Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Membrane Science xxx (2004) xxx–xxx

Numerical simulation of the flow in a plane-channel containing


a periodic array of cylindrical turbulence promoters
C.P. Koutsou, S.G. Yiantsios, A.J. Karabelas∗
Department of Chemical Engineering, Chemical Process Engineering Research Institute, Aristotle University of Thessaloniki,
P.O. Box 361, GR 570 01, Thermi, Thessaloniki, Greece

Received 26 August 2003; accepted 14 November 2003

Abstract

This work is aimed at obtaining a better understanding of transport phenomena in membrane elements, where feed-flow spacers (employed
to separate membrane sheets and create flow channels) tend to enhance mass transport characteristics, possibly mitigating fouling and
concentration polarization phenomena, while augmenting pressure drop. A model flow geometry is considered consisting of a plane-channel,
where a regular array of cylinders is inserted, acting as turbulence promoters. Direct numerical simulations using the Navier–Stokes equations
are performed over a range of Reynolds numbers typical of such membrane modules. The results show that the flow becomes unstable at a
critical Reynolds number of 60, and progressively tends to a chaotic state. Qualitative flow features, such as the development and separation of
boundary layers, vortex formation, the presence of high-shear regions and recirculation zones, and the underlying mechanisms are examined.
In addition, quantitative statistical characteristics such as time-averaged velocities, Reynolds stresses, wall-shear rates and pressure drop are
obtained, which are directly related to mass transport enhancement.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Membrane spacers; CFD; Direct numerical simulation

1. Introduction well to the formation of localized stagnation or dead zones,


where the above phenomena may be intensified. Thus, in re-
Membrane processes are among the most advanced meth- cent years the important role of membrane spacers has been
ods in water treatment and desalination. Spirally wound recognized, and several experimental and theoretical stud-
membrane modules are predominantly employed in reverse ies have appeared aiming at understanding the underlying
osmosis and nanofiltration, and they also find use in ultra- phenomena and optimizing spacer configurations.
filtration and microfiltration. A characteristic of this type of Shock and Miquel [1] studied experimentally various
modules is the presence of spacers in the feed-flow and per- commercial spacers, both for the feed and permeate chan-
meate channels. The feed-flow spacers, in particular, which nels, and obtained correlations for the friction and the mass
have the form of non-woven crossed cylinders, serving to transfer coefficients. On the basis of the results obtained,
separate adjacent membrane leaves and create flow passages, they analyzed the effect of permeate flow channel length
tend to promote turbulence and enhance mass transport. and concluded that there is an optimum number of mem-
The term “turbulence promoters” has been accepted in the brane leaves to minimize pressure losses in the permeate
literature, although in reality the unstable flows may not nec- channel. Zimmerer and Kottke [2] studied experimentally
essarily manifest the characteristics of fully developed turbu- the effect of geometrical spacer characteristics on flow pat-
lence. In this way, the undesirable fouling and concentration terns, fluid residence times, pressure drop and mass transfer
polarization phenomena are mitigated. In parallel, however, coefficients and proposed optimal configurations that would
the presence of spacers leads to increased pressure drop, as improve mass transfer at acceptable energy costs due to
increased pressure drop.
∗ Corresponding author. Tel.: +30-2310-996201; Today, with the development of powerful computational
fax: +30-2310-996209. tools and high accuracy numerical methods, direct numer-
E-mail address: karabaj@cperi.certh.gr (A.J. Karabelas). ical simulation of turbulent flows in complex geometries

