You are on page 1of 8

Desalination 357 (2015) 225–232

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

Modeling concentration polarization in crossflow microfiltration of


oil-in-water emulsion using shear-induced diffusion; CFD and
experimental studies
Akbar Asadi Tashvigh, Amir Fouladitajar ⁎, Farzin Zokaee Ashtiani
Department of Chemical Engineering, Amirkabir University of Technology, No. 424, Hafez Ave., Tehran, Iran

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• A molecular and shear induced diffu-


sion model was developed for MF of
O/W emulsion.
• Predicted permeate flux showed good
agreement with the experimental data.
• CP was predicted using this new model
in MF of oily wastewater.
• The model showed high capability to
describe CP in an unsteady-state condi-
tion.
• The model successfully predicted flux
for a wide range of oil concentrations.

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

Article history: Membrane filtration has become firmly established as a primary technology for ensuring treatment of oily waste-
Received 30 September 2014 water. A detailed study on concentration polarization (CP) during microfiltration (MF) of oil-in-water emulsions
Received in revised form 29 November 2014 was carried out in order to predict permeate flux decline. A two dimensional computational fluid dynamics (CFD)
Accepted 1 December 2014
modeling was conducted using the finite difference method. A combination of molecular and shear-induced dif-
Available online 5 December 2014
fusion coefficient was employed to describe back-diffusion of oil droplets from the membrane wall toward the
Keywords:
bulk. Evaluation of the proposed approach was carried out by comparing the results to the experimental data
Microfiltration in which good agreement was observed almost in every operating condition. The effects of oil concentration,
Oil-in-water emulsion trans-membrane pressure (TMP), and cross-flow velocity (CFV) on the model precision were also studied. Final-
Concentration polarization ly, the growth rate of CP layer along the membrane length was predicted in different operating conditions, in
Shear-induced diffusion which the obtained trend was found to be in agreement with the CP theory.
CFD modeling © 2014 Elsevier B.V. All rights reserved.

1. Introduction of a higher concentration of retained components at the membrane wall


compared to that of the bulk solution, has remained one of the most
A large amount of wastewater in the form of oil-in-water emulsion is important bottlenecks in pressure-driven membrane processes. This is
generated from various industries such as metallurgical, transportation, governed by two simultaneous processes: (1) convective transport of
oil, gas, and petrochemical companies [1–4]. Microfiltration (MF) has droplets towards the membrane along with the permeate flow and
proven to be a suitable technique in a large variety of applications in- (2) back-transport of droplets from the concentrated layer into the bulk
cluding food industry, biotechnology, and treatment of oil-in-water phase [9]. In MF processes, concentration polarization mainly leads to
emulsion [5–8]. Nevertheless, concentration polarization (CP), formation the formation of a cake layer at the membrane wall, which has a negative
effect on permeate flux [10–12].
⁎ Corresponding author. The effects of CP on the membrane performance have been experi-
E-mail address: fouladi@aut.ac.ir (A. Fouladitajar). mentally investigated in recent decades [1,13–16], but fewer theoretical

http://dx.doi.org/10.1016/j.desal.2014.12.001
0011-9164/© 2014 Elsevier B.V. All rights reserved.
226 A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232

(a)

100 mm

5 z
yx 3 mm
0

(b)

100 mm

yx 3 mm Retentate
Feed

Porous MF

Local flux

Fig. 1. (a) Module geometry; (b) the solution domain.

