Anda di halaman 1dari 22

Journal of Membrane Science 156 (1999) 109130

Flux improvement by Dean vortices: ultraltration of


colloidal suspensions and macromolecular solutions
P. Moulin*, P. Manno, J.C. Rouch, C. Serra, M.J. Clifton, P. Aptel
Laboratoire de Genie Chimique (CNRS-UPS-INP UMR 5503), Universite Paul Sabatier,
118 route de Narbonne, 31062 Toulouse Cedex, France
Received 26 March 1998; received in revised form 16 June 1998; accepted 26 October 1998

Abstract
Coiled and straight hollow-bre modules have been built and tested; the permeate ux obtained in ultraltration with these
two geometries is compared for two feeds: a colloidal bentonite suspension and a dextran solution. In the case of colloidal
suspensions, the secondary ows induced by the coiled geometry allow fouling to be reduced and the permeate ux is
multiplied by a factor of up to 2. An empirical relationship is proposed to express the limiting ux of permeate as a function of
both the velocity and some geometrical parameters of the coiled modules. Analogous results are obtained during the
ultraltration of dextran. It is also shown that under certain conditions almost no deposit was formed; the permeate ux under
these conditions is three times higher for coiled modules than for straight ones. For a given energy expenditure and
ultraltration process, the gain in permeate ux can reach a factor of 1.8. # 1999 Elsevier Science B.V. All rights reserved.
Keywords: Ultraltration; Hollow-bre modules; Membrane fouling; Secondary ows; Dean vortices

1. Introduction
Over recent years, much research work has been
devoted to the improvement of both the design and
operation of membrane separation modules [14]. It
has been found that Dean vortices, generated by a
coiled geometry, can be used to modify the hydrodynamics near the membrane surface. In the case of
two membrane processes using gasliquid contactors
[5,6], it was shown that the vortices improved mass
*Corresponding author. Present address: Laboratoire d'Etudes et
d'Applications de Procedes Separatifs (LEAPS), Universite d'AixMarseille, Avenue Escadrille Normandie Niemen, 13397 Marseille
Cedex 20, France. Tel.: +33-491-027776; fax: +33-491-288600;
e-mail: philippe.moulin@leaps.u-3mrs.fr

transfer by decreasing the thickness of the polarisation


layer. So it seemed interesting to investigate the effect
of Dean vortices on other systems limited by the mass
transfer: fouling by a lter cake or macromolecular
concentration polarisation. Four recent studies show
the interest of such a geometry [710]. Mallubhotla et
al. [7] and Belfort [8] have studied the effect of Dean
vortices on membrane ltration in a curved-channel
membrane system. Ophoff et al. [9] proposed new
helically twisted tubular membrane modules for membrane ltration. Manno et al. [10] studied the use of
Dean vortices for ultraltration of yeast suspensions.
The rst aim here is to observe the inuence of
Dean vortices on mass transfer resistance due to cake
formation with bentonite and concentration polarisation with dextran. To facilitate analysis of the results,

0376-7388/99/$ see front matter # 1999 Elsevier Science B.V. All rights reserved.
PII: S0376-7388(98)00333-0

110

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

the membranes used were hollow-bre ultraltration


membranes designed for water treatment that have a
cut-off that allows complete retention for the selected
species.
The second aim is to determine to what extent Dean
vortices could be utilised industrially. This study is not
concerned with dening detailed physical mechanisms but represents the experimental results by phenomenological relationships that will be useful for a
technicaleconomical evaluation.
2. Theory
Fig. 1. Dean vortices in a coiled tube.

2.1. Variation of permeate flux


If the feed is only a pure solvent, the ux (J0)
through the membrane (permeate) is proportional to
the transmembrane pressure (Ptm) according to Darcy's law:
J0

Ptm
Lp0 Ptm :
 0 Rm

(1)

In the case of ultraltration of colloidal suspensions


[11], the curve showing the permeate ux as a function
of the transmembrane pressure is identical at low
pressure to the straight line for the solvent, then at
higher pressures it diverges from it progressively.
Beyond a certain value of transmembrane pressure,
the ux becomes independent of the pressure and
levels off at a value that is called the limiting-ux
Jlim. This ux limitation is due to the accumulation of
particles near the membrane surface. In order to
represent the additional resistance (Ra) that lowers
the permeate ux, the resistance-in-series model can
be used. Then, in the case of ultraltration, Darcy's
law becomes
J

Ptm
Lp Ptm :
0 Rm Ra

(2)

In the case of laminar ow of a uid through a


cylindrical tube, the axis of which forms a helix of
small pitch, the primary (axial) ow eld is accompanied by a secondary ow eld which acts in a plane
perpendicular to the tube axis and which is symmetrical to the plane of curvature of the tube (Fig. 1).
Dean showed that the parameter De, now known as
the Dean number, is the dynamic parameter governing
uid motion in such a curved tube:
r
di
De Re
:
(3)
dc
In a helical tube, a modied Dean number can be
used to take into account the torsion effect:
r
di
:
(4)
De0 Re
dc 0
Here dc 0 is the effective coil diameter which varies
with the pitch of the coil b:
"
 2 #
b
0
:
(5)
dc dc 1
dc

2.2. Dean vortices

2.3. Pressure drops

In ow through a curved pipe, secondary ows


appear that are induced by the centrifugal force. Dean
[12,13] rst studied the complex ow pattern in a
curved pipe using a concentric toroidal co-ordinate
system.