0376-7388/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2003.11.005
2 C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx

is feasible for Reynolds numbers on the order of a few sidered. In the next section the model problem and the
thousands. Thus, CFD is becoming a useful tool to assist in numerical techniques as well as computational resources
such studies. Karniadakis et al. [3] in a theoretical study of are described. Subsequently, the results obtained as regards
turbulence promoters for heat transfer enhancement stud- the dominant flow features and the underlying mechanisms
ied the flow in a plane-channel, where a periodic array of are discussed. In addition, statistical flow characteristics
small-diameter cylinders is placed. They report that the relevant to mass transfer optimization are described. Fur-
presence of the small cylinders leads to the destabilization thermore, the temporal and spatial resolution of the nu-
of the flow by essentially the same mechanisms as in a merical procedures employed is discussed. Finally, a short
plane-channel, namely formation of Tollmien–Schlichting discussion is provided of the issue of periodic boundary
waves, but at greatly reduced Reynolds numbers. These conditions employed in the flow simulations.
are close to 150, as compared to 5772 for a plane-channel
according to linear hydrodynamic stability theory [4]. Rel-
evant to turbulence promotion are also the studies of Chen 2. Problem formulation and numerical method
et al. [5] and Zovatto and Pedrizzetti [6], who treat various
geometric configurations of a single cylinder located in a The flow geometry consists of a plane-channel of thick-
plane-channel and analyze flow features and stability. ness H, in which an array of equally spaced cylinders is in-
Several theoretical studies have focused closer to serted, as shown in Fig. 1. The cylinders, of diameter D, are
membrane spacer configurations. Karode and Kumar [7] separated by a distance L, and are located symmetrically in
performed three-dimensional simulations. However, they the channel. The channel is infinite in extend in the stream-
employed a k–ε turbulence model, which is of questionable wise and spanwise directions and the flow is assumed to be
usefulness in such a complex geometry. Furthermore, spa- fully developed and two-dimensional. The ratios L/D and
tial resolution appeared not to be adequate. Cao et al. [8] H/D are nominal parameters of the problem under consid-
studied two-dimensional flow in a short channel containing eration. However in this study, taking into account typical
two cylinders at various arrangements. Turbulence modeling characteristics of membrane spacers, these ratios are fixed
was employed in that study as well. Schwinge et al. [9,10] to the values of 6 and 2, respectively.
performed direct numerical simulations of flow and mass The fluid is assumed to be Newtonian and incompress-
transfer in channels containing five cylinders. The effect of ible and the flow is governed by the Navier–Stokes and the
various configurations on flow characteristics such as pres- continuity equations:
sure drop, eddy formation, wall-shear stresses, and mass ∂u 1 2
transfer was examined. Li et al. [11] presented a study of + u · ∇u = −∇P + ∇ u (1a)
flow and mass transfer by performing a three-dimensional ∂t Re
direct numerical simulation in a geometry closely represent- ∇ ·u=0 (1b)
ing membrane spacers. Periodic boundary conditions were
employed, thus enabling them to simulate just one cell of The cylinder diameter, D, which is half of the channel thick-
the pattern formed by the spacer. The effect of spacer geo- ness, is chosen as a length scale, and the average velocity
metrical characteristics was studied and optimal parameters U0 as the velocity scale, defined by

were proposed. Although no information about the spatial 1 H/2
and temporal resolution of these results was provided, lim- U0 = u(x, y, t)dy (2)
H −H/2
ited experimental data on average mass transfer coefficients
appeared to support their simulations. Then, the time scale t0 is D/U0 , and the Reynolds number is
The aim of the present study is to obtain a better under- Re = U0 D/ν, where ν is the kinematic viscosity of the fluid.
standing of the flow behavior, features and structures, as It may be noted that various definitions of Reynolds number
well as statistical characteristics in a model two-dimensional have been used in the literature, e.g. taking into account the
geometry which resembles membrane spacer configura- porosity of the channel due to the presence of the inserts [1].
tions. This understanding is expected to facilitate spacer The boundary conditions are the no-slip and no-penetration
configuration optimization studies, where three-dimensional conditions on the cylinder and channel surfaces. In addi-
simulations and more realistic geometries need to be con- tion, the flow is assumed to be periodic in the streamwise