works have been conducted to model CP and its effects on the permeate Darcy's law, and mass transfer equation based on the finite difference
flux [17,18]. Most of these studies have only predicted the final steady method, and used a constant molecular diffusion coefficient.
state permeate flux and could not describe the permeate flux as a func- There are some diffusion coefficient models which can be considered
tion of time [19–24]. responsible for the back-diffusion of retained components from mem-
The first simulation of flow in a membrane was undertaken in lam- brane surface toward the bulk. Ripperger and Altmann [30] reported a
inar conditions in channels with porous walls [25]. Nassehi [26], used list of various diffusion coefficient models in cross-flow filtration
Darcy's equation to represent the porous wall conditions. For prediction processes. Belfort et al. [31] showed that shear-induced diffusion can be
of permeate flux of MF processes, Damak et al. [27], used a two dimen- considered as the predominant back-diffusive mechanism during
sional convective diffusion equation with a constant molecular diffusion microfiltration of solutions with particles in the range of 0.5–30 μm.
coefficient in mass balance equation. They solved equations using a fi- Shear-induced diffusion was first introduced in microfiltration by Zydney
nite difference scheme and predicted the final steady state permeate and Colton [32]. Kramadhati et al. [33], presented a modified shear-
flux. Vela et al. [28] used a semi-empirical cross-flow ultrafiltration induced diffusion model for membrane dominated microfiltration exper-
model to predict permeate flux as a function of time for different oper- iments to account polydispersity of suspension particle. Afterwards,
ating conditions. As a negative point, their model needed some empiri- Kromkamp et al. [34] developed a computer simulation model based on
cal parameters which could not be estimated theoretically. Pak et al. the lattice Boltzmann method which was capable of studying concentra-
[29] developed a 2D numerical solution of the coupled Navier–Stokes, tion polarization behavior in cross-flow microfiltration. Their model
Concentraon

Fit 5000
Fit 10000
Fit 30000
Fit 70000
Fit 90000

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


Normalized membrane length

Fig. 2. Mesh dependency test (concentration vs. position for different grids).
A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232 227

regime kept in laminar condition (Reynolds numbers were 565, 1130


CP and 1700) [35]. It was assumed that, for a long time pure water was
C flowing in the slit and immediately feed concentration was changed,
y
x cake
l Input starting estimation
Porous membrane of local flux; v(x)

Fig. 3. Concentration gradient on the membrane surface.

Solve Eqs. 1-7


solved the discrete Boltzmann equation to simulate the flow instead of
Navier–Stokes equations. The shear-induced diffusion coefficient was Reiterate
used, as well. Finally, the permeate flux decline was predicted by solving
those complex equations. Calculate the concentration
A complete description of the process requires solving the fully polarization layer thickness ( c)
coupled Navier–Stokes, continuity, and convection–diffusion equations
which is a complex and time consuming problem. In the present work,
from Eq. (11)
since the permeation velocity is small and does not significantly alter
the axial flow profile in membrane processes, and due to laminar flow
in the slit, a Poiseuille flow profile was assumed to simplify the coupled
momentum and mass transport problems. A new numerical convec-
tion–diffusion CP model was developed which used a combination of
Recalculate the local flux from
molecular and shear-induced diffusion coefficient. Finally the predicted Darcy’s law, Eq. (8)
data were compared to the experimentally-obtained permeate flux and
the accuracy and applicability of the model for microfiltration of oil-in-
water emulsion were discussed.

2. Theory
Local flux; v(x)
2.1. Module geometry and problem description converge?

In the current study, the focus was on modeling CP in a rectan-


gular membrane module. The module was a slit with dimensions
of 3 × 50 × 100 mm3 as height, width and length of the rectangle, re-
spectively. The MF membrane was the lower 50 × 100 mm2 wall of
the slit. The module is depicted schematically in Fig. 1(a). The feed
was a mixture of oil droplets with a diameter of about 1 μm in water.
Write final v(x); for this
2.2. Geometry reconstruction time step

As the ratio of height to width of the feed channel of module was far
than unity (H/W = 0.06), therefore, the effects of the beginning and end Go to next time step
of walls on flow were neglected, and the model was considered to be
two-dimensional. Hence, a 3 × 100 mm2 rectangle was considered as
the solution domain, which, has been schematically depicted in
Fig. 1(b). In order to eliminate the effect of the mesh size on the results, No
initial simulations were carried out until the finer meshes resulted in Time reaches to
constant results. After mesh dependency tests (Fig. 2) it was proven
it’s final value?
that 90,000 cells were adequate to discretize the 2D simulation domain.

2.3. Governing equations


Yes
In this study, the oil-in-water emulsion was assumed to be an in-
compressible, isothermal, and uniform solution. A Poiseuille flow profile
was assumed because the permeation velocity was small enough and Write the matrix of
did not significantly alter the axial flow profile, as well as the flow v(x,t)

Table 1
The values of parameters used in Eq. (10).