During the ultraltration of distilled water with the


different modules, the experimental pressure drops
were compared with the theoretical values. In the case
of a straight module in the laminar regime, the pressure drops inside the bres P are given by the

0.57
0.48
0.24

0.57
0.48
0.24

2S9
2C9
2c9

2S7
2C7
2c7

0.47
0.47
0.23

0.47
0.47
0.23

Active
length l (m)

20
20
5

9
10
3

Number of
fibres n

0.70
0.70
0.70

0.93
0.93
0.93

Int. diam.
di (mm)

2.05
2.05
0.25

1.22
1.36
0.21

Surface area
(m2)100

1
1

1
1

No. of
coil

11
4.1

11
4.1

Coil diameter
dc (cm)

1
10
6

1
10
6

Pitch b
(mm)

6.1
5.8
4.1

6.1
6.2
5.3

Permeability Lp0 1010


(m Pa1 s1) (208C)

Numbering: the first number indicates the module series; S: straight; C: coiled (11 cm diameter); c: coiled (4.1 cm diameter); the second number indicates the hollow-fibre
diameter: 7 for 0.70 mm and 9 for 0.93 mm.

Total length
ltot 0 (m)

Name

Table 1
Characteristics of ultrafiltration hollow-fibre modules

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130


111

112

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

HagenPoiseuille law. The friction factor f in a tube is


dened as
f

di P
:
2v2 lT

(6)

In the case of coiled modules, the pressure drop can


be calculated according to Eq. (7), a correlation from
Mishra and Gupta [14] giving the ratio between the
friction factor in a coiled tube fc and in a similar
straight tube fs:
fc
1 0:033log10 De0 4 :
fs

(7)

This correlation is valid for the range of values:


0:003 < di =dc < 0:15;

0 < b=dc < 25:4;

1 < De < 3000:


The characteristics of the modules used (Table 1)
are consistent with this range of validity. From
Eqs. (6) and (7), it is possible to deduce the pressure
drop in a coiled hollow-bre bundle.
2.4. Zero-deposit critical point
The critical point for zero deposit [11,15,16] can be
dened as the point of separation between the ltration
curve for pure solvent and the curve for a suspension.
The gures giving the variation of permeate ux as a
function of the transmembrane pressure, for various
velocities, allow the the co-ordinates of these points to
be estimated approximately and show that at a given

velocity, this point of no deposit varies with the


module geometry.
By combining Eqs. (1) and (2), Eq. (8) is obtained.
It allows the additional resistance due to the deposit of
particles [17] to be calculated as a function of the
transmembrane pressure


Ptm 1 1
Ra

:
(8)
 0 J J0
The ux at the zero-deposit critical point is obtained
by plotting Ra vs. Ptm and extrapolating the curve to
the abscissa.
3. Experimental
3.1. Ultrafiltration of a colloidal suspension of
bentonite
The experimental apparatus is represented in Fig. 2.
The suspension (4 l) contained in a tank (A) is maintained under agitation (I) and at constant temperature.
The colloidal suspension used in this study is a
bentonite suspension whose the characteristics are
given in Table 2. A feed screw pump (PCM
P2MGI, Moineau, Vanves, France) (B) ensures the
circulation and a by-pass (G) allows the feed suspension to be pumped at low velocity. The permeate and
the retentate leaving the module (D) are recycled back
to the tank so that the composition in the tank can be
assumed constant if the quantity of matter retained on
the membrane is considered as invariant. The pressure

Fig. 2. Schematic diagram of experimental apparatus: tank (A), screw pump (B), flowmeter (C, C0 ), module (D), pressure indicators (E, E0 ),
retentate valve (F), by-pass (G).

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130


Table 2
Characteristics of bentonite suspension
Characteristics

Values

Manufacturer
Bentonite reference
Concentration
Solvent
Size
pH
Zeta potential
Conductivity

Ceca (Paris, France)


Clarsol FB2
3.27 g kg1
Distilled water
430440 nm
10.2
44.8 mV (pH10)
234 mS cm1, distilled water
(3.1 mS cm1)
1770 NTU
1.107106 m2 s1,
water1.005106 m2 s1

Turbidity
Kinematic viscosity at 208C

setting is obtained by progressively closing the valve


(F). The value of the transmembrane pressure is taken
equal to the average of the pressures indicated (TB
233, Sedeme Bourdon, Les Ulis, France) downstream
(E0 ) and upstream (E) of the modules, the pressure
drop is the difference between these two values. The
ow rates are measured by electromagnetic owmeters (C) Deltaux GK2,258 (Khrone, Epinay-surSeine, France) for the feed and Picomag DN2 (Endress
& Hauser, Eysines, France) for the low permeate ow
rate. Both these owmeters have 0.5% precision.