Fig. 1. Schematic of the flow geometry of a plane-channel and a periodic array of turbulence-promoting cylinders.
C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx 3

direction with periodicity L. Thus, for the velocity It may be emphasized that the results in this study were
obtained through a direct numerical simulation (DNS) of the
u(x + L, y, t) = u(x, y, t) (3)
Navier–Stokes equations, without the introduction of any
Similarly, the pressure is composed of a constant mean pres- approximation by turbulence modeling. The numerical sim-
sure gradient and a periodic part: ulations covered the range of Reynolds numbers up to 200,
  which is representative of the operation of spirally wound
dP membrane modules. In typical desalination membrane ap-
P(x, y, t) = x + P̂(x, y, t) (4a)
dx plications the feed-flow channel Reynolds number does not
exceed 1000. However, this is defined on the basis of the
P̂(x + L, y, t) = P̂(x, y, t) (4b) channel hydraulic diameter, which is four times the cylinder
diameter used in the present definition of Reynolds number.
The assumption of periodicity, which is a typical practice in
A CFD code (FLUENT, v. 6.0.12), which employs the
direct numerical simulations of turbulent flow (i.e. [12]), as
finite-volume method, was used. During each simulation the
well as in other applications of computational methods (i.e.
governing equations were integrated in time by imposing a
molecular dynamics), allows the computational domain to
constant mean pressure gradient until the flow reached a sta-
be restricted to just one cell containing one cylinder. This
tistically steady state, that is when all mean variables such as
issue is further discussed in the next section.
velocities and Reynolds stresses reach steady values. Time
was advanced by a second-order Adams–Moulton scheme
and the convective terms were discretized by a second-order
upwind scheme. In all the simulations the dimensionless
time step was less than 0.0015. The computational grid was
chosen to be finer near the cylinder and the solid surfaces,
in order to closely capture boundary layers and high-shear
zones. By numerical experimentation it was found that a
grid containing 5400 nodes was adequate for the range of
Reynolds numbers covered in this study. This issue is dis-
cussed in more detail in the next section.

3. Results and discussion

3.1. Flow features and structures

For Reynolds numbers up to 60 the flow is steady and


symmetric with respect to the channel symmetry plane. As
may be seen from Fig. 2, the boundary layers which develop

Fig. 2. Steady flow-streamfunction (a and b) and vorticity (c and d) Fig. 3. Unsteady flow instantaneous streamfunction (a) and vorticity (b)
contours at Reynolds numbers of 55 (a and c) and 60 (b and d). contours at Re = 70.
4 C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx

2
Re=70
Re=90
Re=110
Re=170
1
v

-1

-2
Fig. 4. Instantaneous streamfunction contours of the fluctuating part of 0 2 4 6 8 10 12 14
the flow field, at two different instances in time (a and b), at Re = 70. t
Red and purple points correspond to the minimum and maximum values,
respectively. (Please refer to colour version on Science Direct). Fig. 5. Fluctuations of the normal velocity component as a function of
time for various Reynolds numbers, at a position midway between two
consecutive cylinders on the channel symmetry plane.

Fig. 6. Instantaneous streamfunction contours at various time intervals during an eddy turnover period T, for Re = 170.
C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx 5

around the cylinder are separated, thus resulting in the for- 0.5
mation of two symmetric standing eddies behind the cylin-
der. As the Reynolds number increases, the standing eddies 0.4
increase in length and the boundary layer separation points
move towards the cylinder front stagnation point. Looked
wall shear rate

0.3
from the point of view of vorticity transport, as presented in
Fig. 2(c) and (d), the wake of the cylinder consists of two
layers of vorticity, shed in a similar fashion as in the case 0.2
of a cylinder in unbounded flow. An additional important
feature here is the presence of two additional vorticity lay- 0.1
ers bound on the channel walls. Another characteristic of
the flow is the two high-velocity jets formed at the two flow 0
constriction sections above and below the cylinder. These
jets merge after the recirculation zone, leading to a veloc-
-0.1
ity profile with an inflection point at the channel symmetry 0 1 2 3 4 5 6
plane. (a) x
As the Reynolds number is increased beyond the critical
value of 60, the flow is destabilized, apparently through a 0.5
Hopf bifurcation, and becomes periodic. As may be seen
in Fig. 3, the recirculation eddies cease to be symmetric,
0.4
they oscillate and are periodically detached. In parallel, there
is an oscillation in the strength of the two high-velocity
wall shear rate