Parameter Value Unit

rm 2.49415 × 1012 m−2


rC 9.7155 × 1015 m−2
l 1.25 × 10−4 m
Fig. 4. Iterative solution procedure.
228 A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232

Table 2 In addition, longitudinal velocity in the slit is expressed by Eq. (3)


Physical properties of the feed and the membrane. [36]:
Feed properties
   2 
Oil type Gasoil ΔP x H 2 H
Oil density at 15 °C 845 kg/m3 u¼ y− − ð3Þ
2μL 2 2
Kinematic viscosity of oil at 37.8 °C 3.8 cSt
Surfactant Polyoxyethylene (80) Sorbitan
Monooleate (Tween 80, Merck) where, △Px is the pressure drop in the module length, H and L are the
height and length of the membrane module, and μ indicates the fluid
Membrane properties
Material (PVDF, Millipore Co) viscosity.
Mean pore size 0.45 μm Initial condition, (it was assumed that, at first, pure water was
Thickness 125 μm flowing in the slit), t = 0:
Porosity 70%
Effective membrane area 50 cm2
C ðx; y; 0Þ ¼ 0: ð4Þ

Boundary conditions.
Inlet boundary condition, x = 0:
so the longitudinal velocity profile was not considered a function of time
and remains fully developed as inlet concentration changed. C ð0; y; t Þ ¼ C 0 : ð5Þ
Mass transport equation (dispersion in x direction was neglected) is
written in Eq. (1), [36]:
Membrane surface condition, (intrinsic rejection of the membrane
  was assumed to be 100%), y = 0
∂C ∂C ∂C ∂C ∂C
þu þv ¼ Dy ð1Þ
∂t ∂x ∂y ∂y ∂y
∂C ðx; 0; t Þ
−Dy ¼ vcðx; 0; t Þ: ð6Þ
∂y
where x and y are the longitudinal and transverse coordinates, respec-
tively, t is time, C is concentration of oil, u and v are the longitudinal
and transverse velocities, respectively, and Dy indicates diffusion coeffi- Impermeable wall boundary condition, y = H
cient of oil in water in y direction. It should be noted that a combination
of molecular and shear-induced diffusion coefficient (Eq. (2)) was used ∂C ðx; H; t Þ
¼ 0: ð7Þ
in the mass conservation equation which is described in Eq. (2) [37,38]: ∂y

∂u 2 2  8:8φ
 Darcy's law
Dy ¼ Dm þ 0:33 ap φ 1 þ 0:5e ð2Þ
∂y
TMP
vðxÞ ¼ ð8Þ
where, Dm is the molecular diffusion coefficient of oil in water, φ repre- μRt
sents oil volume fraction, and ap is oil droplet size. The second term in
the right hand side of Eq. (2) indicates shear-induced diffusion. where, Rt is representing the overall membrane resistance, given by the

Fig. 5. The laboratory scale experimental setup.


A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232 229

115

105 Exp. 1 (C=5000 mg/L)


Model. 1 (C=5000 mg/L)

Permeate flux (L/h.m2)


95 Exp. 2 (C=10000 mg/L)
85 Model. 2 (C=10000 mg/L)
Exp. 3 (C=20000 mg/L)
75 Model. 3 (C=20000 mg/L)
65

55

45

35
50 550 1050 1550 2050 2550 3050 3550 4050
Time (s)

Fig. 6. The effect of feed concentration on permeate flux for TMP of 1 bar and CFV of 0.2 m/s.

following equation: previous works [6,39]. Table 1, illustrates the values of parameters
used in Eq. (10).
Rt ¼ rm l þ r cake δcake þ r CP δCP ð9Þ Darcy's law was coupled with the mass balance equation (Eq. (1)) in
an iterative method.
where, rm, rcake and rCP are the specific resistances of the membrane,
cake and CP layers respectively, and l, δcake and δCP are the thicknesses
of membrane, cake and CP layers, respectively, which are depicted in 2.4. Solution method
Fig. 3. In order to simplify the problem, Rt was obtained by the following
equation For the numerical solution of Eq. (1), a finite difference (FD) scheme
was employed to discretize Eqs. (1)–(7), this included a space averaged
Rt ¼ r m l þ r C δ C ð10Þ centered-difference approximation for time derivative, a time averaged
centered-difference approximation for axial derivative, and modified
where, δC is the total thickness of CP and cake layers, and rC represents Crank–Nicholson analogs for the vertical derivatives. More description
the mean specific resistance of the cake and CP layers obtained from has been given in Ref. [40].