113

The ultraltration modules were built using cellulose-acetate hollow bres (BCDA 48 and BCDA 68,
Aquasource, Toulouse, France) having inner diameters of 0.93 and 0.70 mm, respectively. Two module geometries were used: straight and coiled. The
hollow bres are placed in a exible PVC tube which
is then wound around a rigid cylindrical tube so as to
form a coil with touching pitch. The bres are potted
in an epoxy plug at each end of the modules.
The characteristics of the modules are shown in
Table 1. The effective bre length is taken as the total
length minus the length of the resin plug. As the feed
ow is inside the bres, the hydraulic diameter used to
calculate the characteristic parameters (Reynolds
number, pressure drop) is the inner bre diameter.
Before beginning ultraltration, the reference state
of the membrane is measured by the pure water
permeability coefcient Lp0 at 208C; it can be estimated with an uncertainty of 5%.
The ultraltration of a bentonite suspension is carried
out at both constant temperature and constant cross-ow
velocity.Foreachvelocity,thepermeateuxismeasured
as a function of the transmembrane pressure. The pressure is initially set at a low value, then is increased by
successive stages until a limiting ux is reached (Fig. 3).
These values of limiting ux were corrected to 208C

Fig. 3. Variation of permeate flux and transmembrane pressure with time (module 2C9, bentonite suspension, C03.27 g kg1, di0.93 mm,
v1.70 m s1, Re1428, De0 131, T208C).

114

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

taking into account thevariation inviscosity of water as a


function of the temperature.
When a steady limiting ux is obtained, the experiment is stopped. A backush with distilled water is
performed to remove the lter cake, until the permeability coefcient for water returns to the initial value.
The permeate ux is expressed in l h1 m2, as this
unit is used industrially; for some curves used to
establish correlations, the SI system is respected
(m s1). In some gures, error bars are shown: these
represent estimates of uncertainty in the experimental
measurements [23].
3.2. Ultrafiltration of a dextran solution
The experimental apparatus and the modules used
are identical to those used for the ultraltration of the
bentonite suspension. The macromolecular solute is
dextran DT 2000; the characteristics of the feed
solution are given in Table 3.
In the ultraltration of macromolecules, fouling is
caused by adsorption of dextran at the membrane
surface or inside the membrane pores. This adsorption, which depends on the nature of both the uid and
the membrane, causes a reduction in the permeability
coefcient [18,19].
In order to use for each experiment a module having
the same reference state (i.e. the same permeability
coefcient for water), pre-adsorption operations were
performed under more extreme conditions of concentration polarisation than those applied during the later
experiments. The operating conditions for this preadsorption are: a concentration of 2 g kg1, negligible
cross-ow velocity, a transmembrane pressure of
about 200 kPa. These operations are continued until
a constant permeability coefcient is obtained after a
backush with water.
Table 3
Characteristics of dextran solution DT 2000
Characteristics

Values

Dextran
Manufacturer
Concentration
pH
Conductivity
Kinematic viscosity (T208C)

DT 2000
Pharmacia
1 g kg1
5.6
4.1 mS cm1
1.071106 m2 s1,
water1.005106 m2 s1

Table 4
Variation of permeability coefficient before and during the
ultrafiltration of dextran solution
Module

1010Lp0
(m Pa1 s1)

1010Lp
(m Pa1 s1)

2S9
2C9
2c9
2S7
2c7
2C7

6.1
6.2
5.3
6.1
4.1
5.8

2.9
3.6
3.5
3.5
3.1
3.5

Lp0 and Lp are the membrane permeability before and after dextran
adsorption, respectively.

For example, Fig. 4 shows the variation of the


permeability coefcient of the membranes in module
2C9 before, during and after the pre-adsorption operations. Table 4 gives the permeability coefcients of
modules used for dextran ultraltration before and
after pre-adsorption. These permeabilities are lower
than those for the bentonite experiments, so the limiting ux will be more difcult to reach in the operating
range of transmembrane pressure (0200 kPa).
4. Results for ultrafiltration of a colloidal
bentonite suspension
4.1. Pressure drops
Fittings are mounted at each end of the module to
allow it to be connected to the ow circuit: the
pressure drops introduced by these ttings (expansions and contractions) are less than 3 kPa and so are
negligible.
The measurements obtained with a coiled module
(2C7) are represented in Fig. 5. In this gure, the
calculated pressure drop for a straight module of same
length is also plotted to illustrate the increase in the
pressure drop due to the secondary ow. The comparison between the theoretical and experimental values
is satisfactory.
4.2. Influence of the velocity
Figs. 6 and 7 show both the variation of permeate
ux as a function of the transmembrane pressure for
different values of the cross-ow velocity for the

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

115

Fig. 4. Variation of permeability coefficient before, during and after pre-adsorption of dextran (module 2C9, dextran solution, T208C).

Fig. 5. Comparison of predicted pressure drop with experimental values (module 2C7, water, di0.70 mm, dc11 cm, Re660v, De0 58v,
T208C).

straight module 2S9 on the one hand and for the coiled
module 2c7 on the other. Beyond a certain value of
cross-ow velocity, the values of the limiting ux
obtained in a coiled module are always higher than
those obtained in the straight one. The higher the ow
velocity, the greater is the improvement in mass
transfer in the presence of Dean vortices. The higher

the velocity, the higher is the transmembrane pressure


necessary to reach the critical point of zero deposit.
4.3. Straight modules
As in the case of water oxygenation [5], the results
obtained with the straight modules were compared to

116

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 6. Effect of the velocity on permeate flux (module 2S9, bentonite suspension, C03.27 g kg1, di0.93 mm, Re840v, T208C).