jets, as well as in the free vorticity layers shed from the 0.3
cylinder.
The above features suggest that the mechanism of the first 0.2
instability of this flow is the same as that for a cylinder in
unbounded flow. Indeed, if the mean flow is subtracted, the
0.1
streamfunction of the oscillating part, as shown in Fig. 4, is
very similar to the center- or cylinder-modes observed by
Karniadakis et al. [3] and Chen et al. [5] (see their Figs. 9(c) 0

and 8, respectively). In addition, the critical Reynolds num-


ber found by Chen et al. [5] for the case of a single cylin- -0.1
0 1 2 3 4 5 6
der in a channel is 83, which is not very different from 60
(b) x
obtained in this study.
Progressively, as the Reynolds number is increased, the Fig. 7. Dimensionless instantaneous shear rate as a function of position
flow instability is intensified. The waveforms of the fluctu- along the lower wall for various time intervals during an eddy turnover
ating velocities, shown in Fig. 5, are enriched in harmonics, period, at Reynolds numbers 77 (a) and 78 (b). At Re = 78 negative
shear rates appear.
become more complex and tend to a chaotic state. Another
characteristic observed is that the basic frequency of the os-
cillations changes rapidly over a narrow range of Reynolds by calculating instantaneous shear rate profiles along one of
numbers from 70 to 90. the walls. As may be seen from Fig. 7, at Re = 77 shear
In Fig. 6 instantaneous streamfunctions are shown during rates are positive throughout an oscillation cycle, whereas
an oscillation cycle at Re = 170. As may be observed, a at Re = 78 instantaneous negative shear rates appear.
characteristic of the flow is that the two high-velocity jets It is instructive to observe flow vorticity contours, as
periodically increase in strength, collide and break through shown in Fig. 8. As the Reynolds number is increased, the
each other. Another distinct characteristic of the flow at free vorticity layers oscillate and gain strength locally to
higher Reynolds numbers is the periodic appearance of re- form separate vortices. These vortices interact with the lay-
circulation eddies at the channel walls. These eddies form ers of wall-bound vorticity forcing it to move away from the
near the point where the recirculation zone ends behind the walls. As a result of this interaction the free vortices move
cylinder, move along the channel walls, and disappear as away from the walls, as well. This behavior bears resem-
they approach the constrictions close to the next cylinder. blance with the one presented in the numerical study of the
Similar wall eddies were observed in the study of Karni- interaction of a free vortex with a flat plate boundary layer
adakis et al. [3] when the small-diameter cylinders were by Luton et al. [13]. Thus, a distinct characteristic of the
placed close to one of the channel walls. flow around cylinders in a channel is that, in addition to the
The critical Reynolds number above which wall eddies vortices shed from the cylinders, conjugate vortices appear,
first appear, was found to be close to 78. This was determined originating from the wall-bound vorticity. Additionally, as
6 C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx

Fig. 8. Instantaneous vorticity contours at Reynolds numbers of 70 (a), 90 (b), 110 (c), and 170 (d). At the higher Reynolds numbers vortices originating
from the walls appear.

pointed out by Zovatto and Pedrizzetti [6], in contrast to the is equal to the Strouhal number (St = ΩD/U0 ), shifts from
case of a cylinder in unbounded flow, counterclockwise vor- 0.295 to 0.495. This is the same Reynolds number range
tices shed from the lower side of the cylinder move upward where wall eddies appear. Further increase of the Reynolds
and vice versa. number results in enrichment of the frequency spectrum with
Finally, another characteristic feature of the flow is the harmonics and subharmonics, but without significant change
frequency of oscillations. By taking velocity traces at various in the dominant frequency. It may be mentioned here that
locations in the channel and applying FFT transforms the a Strouhal number of 0.369 at the onset of instability was
frequency spectrum of the oscillations can be calculated. As obtained by Chen et al. [5] for an analogous case of a single
shown in Fig. 9, over a narrow range of Reynolds numbers cylinder in a channel, whereas for a cylinder in unbounded
from 70 to 90 the dominant dimensionless frequency, which flow the corresponding number is 0.138.