(a)
0.8
Exp. 1 (TMP=0.5 bar)
0.7 Model. 1 (TMP=0.5 bar)
PF / Initial PF (J/J0)

Exp. 2 (TMP=1 bar)


0.6 Model. 2 (TMP=1 bar)
Exp. 3 (TMP=2 bar)
0.5 Model. 3 (TMP=2 bar)

0.4

0.3

0.2
50 550 1050 1550 2050 2550 3050 3550 4050
Time (s)

(b)

280

TMP=0.5 bar TMP=1 bar TMP=2 bar


Predicted permeate flux

260

240
(L/h.m2)

220

200

180

160
160 180 200 220 240 260 280

Experimentally measured permeate flux (L/h.m2)

Fig. 7. (a) The effect of TMP on permeate flux decline. (b) Initial permeate flux in three different TMPs; predicted versus experimental.
230 A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232

125
Exp. 1 (CFV=0.1 m/s)
115

Permeate flux (L/h.m2)


Model. 1 (CFV=0.1 m/s)
105
Exp. 2 (CFV=0.2 m/s)
95
Model. 2 (CFV=0.2 m/s)
85 Exp. 3 (CFV=0.3 m/s)
75 Model. 3 (CFV=0.3 m/s)
65
55
45
35
50 550 1050 1550 2050 2550 3050 3550 4050
Time (s)

Fig. 8. Effect of CFV on permeate flux.

It should be noted that the v in Eqs. (1) and (6) is unknown and is in duplicate and results were reproducible with a ±5% error. More de-
dependent on the CP layer thickness, δc. This CP layer is equal to the dis- scription regarding module design, the experimental setup and its pro-
tance from the membrane surface where the oil concentration is 99% of cedure have been described in previous studies [11].
the bulk oil concentration [29].
4. Results and discussion
C−C 0
b0:001 ð11Þ
C0 4.1. The effects of oil concentration