Fig. 7. Effect of the velocity on permeate flux (module 2c7, bentonite suspension, C03.27 g kg1, di0.70 mm, Re660v, De0 99v,
T208C).

reference studies [2022]. These studies showed that


the limiting permeate ux depended on the cross-ow
velocity in the following way:
Jlim bva :

(9)

The units used in this relationship are those given in


the list of symbols; the same approach has been
followed here for all the correlations for which an
expression in terms of dimensionless groups has not
been possible.

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

117

Table 5
In a straight module, relation between the limiting flux of permeate and the velocity (bentonite suspension)
Reference
[20]
[21]
[21]a
[22]
Present study

Suspensions
Bentonite
Bentonite
Bentonite
Bentonite
Bentonite

[0.033 g l ]
[1 g l1]
[3 g l1]
[1.431035103 g l1]
[3.27 g l1]

Configuration

Regime

Hollow fibres, di0.98 mm


Hollow fibres, di0.93 mm
Hollow fibres, di0.93 mm
Ceramic membrane, Di7.0 mm
Hollow fibres, di0.93 mm

Laminar
Laminar
Laminar
Laminar
Laminar

a
(Re<1600)
(Re<600)
(Re<200)
(Re<2100)

0.73
0.60
0.80
0.55
0.6

As Wandelt's work is not aimed at determining the limiting flux as a function of velocity, the two exponents were obtained from a small
number of measurements. This could explain the higher value of the exponent.

Table 5 summarises the values of a found by these


authors. For the straight modules 2S9 and 2S7, the
values of a and b are given in Table 6. The values of a
are approximately consistent with the literature, and in
our case are identical for the two internal diameters
used.
Another way of representing the results is to express
the variation of Jlim as a function of the shear stress .
The wall shear stress for a Newtonian uid owing
through a straight cylindrical tube in the laminar
regime is
 8v=di :

(10)

Fig. 8 shows clearly that for a given shear stress the


limiting ux in straight modules does not vary with the
internal diameter of the bre. The following relationship is found:
Jlim 6:9  106  0:6 ;

R2 0:998:

(11)

values of the cross-ow velocity and the internal


diameter of the hollow bres. The limiting permeate
ux is greater in the presence of secondary ow. This
ux increases with the Dean number (for a small coil
diameter).
Figs. 11 and 12 show the limiting ux plotted as a
function of the cross-ow velocity for hollow bres
with an internal diameter of 0.93 and 0.70 mm. The
higher the cross-ow velocity, the greater is the effect
of the secondary ow. In fact if the limiting ux is
expressed as a function of the velocity (Jlimbva) for
each module, the result is given in Table 6.
If we try to correlate the results as a function of the
Dean number (or other dimensionless groups), the
curves obtained are completely independent of each
other. On the other hand, if we express the limiting ux
as a function of the ratio v/di and of the coil diameter
with a Reynolds number in the range 1701600, the
relationship is
Jlim 3:9  108 v=di 0:9 dc0:25 ;

4.4. Coiled modules


Figs. 9 and 10 give some examples of the variation
of permeate ux as a function of the transmembrane
pressure for different module geometries for given
Table 6
For straight and coiled modules, a and b coefficients for the
relationship: Jlimbva
Module

b105

R2

2S9
2C9
2c9
2S7
2c7
2C7

2.71
3.93
4.62
3.24
5.85
4.29

0.60
0.85
0.96
0.60
0.88
0.84

0.998
0.998
0.998
0.997
0.982
0.990

R2 0:98:

(12)

Fig. 13 illustrates the correlation between the


experimental results and the values calculated from
Eq. (12). This shows that the smaller the coil diameter,
the greater is the permeate ux. The 0.9 exponent
compared with the 0.6 exponent obtained with straight
modules shows that the effect of the secondary ow is
greatest at high cross-ow velocities.
The zero-deposit critical point (Fig. 14) is located
at higher values of transmembrane pressure and
permeate ux when Dean vortices are present. The
position of this point also varies with the coil diameter:
the smaller the coil diameter, the higher is the transmembrane pressure at that point. This means that the
coiled modules can work under zero-deposit conditions with permeate uxes higher than for straight
modules. For example, at v2.1 m s1, for a Reynolds

118

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 8. In laminar regime and for straight modules, limiting flux of permeate vs. shear stress  (module 2S9 and 2S7, bentonite suspension,
C03.27 g kg1, T208C).

Fig. 9. Permeate flux vs. transmembrane pressure at a given velocity for different geometry of modules (bentonite suspension,
C03.27 g kg1, di0.93 mm, v0.80 m s1, Re672, T208C).

number of 1400, and over the whole range of transmembrane pressure, the 2c7 module gives a permeate
ux for the ultraltration of a bentonite suspension

(Fig. 7) that is equal to the ltration ux for water. The


highest value reached is almost 220 l h1 m2. For the
equivalent straight module 2S7 the zero-deposit cri-

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

119

Fig. 10. Permeate flux vs. transmembrane pressure at a given velocity for different geometry of modules (bentonite suspension,
C03.27 g kg1, di0.70 mm, v0.80 m s1, Re525, T208C).