3.2. Statistical flow characteristics

In Fig. 10 time-averaged flow velocity and Reynolds


stress contours are shown. The velocity contours, apart from
their obvious symmetry characteristics (i.e. u is symmetric
and v is anti-symmetric with respect to the channel cen-
terline), reveal that at relatively low Reynolds numbers the
two high-velocity jets retain their identity, whereas as the
Reynolds number increases the intense oscillations result
in a smooth mean profile approaching plug flow. From the
Reynolds stress contours, it may be observed that intense
fluctuations move closer to the cylinder as the Reynolds
number is increased and progressively occupy a larger
extent of the flow domain on both sides of the cylinder.
Apart from flow mixing, a characteristic of spacers
and turbulence promoters in channels is the increase of
wall-shear rates, which is directly related to the increase of
Fig. 9. Dimensionless frequency spectra of the fluctuations of the normal mass transfer coefficients in high Schmidt number transport
velocity for various Reynolds numbers, at a position midway between processes, such as salt transport in membrane desalina-
two consecutive cylinders on the symmetry plane. tion. As mentioned already, optimization of mass transfer
C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx 7

Fig. 10. Contours of time-averaged velocity components and Reynolds stresses for various Reynolds numbers. Red and purple points correspond to the
minimum and maximum values, respectively. (Please refer to colour version on Science Direct).

characteristics is beyond the scope of this paper, since this


would require a more realistic representation of the mem-
brane spacer geometry and a three-dimensional simulation.
However, it is interesting to observe some relevant character-
40
Re=70
istics of the flow studied. In Fig. 11 time-averaged wall-shear
Re=90 rates and their fluctuations (r.m.s. values) are shown. Given
Re=110
D and U0 , as length and velocity scales, the shear rates are
mean wall-shear rate

Re=145
30 Re=170 normalized by U0 /D. As expected, for all Reynolds numbers
the maximum mean shear rate occurs near the constrictions
above and below the cylinder. This maximum is significantly
larger than that in a plane-channel, where the corresponding
20
dimensionless shear rate for laminar flow is uniform and
equal to 3. Fluctuations in shear rate appear to be significant
as well. In all cases a local maximum occurs right before
10 the cylinder where velocity fluctuations are more intense.
Above a Reynolds number of 90, secondary local maxima
appear downstream of the cylinder, which are related to wall
boundary layer separation and the appearance of wall eddies.
0
0 1 2 3 4 5 6 Finally, in Fig. 12 the dimensionless pressure drop,
(a) x which is proportional to a friction coefficient, is shown

8
Re=70 1
Re=90 Steady flow
Re=110
wall shear rate fluctuations

Unsteady flow
Re=145
Re=170 Plane-channel
6
-dP/dx

4
0.1

0
0 1 2 3 4 5 6 0.01
(b) 50 60 70 80 90 100 200
x
Re
Fig. 11. Dimensionless time-averaged shear rate (a) and r.m.s. values
of the fluctuations (b) for various Reynolds numbers, as a function of Fig. 12. Dimensionless mean pressure gradient as a function of Reynolds
position along the lower wall. number. Plane-channel laminar flow correlation: −dP/dx = 3Re−1 .
8 C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx

as a function of Reynolds number. This is also an impor- 0.25


tant parameter in spacer optimization, since it is related N=5400
to the additional energy costs incurred to achieve im- N=10000
0.2 N=21600
proved mass transfer benefits. The correlation found for the
present geometry in the unstable flow regime is −dP/dx =
3.1Re−0.47 . As expected, the presence of the cylinders
Amplitude

0.15
leads to a significant increase in pressure drop compared
with that in a plane-channel, where for laminar flow
−dP/dx = 3Re−1 .
0.1

3.3. Validation of the results


0.05
To test the validity of the simulation results, tests with
grids containing approximately twice and four times the
original number of nodes were performed. In Fig. 13 time- 0 0.5 1 1.5 2
averaged wall-shear rates and r.m.s. values of their fluctu- Frequency
ations are shown. In Fig. 14 the frequency spectra of the
Fig. 14. Dimensionless frequency spectra of the fluctuations of the normal
velocity for three different grids comprised of 5400, 10 000 and 21 600
40 nodes, at Re = 170, at a position midway between two consecutive
cylinders on the channel symmetry plane.
N=5400
N=10000
N=21600
velocity fluctuations at a certain position in the channel are
30
shown. Inspection of Figs. 13 and 14 suggests that not only
mean shear rate

statistical quantities but the transient flow behavior is also


adequately resolved, for the range of Reynolds numbers cov-
20 ered in this study.