An iterative technique was required for solution and a time step As depicted in Fig. 6, the effects of three different oil concentrations
equal to 0.05 was chosen. A computer code was written in MATLAB® on the permeate flux were theoretically modeled and compared with
software for the model described above and was solved by two Intel® the experimental data. As shown in Fig. 6, the model provided excellent
Xeon® E5620 2.4 GHz CPU and 20 GB of RAM. The procedure of the it- prediction for permeate flux and the shape of predicted curves showed
erative solution is depicted in Fig. 4. a good agreement with the experimental results. With decreasing the
feed concentration, the final permeate flux tended to increase. This be-
3. Experimental havior was also satisfied by the model, but the model overpredicted the
permeate flux in the beginning of filtration time. As shown in Fig. 6, the
3.1. Feed preparation and the membrane model provided good agreement in all tested concentrations. The best
agreement between experimental and predicted data was obtained for
The oil-in-water emulsion was prepared by mixing gasoil and sur- the feed concentration of 20,000 mg/L. It should be noted that previous
factant in distilled water at a mixing rate of 12,000 rpm for 30 min. models [20,41] which used a constant molecular diffusion coefficient
The surfactant was used at a concentration of 100 mg/L. A brief descrip- failed to provide accurate predictions in high concentrations. It implies
tion about feed and membrane properties is presented in Table 2. that the shear-induced part of diffusion coefficient is of great impor-
tance as the feed concentration increases and should be considered in
3.2. Experimental setup the model. Therefore, using a combination of molecular and shear-
induced diffusion coefficient seems to be necessary in the modeling of
A schematic diagram of the experimental setup is shown in Fig. 5. oil-in-water emulsion microfiltration.
The stable oil-in-water emulsion was kept in a 10 L feed tank and was
delivered to the membrane module by a centrifugal recirculation 4.2. Effect of TMP
pump, controlled by an inverter. This pump also provided the required
constant operating pressure to the feed. Two needle valves were Fig. 7(a) shows the permeate flux versus the filtration time for three
installed on the feed and retentate lines to adjust the stream flow rate. different TMPs at the oil concentration of 5000 mg/L and CFV of 0.2 m/s.
Besides, the valve installed after the membrane module, could exert a As shown in Fig. 7(a), the shape of the predicted curve was very similar
backpressure along the membrane unit. More details about the appara- to the experimental results. The rate of flux decline was overpredicted
tus can be found in previous publications [12,39]. at lower TMPs, while the model predictions tended to be underestimated
The effects of TMP, CFV, and oil concentration were experimentally
and numerically studied. The feed with three different oil concentra-
Table 3
tions (namely 5000, 10,000, and 20,000 mg/L) was fed into the mem- Experimental and predicted permeate flux.
brane module. The process was conducted in different TMPs of 0.5, 1,
Operating condition Received filtrate volume Percentage
and 2 bar, and CFV of 0.1, 0.2, and 0.3 m/s. According to the module di-
per unit membrane area error
mensions, these velocities resulted in Reynolds number of 565, 1130, (L/m2)
and 1700, respectively, which can be called laminar flow regime. Finally,
Oil concentration TMP CFV Experimental Predicted
27 designed experiments and CFD simulation runs, with full factorial
(mg/L) (bar) (m/s)
design, were conducted to investigate the effects of operating condi-
5000 1 0.2 72.5 75.1 3.41
tions on the permeate flux decline and CP layer in microfiltration of
10000 1 0.2 61.6 66.4 7.77
oil-in-water emulsion. The procedure of each run was a combination 20000 1 0.2 64.4 62.4 −3.10
of measuring pure water permeability and then pumping the oil-in- 5000 0.5 0.2 67.5 68.1 0.89
water emulsion into the module. The system was operated in a recycling 5000 2 0.2 183.5 150.7 −17.90
mode for 70 min and then the cleaning-in-place (CIP) procedure was 5000 1 0.1 68.2 68.4 0.21
5000 1 0.3 81.6 83.1 1.69
done. In addition, it should be added that all experiments were done
A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232 231

0.4

Relative CP layer thickness ( c/H)


0.35
CFV=0.1 m/s CFV=0.2 m/s CFV=0.3 m/s
0.3

0.25

0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Relative membrane length (x/L)

Fig. 9. Effect of CFV on growth of concentration polarization layer; TMP of 1 bar and concentration of 5000 mg/L.