Fig. 11. Limiting permeate flux vs. velocity for different module geometries (bentonite suspension, C03.27 g kg1, di0.93 mm, Re840v,
T208C).

tical point is reached at a lower permeate ux, around


150 l h1 m2.
These results can be presented in a different way: if
the ratio between the limiting ux obtained with and
without secondary ows is expressed as a function of
the Dean number, the following relationships are

obtained:
J 2C9=J 2S9 0:5De00:25 ;

R2 0:996;
(13a)

J 2c9=J 2S9 0:3De0

0:36

R2 0:996;

(13b)

120

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 12. Limiting permeate flux vs. velocity for different module geometries (bentonite suspension, C03.27 g kg1, di0.70 mm, Re660v,
T208C).

Fig. 13. For coiled modules, experimental values of limiting flux of permeate vs. values calculated by Eq. (12) (bentonite suspension,
C03.27 g kg1, T208C).

J 2C7=J 2S7 0:6De0


J 2c7=J 2S7 0:5De

0:28

0 0:24

R2 0:979;

4.5. Large modules


(13c)

R 0:987:

(13d)

The exponents (0.240.36) are close to those


obtained in heat transfer [29] and during the microltration of yeast [7].

To test the possibility of building bigger modules, to


see whether such modules perform like the laboratory
modules and to show the feasibility of using Dean
vortices on the industrial scale, larger modules were
built, both straight and coiled. The measurements with

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

121

Fig. 14. Variation with the transmembrane pressure of additional resistance due to the filter cake for different module geometries (bentonite
suspension, C03.27 g kg1, di0.93 mm, v0.80 m s1, Re672, T208C).

Table 7
Characteristics of large hollow-fibre modules
Characteristics

Large straight
module

Large coiled
module

Effective length (m)


dide (mm)
Number of fibres
Surface area (m2)
Lp0 (m Pa1 s1)

0.85
0.701.26
450
0.84
4.71010

0.85
0.701.26
436
0.82
4.41010

these modules were performed in the same way as for


the laboratory modules, with the tank volume
increased from 4 to 17 l. Table 7 shows the geometrical parameters of the two modules used for these
experiments.
If the limiting permeate ux is expressed as a
function of the cross-ow velocity using Eq. (9),
the value of a is 0.6 and b is 2.9105, for straight
modules. This should be compared with Table 6 for
laboratory modules. The limiting-ux measurements
with the straight module conrm the 0.6 exponent
relating the limiting permeate ux to the cross-ow
velocity.
If the permeate ux for the coiled and straight
geometries is compared (Fig. 15) as a function of

the transmembrane pressure at a constant velocity,


the permeate ux is higher with secondary ow. The
improvement at this velocity of 0.35 m s1 is about
50%, in agreement with the data obtained at the same
velocity with the laboratory modules.
Fig. 16 gives the experimental limiting ux and the
limiting ux calculated from Eq. (12). It can be seen
that the experimental values are slightly higher than
the calculated values.
5. Results for ultrafiltration of a dextran
solution
5.1. Straight modules
Table 8 gives the values of a and b corresponding to
the limiting ux for the two straight modules, together
with those previously obtained in our laboratory: a
good agreement can be noted. The limiting ux for the
straight modules as a function of the shear stress  ts
the following relationship:
Jlim 7:2  106  0:40 ;

R2 0:986:

(14)

122

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 15. For large modules, permeate flux vs. transmembrane pressure (bentonite suspension, C03.27 g kg1, di0.70 mm, v0.35 m s1,
Re231, T208C).

Fig. 16. Comparison of experimental limiting permeate flux with calculated values (Eq. (12)) (bentonite suspension, C03.27 g kg1,
di0.70 mm, v0.35 m s1, Re231, T208C).

5.2. Influence of secondary flows


Fig. 17 shows the variation of permeate ux as a
function of the transmembrane pressure for different

module geometries, at a xed cross-ow velocity.


The permeate uxes are higher in the presence of
secondary ows and increase as the coil diameter is
reduced.

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130


Table 8
In a straight module, relationship between the limiting permeate
flux and the velocity (dextran solution)
Modules

Relations

Reference

Dextran

Straight
Straight
2S9
2S7

Jlim2.3105v0.4
Jlim1.6105v0.4
Jlim1.7105v0.4
Jlim1.9105v0.4

[27]
[28]
This study
This study

DT
DT
DT
DT

70
70
2000
2000

Figs. 18 and 19 show the limiting ux as a function


of the cross-ow velocity for different module geometries and for different hollow-bre diameters. The
limiting ux is higher with coiled modules than with
straight: it is multiplied by 1.5 for a cross-ow velocity
of about 1.0 m s1 (Fig. 18). At low cross-ow velocity, the smaller the coil diameter, the greater is the
limiting permeate ux. However, the difference
between the two geometries (at given internal bre
diameter) decreases when the velocity rises. The
relationship between the limiting permeate ux and
the cross-ow velocity for modules 2C9 and 2C7 is
5

0:5

Jlim 2:5  10 v=di :

(15)

For modules 2c9 and 2c7 (dc4.1 cm) this correlation does not t as well. This is probably due to the

123

greater experimental error involved in measuring


small permeate ow rates.
Fig. 20 shows that for all transmembrane pressures
the resistance due to concentration polarisation is
lower in the presence of secondary ows. This resistance decreases with decreasing coil diameter. An
absence of additional resistance indicates the apparent
absence of concentration polarisation and the ux
beyond which this resistance appears corresponds to
the critical ux for zero concentration polarisation.
For modules 2C7 and 2c7, this value is reached at 45
and 50 kPa, respectively. On the other hand, in the case
of a straight module, this critical value is much lower,
about 10 kPa.
If the results are given in terms of the ratio
of limiting ux for coiled and straight modules
the following relationships are obtained: the
exponent of 0.1 is smaller than for the bentonite
results:
0:1

(16a)

0:1

(16b)

J 2C9=J 2S9 0:94De0 ;


J 2C7=J 2S7 0:89De0 :

This lower exponent is in agreement with the results


obtained by Mallubhotla et al. [7] for the ultraltration
of dextran solutions.