3.4. Effect of periodic conditions


10
As mentioned in the previous section, periodic boundary
conditions are usually employed in turbulent flow numerical
simulations. The presence of the periodic array of cylinders
0
0 1 2 3 4 5 6
imposes a strong constraint on the flow to acquire spatially
(a) x periodic characteristics. However, the periodicity of geom-
etry does not necessarily coincide with that of the flow, es-
8
pecially in the unstable flow regime. Ideally, a large number
of cells, each containing one cylinder, should be considered.
N=5400 It is assumed in this study that considering only a single
wall shear rate fluctuations

N=10000
N=21600 periodic cell adequately captures the dominant flow features
6 and the main statistical quantities. To test this assumption,
the flow in a periodic cell of twice the original length,
containing two cylinders was simulated. The same spatial
4 resolution was kept resulting in a grid containing twice the
number of nodes. As may be seen from Fig. 15 a qualitative
difference exists in that the oscillations in two neighboring
cells are now exactly out of phase, whereas the single-cell
2 simulations constrain neighboring cells to oscillate in phase.
It may thus be hypothesized that in a simulation containing
more cells, phase differences would exist between neighbor-
0 ing cells. However, as may be observed, the dominant flow
0 1 2 3 4 5 6 features remain the same as in the single-cell simulations.
(b) x In Fig. 16 time-averaged wall-shear rates and r.m.s. values
Fig. 13. Dimensionless time-averaged shear rate (a) and r.m.s. values of are shown. It is evident that there is qualitative similarity
the fluctuations (b), as a function of position along the lower wall, for three of the flow features, and that the statistical quantities retain
different grids comprised of 5400, 10 000 and 21 600 nodes, at Re = 170. the basic periodicity of the geometry. Moreover, there is
C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx 9

satisfactory quantitative agreement between the single and


the double-cell results. On the basis of these test results,
it is suggested that simulations considering a flow peri-
odicity coinciding with that of the geometry are capable
of adequately capturing the dominant flow features and
may also be used to obtain statistical quantities of interest,
such as pressure drop, wall-shear rates and mass transfer
coefficients.

4. Concluding remarks
Fig. 15. Streamfunction (a) and vorticity contours (b) at Re = 170,
obtained from a simulation where the flow periodicity interval is taken
to be twice that of the geometry.
The flow in a model two-dimensional geometry, con-
sisting of a plane-channel in which an array of cylinders
is located, has been studied in order to obtain a better
understanding of the behavior, the dominant features and
structures, as well as the statistical characteristics. Above
a critical Reynolds number of 60 the flow becomes unsta-
ble due to mechanisms similar to those operating in the
40 case of a cylinder in unbounded flow. Above a Reynolds
number of 78 wall eddies appear, conjugate to those shed
by the cylinder. This distinct characteristic of the flow in
the channel wall is due to the interaction of the vorticity
mean wall shear rate

30
shed by the cylinders with the vorticity layers created on
the channel walls. Three-dimensional effects (i.e. along the
cylinder axis) have not been considered in this study. It is
20 known, for example, that above a Reynolds number close
to 180 three-dimensional instabilities appear in the case of
a cylinder in unbounded flow. However, the dominant flow
10 features remain unaltered [5].
By employing periodic boundary conditions in the
streamwise direction, significant savings in computational
0
costs may be made since the computational domain can
0 2 4 6 8 10 12 be restricted to just one cell containing one cylinder. The
(a) x approach adopted, which is usual practice in turbulent flow
numerical simulations, may be also applied to tackle the
10 mass transfer problem in order to obtain local mass transfer
coefficients. The interested reader may consult Karniadakis
et al. [3] for a similar study of heat transfer. Although
8
the flow periodicity does not coincide with that of the ge-
ometry, its statistical characteristics do so. Thus reliable
mean wall shear rate