as TMP increased. In Fig. 7(b), predicted initial permeate flux is presented phenomenon in microfiltration of oil-in-water emulsions in a flat
versus that experimentally measured. It can be seen from Fig. 7(b) that sheet membrane module under laminar flow. Mass transport equa-
model prediction is lower than experimental for higher TMPs. Since the tion was mathematically expressed using a two dimensional convec-
internal fouling was ignored in this model, the model underestimated tive diffusion equation in a Cartesian coordinate system and was
the permeate flux at higher TMPs. In addition, rc has been considered to solved based on the finite difference scheme. The obtained results
be constant in all operating conditions, but it can change in different showed that the model is capable to predict permeate flux for
TMPs which can result in some errors. This is worth to mention that every feed concentration and CFV. However, at higher TMPs the pre-
some other studies [28,42] which have employed semi-empirical models dicted results showed some differences with the experimental data.
to predict flux decline, failed to provide accurate results at higher TMPs Additionally, since the effects of CFV, oil concentration, and the oil
and more discrepancy between model and experimental data were ob- droplet size were considered in the applied diffusion model, the pre-
served. It confirms that the theoretical model developed in the present dicted results were in good agreement with the experimental data,
study is more capable than semi-empirical models, even at higher TMPs. almost in every operating condition.
The CP layer thickness grew with a sharp rate in the entrance of the
4.3. Effect of CFV module and kept on with an approximately constant rate when about
40% of the membrane length was passed. Since the diffusion coefficient
The experimental and predicted permeate fluxes in three CFVs and a was assumed to be a function of oil concentration, droplet size, CFV and
concentration of 5000 mg/L and TMP of 1 bar are presented in Fig. 8. oil physicochemical properties, the model is applicable for prediction of
As expected, the higher the feed velocity, the higher the permeate oil-in-water emulsion microfiltration in a wide range of operating
flux. It is due to the fact that higher shear rate can cause a thinner con- conditions.
centration polarization layer and faster particle back-transport. Since
the effect of CFV on the diffusion was considered in the model, the pre- Nomenclature
dicted permeate flux was in good agreement with the experimental CFV cross-flow velocity (m/s)
data for any feed velocity. As it is depicted in Fig. 8, with an increment TMP trans-membrane pressure (Pa)
of CFV from 0.1 to 0.2 m/s, the enhancement of the final permeate flux C oil concentration (mg/L)
was found to be 15.8%, whereas this enhancement was 8.6% when in- Dy diffusion coefficient in y direction (m2/s)
creasing from 0.2 to 0.3 m/s. This means that this velocity increment Dm molecular diffusion coefficient (m2/s)
was not efficient enough to make the flow more turbulent. Therefore, H height the membrane module (mm)
considerable improvement was not achieved in this operating condi- L length of the membrane module (mm)
tion. This was found to be in agreement with the literature [43]. The Rt overall membrane resistance (m−1)
error analysis for the model predictions in different operating condi- W width of the membrane module (mm)
tions have been given in Table 3. ap oil droplet size (m)
rm intrinsic resistance of the membrane (m−2)
4.4. Prediction of CP layer thickness rc mean specific resistance of the cake and CP layer (m−2)
rcake specific resistance the cake layer (m−2)
Fig. 9 illustrates the predicted concentration polarization layer thick- rCP specific resistance the CP layer (m−2)
ness along the membrane length, for a concentration of 5000 mg/L and l membrane thickness (m)
TMP of 1 bar, under steady-state condition. As shown in Fig. 9, with an t time (s)
increase in CFV, the CP layer thickness decreased. In addition, as the u X-component of velocity (m/s)
fluid passed almost 40% of the membrane length, the thickness of CP v Y-component of velocity (m/s)
layer grew with a constant slope. This is of great importance in design- x longitudinal coordinate
ing the optimum length of a module. The similar trend was observed in y transverse coordinate
tubular membranes in which the growth rate of the CP layer got almost ρ density (kg/m3)
constant in x/R N 150 [29]. μ viscosity of fluid (Pa·s)
φ oil volume fraction
5. Conclusions δc summation of CP and cake layer thicknesses (m)
δcake cake layer thickness (m)
This study investigated the applicability of the combination of δCP CP layer thickness (m)
molecular and shear-induced diffusion model to describe CP ΔPx pressure drop in the x direction (Pa)
232 A. Asadi Tashvigh et al. / Desalination 357 (2015) 225–232

References [22] G. Denisov, Theory of concentration polarization in cross-flow ultrafiltration: gel-