Fig. 17. Permeate flux vs. transmembrane pressure at a given velocity for different module geometries (dextran solution, C01 g kg1,
di0.70 mm, v0.40 m s1, Re294, De0 38 (2c7); 23 (2C7), T208C).

124

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 18. Limiting permeate flux vs. velocity for different module geometries (dextran solution, C01 g kg1, di0.93 mm, Re840v,
T208C).

Fig. 19. Limiting permeate flux vs. velocity for different module geometries (dextran solution, C01 g kg1, di0.70 mm, Re660v,
T208C).

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

125

Fig. 20. Variation of additional resistance due to the concentration polarisation with the transmembrane pressure and the module geometries
(dextran solution, C01 g kg1, di0.70 mm, v0.40 m s1, Re294, De38 (2c7); 23 (2C7), T208C.

6. Discussion

Table 9
Critical Dean number for various processes

6.1. Limiting flux in ultrafiltration: some different


mechanisms

Process

Critical Dean
number

Heat transfer [30]


NMR imaging [31,32]
Water oxygenation [5,23]
Pervaporation [6,23]
This study

23.4
25
25
19
1628

In ultraltration with coiled modules there is a


difference in behaviour between colloidal suspensions
and macromolecular solutions: the curves showing the
limiting ux as a function of velocity (Figs. 11, 12, 18,
19) do not have the same form. In the ultraltration of
bentonite suspensions the curves giving the ux in the
two coiled geometries diverge as the velocity
increases (Figs. 11 and 12), whereas in the ultraltration of dextran solutions the two ux curves converge
(Figs. 18 and 19). This seems to indicate that
the mechanisms responsible for mass transfer are
fundamentally different for these two different types
of feed. This could correspond to the difference
between molecular diffusion for macromolecules
and shear-induced migration for particles in suspension [15,23].
6.2. Critical Dean number
In several studies concerning the use of Dean
vortices in ultraltration [2426], analytical techni-

ques are used, together with optical and magneticresonance imaging methods to characterise Dean vortices in a curved channel. The authors use a critical
Dean number Dec to characterise that ow rate above
which the ow pattern changes and vortices are
detected. We can dene the critical Dean number in
membrane operations as the value of the Dean number
beyond which an improvement in mass transfer
occurs: Table 9 summarises the critical Dean numbers
found in various studies. There is an agreement
between all these values obtained with coiled geometries for different applications. When the Dean
number is below 20, the secondary ows do not have
a measurable effect.

126

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 21. Energy consumption vs. transmembrane pressure (module 2C9, bentonite suspension, C03.27 g kg1, di0.93 mm, v0.5 m s1,
Re420, T208C).

6.3. Energy expenditure and mass transfer


enhancement
As the pressure drops are higher in coiled than in
straight modules, it is important to consider the performance of the coiled geometry in terms of energy
expenditure. To concentrate on the effect of modifying
module geometry, only the energy dissipated within
the module itself will be considered. The overall
power dissipated within the module Pd is the sum
of two terms: the rst is the power dissipated by the
cross ow and the second is the power required for
membrane permeation. As the permeate ow rate is
small compared with the cross ow, the latter will be
taken equal to Q the ow rate at the module inlet. The
overall power expenditure within the module can then
be taken, to a good approximation, as
Pd QP Qp Ptm :

(17)

The dissipated power is expressed in W, the volumetric ow rate in m3 s1 and pressures in Pa. The
energy expenditure per unit volume of permeate (in
kW h m3) is


Pd
QP
E

Ptm
3:6  106 :
3:6  106 Qp
Qp
(18)

For a given cross-ow velocity, the plot of E as a


function of the transmembrane pressure passes
through a minimum (Fig. 21). At low pressures, the
energy decreases with increasing pressure because the
permeate ow rate increases and the rst term in
brackets on the right-hand side of Eq. (18) decreases
more quickly than the second increases. On the other
hand, when the limiting ux is reached, the rst term
stays constant and the energy expenditure increases
linearly with Ptm. Thus, a minimum appears and this
estimated value of the minimum energy will be used to
compare the straight modules to the coiled ones.
Figs. 22 and 23 give the ratio of the limiting
permeate ux for a coiled to a straight module as a
function of the energy expenditure, for the ultraltration of bentonite suspensions and dextran solutions,
respectively. It should be noted, however, that at a
given energy expenditure the Reynolds numbers in the
two modules are not the same. These two gures show
that the limiting permeate ux at a given energy
expenditure is greater in presence of Dean vortices,
with an improvement factor of 2 for the ultraltration
of colloid suspension and 1.5 for the macromolecular
solution. The curves in these two gures have quite
different shapes. No obvious explanation for this is
available: we can only suppose that this difference is
due to a difference in the mass transfer mechanisms.