6
information on features important for membrane spacer
optimization studies can be obtained in this way. The un-
derstanding obtained in this study is expected to facilitate
4 optimization studies of membrane spacer configuration,
where three-dimensional simulations and more realistic
geometries need to be considered.
2

0 Nomenclature
0 2 4 6 8 10 12
(b) x dP/dx pressure gradient (Pa/m)
Fig. 16. Time-averaged shear rate (a) and r.m.s. values of the fluctuations
D diameter of cylinders (m)
(b), as a function of position along the lower wall, for Re = 170. The H channel height (m)
continuous and dotted lines correspond to simulations where the flow L cylinder spacing (m)
periodicity interval is the same and twice that of the geometry, respectively.
10 C.P. Koutsou et al. / Journal of Membrane Science xxx (2004) xxx–xxx

[2] C.C. Zimmerer, V. Kottke, Effects of spacer geometry on pressure


Nomenclature drop, mass transfer, mixing behavior, and residence time, distribution,
Desalination 104 (1996) 129–134.
N number of nodes [3] G.E. Karniadakis, B.B. Mikic, A.T. Patera, Minimum-dissipation
P pressure (Pa) transport enhancement by flow destabilization: Reynolds’ analogy
P̂ periodic part of pressure (Pa) revisited, J. Fluid Mech. 192 (1988) 365–391.
Re Reynolds number [4] P.G. Drazin, W.H. Reid, Hydrodynamic Stability, Cambridge Uni-
versity Press, Cambridge, 1982.
St Strouhal number [5] J.-H. Chen, W.G. Pritchard, S.J. Tavener, Bifurcation for flow past
t time a cylinder between parallel planes, J. Fluid Mech. 284 (1995) 23–
T turnover period (s) 41.
u u-velocity [6] L. Zovatto, G. Pedrizzetti, Flow about a circular cylinder between
u mean u-velocity parallel walls, J. Fluid Mech. 440 (2001) 1–25.
[7] S.K. Karode, A. Kumar, Flow visualization through spacer filled
u u Reynolds stress (or r.m.s. u-velocity) channels by computational fluid mechanics. I. Pressure drop and
U0 average velocity (m/s) shear rate calculations for flat sheet geometry, J. Membr. Sci. 193
v v-velocity (2001) 69–84.
v mean v-velocity [8] Z. Cao, D.E. Wiley, A.G. Fane, CFD simulations of net-type tur-
v v Reynolds stress (or r.m.s. v-velocity) bulence promoters in a narrow channel, J. Membr. Sci. 185 (2001)
157–176.
x x-coordinate [9] J. Schwinge, D.E. Wiley, D.F. Fletcher, Simulation of the flow around
y y-coornidate spacer filaments between narrow channel walls. 1. Hydrodynamics,
Ind. Eng. Chem. Res. 41 (2002) 2977–2987.
Greek letters [10] J. Schwinge, D.E. Wiley, D.F. Fletcher, Simulation of the flow around
spacer filaments between narrow channel walls. 2. Mass transfer
v kinematic viscosity (m2 /s)
enhancement, Ind. Eng. Chem. Res. 41 (2002) 4879–4888.
Ω shedding frequency (Hz) [11] F. Li, W. Meindersma, A.B. de Haan, T. Reith, Optimization of com-
mercial net spacers in spiral wound membrane modules, J. Membr.
Sci. 208 (2002) 289–302.
References [12] J. Jimenez, P. Moin, The minimal flow unit in near-wall turbulence,
J. Fluid Mech. 225 (1991) 213–240.
[1] G. Shock, A. Miquel, Mass transfer and pressure loss in spiral wound [13] A. Luton, S. Ragab, D. Telionis, Interaction of spanwise vortices
modules, Desalination 64 (1987) 339–352. with a boundary layer, Phys. Fluids 7 (1995) 2757–2765.

You might also like