Thermophysics 26, 72–75.
normally distributed random number, dimen- Gunjal, P.R., Ranade, V.V., Chaudhari, R.V., 2003. Liquid distribution and
sionless RTD in trickle bed reactor: experiments and CFD simulations. Canadian
 dynamic viscosity, Pa s Journal of Chemical Engineering 81, 821–830.
 bubble induced turbulent of dissipation rate Jakobsen, H.A., Sannas, B.H., Grevskott, S., Svendsen, H.F., 1997. Modeling
source term, kg m−1 s−4 of vertical bubble-driven flows. Industrial and Engineering Chemistry
Research 36 (10), 4052–4074.
k bubble induced turbulent kinetic energy source Jakobsen, H.A., Lindborg, H., Dorao, C.A., 2005. Modeling of buble column
term, kg m−1 s−3 reactors: progress and limitations. Industrial and Engineering Chemistry
 density, kg m−3 Research 44, 5107–5151.
 surface tension, N m−1 Jamialahmadi, M., Branch, C., Muller-Steinhagen, H., 1994. Terminal bubble
rise velocity in liquids. Chemical Engineering Research and Design 72,
 ,  k constants in the k. model or RSM model, 119–122.
dimensionless Lain, S., Bröder, D., Sommerfeld, M., Göz, M.F., 2002. Modeling
ct eddy crossing time, s hydrodynamics and turbulence in a bubble column using the
Euler–Lagrange procedure. International Journal of Multiphase Flow 28,
e eddy lifetime, s 1381–1407.
Subscripts Lapin, A., Lübbert, A., 1994. Numerical simulation of the dynamics of two-
g gas phase gas–liquid flows in bubble columns. Chemical Engineering Science
49, 3661–3674.
i, j, k spatial coordinates
Leon-Becerril, E., Cockx, A., Liné, A., 2002. Effect of bubble deformation
l liquid on stability and mixing in bubble columns. Chemical Engineering Science
p particle 57, 3283–3297.
q phase index: l or g Levin, M.A., Gealt, M.A., 1993. Biotreatment of Industrial and Hazardous
t turbulent Waste. MCGraw-Hill, New York. 71–72.
Morsi, S.A., Alexander, A.J., 1972. An investigation of particle trajectories in
tr tracer two-phase flows systems. Journal of Fluids Mechanics 55 (2), 193–208.
Operator Mudde, R.F., Simonin, O., 1999. Two- and three-dimensional simulations of
a bubble plume using a two-fluid model. Chemical Engineering Science
az averaged value of a on a line along z 54, 5061–5069.
ax,y averaged value of a on a (x, y) plane Olmos, E., Gentric, C., Midoux, N., 2003. Numerical description of flow
regime transitions in bubble column reactors by multiple gas phase model.
Chemical Engineering Science 58, 2113–2121.
Acknowledgements Pfleger, D., Becker, S., 2001. Modelling and simulation of the dynamic
flow behaviour in a bubble column. Chemical Engineering Science 56,
The authors would like to thank Richard Lainé for his help 1737–1747.
Potier, O., Leclerc, J.P., Pons, M.N., 2005. Influence of geometrical and
with the LDV experimental setup, Benoit Fiers for his help operational parameters on the axial dispersion in an aerated channel reactor.
in the experimental works and Noël Midoux for his scientific Water Research 39, 4454–4462.
contribution. Ranade, V.V., 2002. Process system engineering. Computational Flow
Modelling for Chemical Reactor Engineering, vol. 5. Academic Press,
New York, pp. 57–112.
References Roustan, M., Line, A., 1996. Rôle du brassage dans les procédés biologiques
d’épuration. Tribune de l’eau 5–6, 109–115.
Arnold, F.C., 1988. Physical model for two-phase flow in steam injection Sato, Y., Sekoguchi, K., 1975. Liquid velocity distribution in two-phase bubble
wells. Conference Paper, American Institute of Chemical Engineers, flow. International Journal of Multiphase Flow 2 (1), 79–95.
National Meeting, New York, United States. pp. 42–76. Sokolichin, A., Eigenberger, G., Lapin, A., 2004. Simulation of buoyancy
CD-Adapco-Group, Computational Dynamics Limited, Manual – Methodo- driven bubbly flow: established simplifications and open questions.
logy, STAR-CD version 3.2, 2003. A.I.Ch.E. Journal 50, 24.
Cockx, A., Do-Quang, Z., Audic, J.M., Liné, A., Roustan, M., 2001. Global Stropky, D., Pouqatch, K., Nowak, P., Salcudean, M., Pagorla, P., Gartshore,
and local mass transfer coefficients in waste water treatment process by I., Yuan, J.W., 2007. RTD (residence time distribution) predictions in large
Y. Le Moullec et al. / Chemical Engineering Science 63 (2008) 2436 – 2449 2449

mechanically aerated lagoons. Water Science and Technology 55 (11), Villermaux, J., 1993. Génie de la réaction chimique, conception et
29–36. fonctionnement des réacteurs. TEC & DOC - Lavoisier, 159–193.
Talvy, S., Cockx, A., Line, A., 2007. Modeling hydrodynamics of gas–liquid Zhang, L., Pan, Q., Rempel, G.L., 2007. Residence time distribution in a
airlift reactor. A.I.Ch.E. Journal 53 (2), 335–353. multistage agitated contactor with Newtonian fluids: CFD prediction and
Thyn, J., Ha, J.J., Strasak, P., Zitny, R., 1998. RTD prediction, modelling experimental validation. Industrial and Engineering Chemistry Research
and measurement of gas flow in reactor. Nukleonika 43 (1), 95–114. 46 (11), 3538–3546.
Troshko, A., 2006. Best practice for modelling bubble column reactor with Zhao, Y., Chen, G., Yuan, Q., 2006. Liquid–liquid two-phase flow patterns
FLUENT. www.fluentusers.com. in rectangular microchannel. A.I.Ch.E. Journal 52, 4052–4060.
Vial, C., Poncin, S., Wild, G., Midoux, N., 2005. Experimental and theoretical
analysis of axial dispersion in the liquid phase in external-loop airlift
reactors. Chemical Engineering Science 60, 5945–5954.

You might also like