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

127

Fig. 22. Ratio of limiting permeate flux between a coiled and a straight module vs. energy consumption (bentonite suspension,
C03.27 g kg1, di0.93 mm, Re840v, T208C).

Fig. 23. Ratio of limiting permeate flux between a coiled and a straight module vs. energy consumption (dextran solution, C01 g kg1,
di0.70 mm, Re660v, T208C).

128

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

Fig. 24. Variation of Sherwood number with energy consumption (water, modules 1S-1D, di3.2 mm and 2S-2D di2.4 mm).

6.4. Gasliquid contactors


In a previous work, we examined the effect of Dean
vortices in gasliquid membrane contactors [5]. The
present treatment in terms of energy expenditure can
also be extended to that application. There the power
dissipated Pd is only determined by the feed ow
rate

Pd QP:

(19)

For the case of the water oxygenation, Fig. 24


shows that for a given energy dissipation, the mass
transfer (Sh) remains higher in presence of secondary
ows. These results obtained in the case of gasliquid
membrane contactors are in agreement with the results
obtained in ultraltration processes.
7. Conclusion
This study is concerned with the inuence on
ultraltration of the secondary ows generated by a
coiled module geometry. The following points have
been established:
 In the ultrafiltration of a colloidal bentonite suspension or of a macromolecular dextran solution,





the results obtained with straight modules are in


agreement with previously published results.
Dean vortices cause an increase in permeate flux
for a given transmembrane pressure. This
improvement gets better as the coil diameter is
reduced: the permeate flux can be as much as
doubled. These results confirm observations made
in the first part of the study on Dean vortices in
membrane processes [5,6,23].
In the case of colloid ultrafiltration, an empirical
relationship allows the limiting permeate flux to
be related to both the module characteristics and to
the cross-flow velocity.
The zero-deposit critical flux is multiplied by a
factor of up to 4 when a straight module is replaced
by a coiled one.
The results obtained with large modules show that
the observations made with laboratory modules
can be transposed to the industrial scale.
The calculation of the energy dissipated to reach
the limiting flux shows that for a given energy, the
permeate flux obtained with the coiled modules is
higher than with the straight modules. The
improvement remains important and goes up to
a factor of 2 in the case of the ultrafiltration of
bentonite suspensions and 1.5 for dextran solutions.

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

129

8. Nomenclature

References

b
d
De
De0

[1] G. Belfort, Fluid mechanics in membrane filtration: recent


developments, J. Membr. Sci. 40 (1989) 123147.
[2] H.B. Winzeler, G. Belfort, Enhanced performance for
pressure-driven membrane processes: the argument for fluid
instabilities, J. Membr. Sci. 80 (1993) 3547.
[3] M. Nystrom, J.A. Howell, Flux enhancement, in: J.A. Howell,
V. Sanchez, R.W. Field (Eds.), Membranes in Bioprocessing:
Theory and Applications, Chapman & Hall/Blackie, London,
1993.
[4] G. Belfort, R.H. Davis, A.L. Zydney, The behavior of
suspensions and macromolecular solutions in crossflow
filtration, J. Membr. Sci. 96 (1994) 158.
[5] P. Moulin, C. Serra, J.C. Rouch, M.J. Clifton, P. Aptel, Mass
transfer improvement by secondary flows Dean vortices in
coiled tubular membranes, J. Membr. Sci. 114 (1996) 235
244.
[6] S. Schnabel, P. Moulin, T. Nguyen, D. Roizard, P. Aptel,
Removal of volatile organic components (VOC's) from water
by pervaporation with dense silicone hollow fibres. Separation improvement by Dean vortices, J. Membr. Sci. 142
(1998) 129141.
[7] H. Mallubhotla, S. Luque, G. Belfort, Membrane filtration
with self-cleaning Spiral vortices, in: A.J.B. Kemperman,
G.H. Koops (Eds.), Euromembrane'97, Progress in Membrane Science and Technology, Univ. of Twente, Netherlands,
1997, pp. 385387.
[8] G. Belfort, Coiled membrane filtration system, US Patent
5 626 758, 1997.
[9] J. Ophoff, G.S. Vos, I.G. Racz, T. Reith, Flux enhancement by
reduction of concentration polarisation due to secondary flow
in twisted membrane tubes, in: A.J.B. Kemperman, G.H.
Koops (Eds.), Euromembrane '97, Progress in Membrane
Science and Technology, Univ. of Twente, Netherlands, 1997,
pp. 388390.
[10] P. Manno, P. Moulin, J.C. Rouch, P. Aptel, Mass transfer
improvement in helically wound hollow-ultrafiltration modules: bentonite and yeast suspensions, in: A.J.B. Kemperman,
G.H. Koops (Eds.), Euromembrane '97, Progress in Membrane Science and Technology, Univ. of Twente, Netherlands,
1997, pp. 388390.
[11] C. Gourgues, P. Aimar, V. Sanchez, Ultrafiltration of
bentonite suspensions with hollow fiber membranes, J.
Membr. Sci. 74 (1992) 5169.
[12] W.R. Dean, The stream line motion of fluid in a curved pipe,
Phil. Mag. 5 (1928) 674695.
[13] S.A. Berger, L. Talbot, L.S. Yao, Flow in curved pipes, Ann.
Rev. Fluid Mech. 15 (1983) 461512.
[14] P. Mishra, S.N. Gupta, Momentum transfer in curved pipes, 1.
Newtonian fluids, 2. Non-Newtonian fluids, Ind. Eng. Chem.
Process Des. Dev. 18 (1979) 130142.
[15] P. Bacchin, P. Aimar, V. Sanchez, Influence of surface
interaction on mass transfer during colloid ultrafiltration,
J. Membr. Sci. 115 (1996) 4963.
[16] J.A. Howell, Sub-critical flux operation of microfiltration,
J. Membr. Sci. 107 (1995) 165171.