layer model and osmotic-pressure model, J. Membr. Sci. 91 (1994) 173–187.
[1] B. Chakrabarty, A. Ghoshal, M. Purkait, Ultrafiltration of stable oil-in-water emulsion [23] A.L. Zydney, Stagnant film model for concentration polarization in membrane sys-
by polysulfone membrane, J. Membr. Sci. 325 (2008) 427–437. tems, J. Membr. Sci. 130 (1997) 275–281.
[2] M. Zare, F. Zokaee Ashtiani, A. Fouladitajar, CFD modeling and simulation of concen- [24] W.F. Blatt, A. Dravid, A.S. Michaels, L. Nelsen, Solute polarization and cake formation
tration polarization in microfiltration of oil–water emulsions; application of an in membrane ultrafiltration: causes, consequences, and control techniques, Mem-
Eulerian multiphase model, Desalination 324 (2013) 37–47. brane Science and Technology, Springer, 1970, pp. 47–97.
[3] H. Ohya, J. Kim, A. Chinen, M. Aihara, S. Semenova, Y. Negishi, O. Mori, M. Yasuda, [25] A.S. Berman, Laminar flow in channels with porous walls, J. Appl. Phys. 24 (1953)
Effects of pore size on separation mechanisms of microfiltration of oily water, 1232–1235.
using porous glass tubular membrane, J. Membr. Sci. 145 (1998) 1–14. [26] V. Nassehi, Modelling of combined Navier–Stokes and Darcy flows in crossflow
[4] A.B. Koĺtuniewicz, R.W. Field, Process factors during removal of oil-in-water emul- membrane filtration, Chem. Eng. Sci. 53 (1998) 1253–1265.
sions with cross-flow microfiltration, Desalination 105 (1996) 79–89. [27] K. Damak, A. Ayadi, B. Zeghmati, P. Schmitz, Concentration polarisation in tubular
[5] A. Fouladitajar, F. Zokaee Ashtiani, H. Rezaei, A. Haghmoradi, A. Kargari, Gas sparging membranes—a numerical approach, Desalination 171 (2005) 139–153.
to enhance permeate flux and reduce fouling resistances in cross flow [28] M.C. Vincent Vela, S.Á. Blanco, J.L. García, E.B. Rodríguez, Application of a dynamic
microfiltration, J. Ind. Eng. Chem. 20 (2014) 624–632. model for predicting flux decline in crossflow ultrafiltration, Desalination 198
[6] Z. Jalilvand, F. Zokaee Ashtiani, A. Fouladitajar, H. Rezaei, Computational fluid dy- (2006) 303–309.
namics modeling and experimental study of continuous and pulsatile flow in flat [29] A. Pak, T. Mohammadi, S. Hosseinalipour, V. Allahdini, CFD modeling of porous
sheet microfiltration membranes, J. Membr. Sci. 450 (2014) 207–214. membranes, Desalination 222 (2008) 482–488.
[7] R. Ghidossi, D. Veyret, P. Moulin, Computational fluid dynamics applied to mem- [30] S. Ripperger, J. Altmann, Crossflow microfiltration—state of the art, Sep. Purif.
branes: state of the art and opportunities, Chem. Eng. Process. 45 (2006) 437–454. Technol. 26 (2002) 19–31.
[8] N. Javadi, F. Zokaee Ashtiani, A. Fouladitajar, A. Moosavi Zenooz, Experimental stud- [31] G. Belfort, R.H. Davis, A.L. Zydney, The behavior of suspensions and macromolecular
ies and statistical analysis of membrane fouling behavior and performance in solutions in crossflow microfiltration, J. Membr. Sci. 96 (1994) 1–58.
microfiltration of microalgae by a gas sparging assisted process, Bioresour. Technol. [32] A.L. Zydney, C.K. Colton, A concentration polarization model for the filtrate flux in
162 (2014) 350–357. cross-flow microfiltration of particulate suspensions, Chem. Eng. Commun. 47
[9] C.M. Silva, D.W. Reeve, H. Husain, H.R. Rabie, K.A. Woodhouse, Model for flux predic- (1986) 1–21.
tion in high-shear microfiltration systems, J. Membr. Sci. 173 (2000) 87–98. [33] N. Kramadhati, M. Mondor, C. Moresoli, Evaluation of the shear-induced diffusion
[10] J. Kromkamp, M. van Domselaar, K. Schroën, R. van der Sman, R. Boom, Shear- model for the microfiltration of polydisperse feed suspension, Sep. Purif. Technol.
induced diffusion model for microfiltration of polydisperse suspensions, Desalina- 27 (2002) 11–24.
tion 146 (2002) 63–68. [34] J. Kromkamp, A. Bastiaanse, J. Swarts, G. Brans, R. Van Der Sman, R. Boom, A suspen-