E
f
J
lT
Lp
P
P
Ptm
Pd
Q
Re
Ra
Rm
Sh
t
T
v

pitch of the coil (m)


diameter (m)
Dean number (dimensionless quantity)
Dean number modified by the pitch (dimensionless quantity)
energy expenditure per unit volume of
permeate (kW h m3)
friction factor in a tube (dimensionless
quantity)
flux (m s1)
total length of hollow fibres (m)
hydraulic permeability coefficient (m Pa1
s1)
pressure (Pa)
pressure drop (Pa)
transmembrane pressure (Pa)
dissipated power (W)
volumetric flow rate (m3 s1)
Reynolds numberd iv/ (dimensionless
quantity)
additional resistance (m1)
membrane resistance (m1)
Sherwood number (dimensionless quantity)
time (s)
temperature (8C)
velocity of the liquid phase (m s1)

Greek letters





dynamic viscosity (Pa s)


kinematic viscosity (m2 s1)
density (kg m1)
wall shear stress (Pa)

Subscripts
a
c
i
lim
m
p
s
tm
0

additional
coil, coiled
internal
limiting
membrane
permeate
straight
transmembrane
solvent

130

P. Moulin et al. / Journal of Membrane Science 156 (1999) 109130

[17] Y. Xu-Jiang, J. Dodds, D. Leclerc, Cake characteristics in


crossflow and dead-end microfiltration, Filtr. Separ. (1995)
795798.
[18] L.A. Errede, Effect of molecular adsorption on water
permeability of microporous membranes, J. Membr. Sci. 20
(1984) 4561.
[19] P. Aimar, V. Sanchez, Le colmatage des membranes
d'ultrafiltration et de microfiltration, Belgian J. Food Chem.
Biol. 44 (1989) 203214.
[20] H. De Balmann, V. Sanchez, Etude du couplage entre les
phenomenes de polarisation de concentration et de colmatage
en ultrafiltration, Entropie 152 (1989) 1927.
[21] B. Wandelt, P. Schmitz, D. Houi, Micro- and ultra-filtration of
dilute suspensions: NMR micro-imaging of deposit formation
in hollow fibers, Euromembrane '92, Progress in Membrane
Science and Technology, Paris, France, 1992, pp. 6, 22.
[22] T. Murase, T. Ohn, K. Kimata, Filtrate flux in crossflow
microfiltration of dilute suspension forming a highly
compressible fouling cake-layer, J. Membr. Sci. 108 (1995)
121128.
[23] P. Moulin, Amelioration du transfert de matiere dans les
operations de separations par membranes en presence de
vortex de Dean, Thesis, Universite Paul Sabatier, Toulouse,
France, 1996.
[24] M.E. Brewster, K.Y. Chung, G. Belfort, Dean vortices with
wall flux in a curved channel membrane system: 1. A new
approach to membrane module design, J. Membr. Sci. 81
(1993) 127137.

[25] K.Y. Chung, W.A. Edelstein, X. Li, Dean vortex in curved


tube flow: 5- 3-D MRI and numerical analysis of the velocity
field, AIChE J. 39 (1993) 15921602.
[26] K.Y. Chung, W.A. Edelstein, G. Belfort, Dean vortices with
wall flux in a curved channel membrane system: 6. Two
dimensional resonance imaging of the velocity field in a
curved impermeable slit, J. Membr. Sci. 81 (1993) 151162.
[27] J.P. Lafaille, V. Sanchez, J. Mahenc, Transfert de matiere dans
les modules d'ultrafiltration equipes de fibres creuses,
Entropie 118 (1984) 2936.
[28] N. Abidine, Polarisation de concentration en ultrafiltration,
Thesis, Universite Paul Sabatier, Toulouse, France, 1985.
[29] F.C. Hass, A.H. Nissan, Proceedings of The International Heat
Transfer Conference, Boulder, CO, USA, vol. 3, 1961, p. 643.
[30] C. Castelain, Etude experimentale de la dynamique des
fluides et des transferts thermiques dans un ecoulement de
Dean alterne en regime d'advection chaotique, Thesis, Ecole
Doctorale Sciences pour l'Ingenieur de Nantes, Nantes,
France, 1995.
[31] H. Mallubhotla, G. Belfort, Aggregation during microfiltration of colloid suspensions, in: Proceedings of the Workshop
on Colloid Science in Membrane Engineering, May 1996,
Toulouse, France.
[32] H. Mallubhotla, G. Belfort, Dean vortices in curved tube
flow: magnetic resonance imaging and microfiltration,
Proceedings of The Eighth Annual Meeting of the North
American Membrane Society (NAMS), Ottawa, Canada, 18
22 May 1996.

Anda mungkin juga menyukai