[11] A. Fouladitajar, F. Zokaee Ashtiani, A. Okhovat, B. Dabir, Membrane fouling in sion flow model for hydrodynamics and concentration polarisation in crossflow
microfiltration of oil-in-water emulsions; a comparison between constant pressure microfiltration, J. Membr. Sci. 253 (2005) 67–79.
blocking laws and genetic programming (GP) model, Desalination 329 (2013) 41–49. [35] S. Kim, E. Hoek, Modeling concentration polarization in reverse osmosis processes,
[12] A. Fouladitajar, F.Z. Ashtiani, B. Dabir, H. Rezaei, B. Valizadeh, Response surface Desalination 186 (2005) 111–128.
methodology for the modeling and optimization of oil-in-water emulsion separa- [36] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, John Wiley & Sons,
tion using gas sparging assisted microfiltration, Environ. Sci. Pollut. Res. (2014), 2007.
http://dx.doi.org/10.1007/s11356-014-3511-6. [37] D. Leighton, A. Acrivos, Measurement of shear-induced self-diffusion in concentrat-
[13] M.C. Porter, Concentration polarization with membrane ultrafiltration, Ind. Eng. ed suspensions of spheres, J. Fluid Mech. 177 (1987) 109–131.
Chem. Prod. Res. Dev. 11 (1972) 234–248. [38] D. Leighton, A. Acrivos, The shear-induced migration of particles in concentrated
[14] W. Bowen, J. Calvo, A. Hernandez, Steps of membrane blocking in flux decline dur- suspensions, J. Fluid Mech. 181 (1987) 415–439.
ing protein microfiltration, J. Membr. Sci. 101 (1995) 153–165. [39] H. Rezaei, F.Z. Ashtiani, A. Fouladitajar, Effects of operating parameters on fouling
[15] H. Bauser, H. Chmiel, N. Stroh, E. Walitza, Control of concentration polarization and mechanism and membrane flux in cross-flow microfiltration of whey, Desalination
fouling of membranes in medical, food and biotechnical applications, J. Membr. Sci. 274 (2011) 262–271.
27 (1986) 195–202. [40] K. Madireddi, R. Babcock, B. Levine, J. Kim, M. Stenstrom, An unsteady-state model
[16] R. Bai, H. Leow, Microfiltration of activated sludge wastewater—the effect of system to predict concentration polarization in commercial spiral wound membranes, J.
operation parameters, Sep. Purif. Technol. 29 (2002) 189–198. Membr. Sci. 157 (1999) 13–34.
[17] G. Keir, V. Jegatheesan, A review of computational fluid dynamics applications in [41] T. Mohammadi, A. Kohpeyma, M. Sadrzadeh, Mathematical modeling of flux decline
pressure-driven membrane filtration, Rev. Environ. Sci. Biotechnol. 13 (2014) 183–201. in ultrafiltration, Desalination 184 (2005) 367–375.
[18] H. Lotfiyan, F. Zokaee Ashtiani, A. Fouladitajar, S. Borhan Armand, Computational fluid [42] M.C. Vincent Vela, S. Álvarez Blanco, J. Lora García, J.M. Gozalvez-Zafrilla, E.
dynamics modeling and experimental studies of oil-in-water emulsion microfiltration Bergantiños Rodríguez, Modelling of flux decline in crossflow ultrafiltration of mac-
in a flat sheet membrane using Eulerian approach, J. Membr. Sci. 472 (2014) 1–9. romolecules: comparison between predicted and experimental results, Desalination
[19] E. Lyster, Y. Cohen, Numerical study of concentration polarization in a rectangular 204 (2007) 328–334.
reverse osmosis membrane channel: permeate flux variation and hydrodynamic [43] T. Mohammadi, M. Kazemimoghadam, M. Saadabadi, Modeling of membrane foul-
end effects, J. Membr. Sci. 303 (2007) 140–153. ing and flux decline in reverse osmosis during separation of oil in water emulsions,
[20] S. Sablani, M. Goosen, R. Al-Belushi, M. Wilf, Concentration polarization in ultrafil- Desalination 157 (2003) 369–375.
tration and reverse osmosis: a critical review, Desalination 141 (2001) 269–289.
[21] I. Sutzkover, D. Hasson, R. Semiat, Simple technique for measuring the concentration
polarization level in a reverse osmosis system, Desalination 131 (2000) 117–127.

You might also like